Next Article in Journal
Dual GSK-3β/HDAC Inhibitors Enhance the Efficacy of Macrophages to Control Mycobacterium tuberculosis Infection
Previous Article in Journal
Prediction of Skin Color Using Forensic DNA Phenotyping in Asian Populations: A Focus on Thailand
Previous Article in Special Issue
60 Years of Studies into the Initiation of Chromosome Replication in Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Post-Translational Modifications in Necroptosis

1
Department of Liver Surgery and Transplantation, Zhongshan Hospital, Fudan University, Shanghai 200032, China
2
Liver Cancer Institute, Zhongshan Hospital, Fudan University, Shanghai 200032, China
3
Key Laboratory of Carcinogenesis and Cancer Invasion, Fudan University, Ministry of Education, Shanghai 200032, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2025, 15(4), 549; https://doi.org/10.3390/biom15040549
Submission received: 28 January 2025 / Revised: 5 March 2025 / Accepted: 6 March 2025 / Published: 9 April 2025
(This article belongs to the Collection Molecular Biology: Feature Papers)

Abstract

:
Necroptosis, a distinct form of regulated necrosis implicated in various human pathologies, is orchestrated through sophisticated signaling pathways. During this process, cells undergoing necroptosis exhibit characteristic necrotic morphology and provoke substantial inflammatory responses. Post-translational modifications (PTMs)—chemical alterations occurring after protein synthesis that critically regulate protein functionality—constitute essential regulatory components within these complex signaling cascades. This intricate crosstalk between necroptotic pathways and PTM networks presents promising therapeutic opportunities. Our comprehensive review systematically analyzes the molecular mechanisms underlying necroptosis, with particular emphasis on the regulatory roles of PTMs in signal transduction. Through systematic evaluation of key modifications including ubiquitination, phosphorylation, glycosylation, methylation, acetylation, disulfide bond formation, caspase cleavage, nitrosylation, and SUMOylation, we examine potential therapeutic applications targeting necroptosis in disease pathogenesis. Furthermore, we synthesize current pharmacological strategies for manipulating PTM-regulated necroptosis, offering novel perspectives on clinical target development and therapeutic intervention.

1. Background

Cell death holds unique physiological and pathological implications. The maintenance of organismal homeostasis depends on the precise regulation of cell death, proliferation, and differentiation [1]. Based on their regulatory mechanisms, cell death can be classified into two major categories: regulated cell death (RCD) and accidental cell death (ACD) [2,3,4,5]. RCD encompasses multiple forms, including apoptosis, necroptosis, autophagy, ferroptosis, cuproptosis, pyroptosis, and NETosis, while ACD primarily refers to necrosis. Notably, necroptosis represents a unique form of RCD that shares morphological features with necrosis, characterized by plasma membrane rupture and the subsequent release of intracellular components [4].
Extensive research has established significant associations between necroptosis and various pathological conditions, including tumorigenesis [6], inflammation disorders [7], neurodegenerative diseases [8], cerebrovascular accidents [9], myocardial infarction [10], aortic aneurysm [11], and respiratory diseases [12]. This caspase-independent form of regulated cell death is governed by intricate signaling pathways, with RIPK1 (Receptor-interacting protein kinase 1), RIPK3 (Receptor-interacting protein kinase 3), and MLKL (Mixed lineage kinase domain-like protein) serving as core regulatory components. The initiation of necroptosis is triggered through specific receptors, notably death receptors (DRs) and pattern recognition receptors (PRRs), along with their associated signaling molecules. Under conditions of caspase-8 inhibition, RIPK1 undergoes autophosphorylation and subsequently recruits and activates RIPK3 through phosphorylation. This interaction leads to the formation of the ripoptosome complex. The assembly progresses with the recruitment and phosphorylation of MLKL, resulting in necrosome formation. Ultimately, MLKL oligomerization facilitates membrane translocation, culminating in cellular membrane disruption and cell death [13,14].
In such a sophisticated signaling cascade, post-translational modifications (PTMs) profoundly influence the progression of necroptosis. PTMs represent essential biochemical processes that modulate protein function through structural alterations, including proteolytic cleavage and side-chain modifications, ultimately affecting protein conformation and activity [15]. While phosphorylation and ubiquitination have been most extensively studied in necroptotic regulation, emerging evidence highlights the involvement of additional modifications. These include methylation, nitrosylation, glycosylation, acetylation, caspase-mediated cleavage, disulfide bond formation, and SUMOylation, collectively orchestrating the necroptotic process.
A comprehensive understanding of necroptosis and its associated post-translational modifications (PTMs) is essential for developing targeted therapeutic approaches. Deciphering the intricate interplay between necroptotic pathways and PTM networks may reveal novel therapeutic targets and intervention strategies, particularly for diseases that currently lack effective treatments. This review systematically examines the molecular mechanisms underlying necroptosis, followed by a detailed analysis of the specific PTMs involved in its regulation. Furthermore, we address current controversies and knowledge gaps in the field, including emerging molecular mechanisms, PTM-related therapeutic targets, and future research directions. Through this comprehensive analysis, we aim to provide both a current perspective and a roadmap for future investigations in this rapidly evolving field.

2. Current Knowledge of Necroptosis

Necroptosis represents a distinct form of programmed cell death characterized by unique mechanisms, pathological consequences, and clinical significance. Unlike immunologically silent apoptosis, necroptosis triggers robust inflammatory responses through plasma membrane rupture and the release of damage-associated molecular patterns (DAMPs) [2]. This inflammatory cascade exceeds even that of pyroptosis, which releases IL-1β and IL-18 through Gasdermin pores [16]. The impact of necroptosis extends beyond the single-cell precision of apoptosis, often causing spreading tissue damage as seen in ischemia–reperfusion injury. This contrasts with ferroptosis, which primarily affects lipid peroxidation-sensitive cells such as renal tubular epithelia [17]. Temporally, necroptosis is completed within 4–6 h following tumor necrosis factor α (TNFα) stimulation combined with cIAP inhibitors or Z-VAD-FMK, positioning it between rapid pyroptosis (under 30 min) and slower ferroptosis (24–48 h) [17,18,19,20].
Molecularly, necroptosis depends on RIPK1 autophosphorylation and necrosome formation with RIPK3, ultimately activating MLKL to form membrane pores [21]. This pathway differs fundamentally from the inflammasome–caspase–Gasdermin axis in pyroptosis and the system Xc/GSH/GPX4 pathway in ferroptosis [16,22]. Morphologically, necroptotic cells exhibit swelling and organelle damage alongside partial chromatin condensation, without the cellular shrinkage or blebbing seen in apoptosis [19].
These regulated cell death mechanisms share upstream signals such as TNF receptor family activation while serving as important therapeutic targets. The critical role of necroptosis in stroke, myocardial infarction, inflammatory bowel disease, and viral infections has made it a focus of modern medical research [18].
The initiation of necroptosis is mediated by diverse upstream signals, such as death receptors (including TNFR1, FAS (also known as CD95 or APO-1), DR3 (also known as TRAMP or APO-3), TRAILR1 (also known as DR4), TRAILR2 (also known as DR5, TRICK, or KILLER), and DR6) [23]; pattern recognition receptors (including Toll-like receptors (TLR4 and TLR3)) [24]; reactive oxygen species (ROS); Z-DNA binding protein 1 (ZBP1); and cytosolic nucleic acid sensors such as RIG-I and STING, which together form the upstream signaling of the necroptosis pathway. These signals converge to activate a conserved molecular cascade: RIPK1 autophosphorylation, RHIM-mediated RIPK1–RIPK3 interaction, and subsequent RIPK3 oligomerization, which phosphorylates MLKL. Phosphorylation plays two pivotal roles in MLKL activation: (i) facilitating conformational transitions and (ii) mediating membrane translocation. Phosphorylated MLKL undergoes a structural shift from an autoinhibited state to an active conformation, forming dimers via its pseudokinase domain. This transition releases the brace region into an elongated helix, which subsequently oligomerizes into trimers or tetramers through coiled-coil assembly, exposing the 4-helical bundle domain (4HBD) [25,26]. Following tetramerization, MLKL dissociates from the necrosome and interacts with molecular chaperones (e.g., Hsp90 and Hsp70), facilitating its translocation to the plasma membrane. The exposed 4HB domain then binds to membrane phosphatidylinositol phosphates (PIPs), compromising membrane integrity and triggering DAMP release [27,28,29,30]. Additionally, MLKL pores may permit ion influx (e.g., Ca²⁺, Na⁺), disrupting osmotic balance and contributing to cellular swelling and rupture [28,31].
Beyond plasma membrane targeting, MLKL interacts with mitochondrial proteins to amplify necroptotic signaling. At mitochondria, MLKL binds to phosphoglycerate mutase family member 5 (PGAM5), which stabilizes MLKL through dephosphorylation, enhancing its oligomerization and membrane-disrupting activity [32,33,34]. Concurrently, MLKL can also interact with dynamin-related protein 1 (DRP1), a downstream molecule of PGAM5, inducing mitochondrial fission that releases more ROS to exacerbate necroptosis [34]. Additionally, mitochondrial DNA (mtDNA) can act as a DAMP to further amplify inflammatory responses by activating the cGAS-STING pathway [35,36]. Additionally, ATP depletion resulting from mitochondrial dysfunction impairs membrane repair mechanisms, enhancing pore formation efficiency [37]. These coordinated interactions significantly accelerate necroptosis progression.
The following sections will systematically examine these upstream signals and their downstream effectors, providing a comprehensive understanding of necroptosis induction mechanisms, including the molecular details of MLKL activation, oligomerization, and membrane targeting.

2.1. TNFR1

TNF-TNFR1 signaling represents the most extensively characterized pathway in necroptosis research. As a member of the TNF receptor superfamily (TNFRSF), TNFR1 contains a cytoplasmic death domain (DD) that orchestrates inflammatory responses, apoptosis, and necroptosis [38,39,40]. Ligand binding initiates the recruitment of TNFRSF1A associated via death domain (TRADD) and RIPK1 through the DD, followed by the assembly of TNF receptor associated factor 2 (TRAF2) and the E3 ubiquitin ligases cellular inhibitor of apoptosis 1 and 2 (cIAP1/2). These components facilitate the attachment of K11-, K48-, and K63-linked polyubiquitin chains to RIPK1 and other signaling proteins, while also promoting the engagement of the linear ubiquitin chain assembly complex (LUBAC, consisting of heme-oxidized IRP2 ubiquitin ligase-1, HOIL-1L; HOIL-1-interacting protein, HOIP; and SHANK-associated RH domain-interacting protein, SHARPIN) [41,42]. TNFR1, TRADD, and RIPK1 are conjugated to the M1-linked polyubiquitin chain mediated via LUBAC, which facilitates the formation of the IkB kinase (IKK) complex consisting of IKKα, IKKβ, and NF-κB essential modulator (NEMO) and the TGF activating kinase 1 (TAK1)-TAK1 Binding Protein 2/3 (TAB2/3) Complex [43,44,45,46,47]. Activation of TAK1 and IKK induces the activation of NF-κB and mitogen-activated protein kinase (MAPK) [48]. These are the major components of the so-called Complex I (Figure 1A). Complex I favors pro-inflammatory gene expression and inhibits cell death. Other kinases in Complex I, such as TANK binding kinase 1 (TBK1), MAPK activated protein kinase 2 (MAPKAPK2, MK2), the tyrosine kinases JAK1 and SRC, ULK1, and adenylate-activated protein kinase (AMPK), can negatively regulate protein activity by phosphorylating RIPK1 [49,50,51,52,53,54,55].
The progression from Complex I to Complex IIa/b depends on multiple critical conditions. These include enhanced RIPK1 autophosphorylation or kinase activity, suppressed RIPK1 ubiquitination, inhibited RIPK1 phosphorylation, altered post-translational modifications (glycosylation, caspase cleavage, SUMOylation), and Complex I destabilization. Various molecular interventions disrupt this transition, such as IAP inhibitors, LUBAC inhibitors, RIPK1 autophosphorylation, RIPK1-DD-mediated dimerization, phosphorylase inhibitors, and kinase knockout [56]. Conversely, cylindromatosis (CYLD) facilitates this progression between complexes [57]. CYLD is linked to LUBAC via the adaptor protein spermatogenesis associated 2 (SPATA2), and it is able to eliminate M1- or K63-linked polyubiquitin, a process that can be detected by A20 and A20 binding and inhibitor of NF-kB (ABIN-1) [58,59,60,61]. Similarly, other deubiquitinating enzymes, such as USP21, OTU deubiquitinase with linear linkage specificity (OTULIN), and zinc finger protein 91 (ZFP91), can likewise also facilitate necroptosis [62,63,64,65,66,67,68]. TRAF2 can function as an inhibitory factor to suppress necroptosis signaling by (1) promoting Complex I stabilization and NF-κB activation by ubiquitinating RIPK1 through binding to cIAP1/2 and (2) inhibiting necrosome formation by binding directly to MLKL, a process that is inhibited by CYLD, providing a direct basis for the idea that CYLD can facilitate necroptosis [57]. Following the destabilization of Complex I, TRADD and RIPK1 dissociate from TNFR1 and translocate to the cytoplasm [69]. Both proteins, through their DDs, interact with FAS associated via death domain (FADD,) an adaptor protein that recruits caspase-8 [58]. This assembly, termed Complex IIa, primarily drives apoptotic signaling and consists of TRADD, FADD, and caspase-8 (Figure 1A). Complex IIa transforms into Complex IIb under specific molecular conditions: RIPK1 ubiquitination inhibition, enhanced autophosphorylation, or stimulated kinase activity. Complex IIb comprises RIPK1, RIPK3, FADD, and caspase-8, initiating RIPK1 kinase activity-dependent apoptosis (RDA) (Figure 1A). The inhibition of caspase-8 activity promotes necrosome formation, resulting in necroptosis rather than apoptosis. Caspase-8 functions as a critical regulatory switch between these cell death pathways. Any agent preventing the caspase-8-mediated cleavage of RIPK1 and RIPK3 facilitates the apoptosis-to-necroptosis transition. At low concentrations, the heterodimerization of cFLIPL (cellular FLICE-inhibitory protein, long isoform) with caspase-8 activates caspase-8 through cleavage, thereby preventing necrosome formation and halting necroptosis. Conversely, normal or elevated cFLIPL levels inhibit both caspase-8 activity and subsequent apoptotic and RDA processes [70,71,72].

2.2. Fas or TRAILR1/2

Upon the binding of FASL and TRAILR1/2 to their corresponding ligands, FADD and caspase-8 are subsequently recruited to form the death-inducing signaling complex (DISC), which cleaves caspase-3/7 and triggers apoptosis [73,74]. Similar to TNFR1 signaling in Complex IIb, DISC can likewise bind to RIPK1, leading to the onset of RDA. When caspase-8 activity is inhibited, cell necroptosis can be triggered further [75]. In the cytoplasm, FADD and caspase-8 assemble a multiprotein complex resembling Complex I, termed the “FADDosome.” This structure incorporates TRADD, TRAF2, cIAP1/2, RIPK1, TAK1, and the IKK complex, thereby activating the NF-κB and MAPK pathways to initiate inflammatory responses [76]. Not surprisingly, in contrast to TNFR1-induced signals, the FAS- and TRAILR1/2-induced signals form DISCs to trigger the initiation of death before the formation of intracytoplasmic complexes to trigger inflammatory signaling (Figure 1B).

2.3. DR3 and DR6

Death receptor 3 (DR3, also known as TNFRSF25) was originally identified as an essential T cell co-stimulatory molecule [77]. In addition to apoptosis, DR3 can mediate necrosome formation in the presence of its ligand TL1A [78] (Figure 1A). In allergic inflammation, such as asthma, TL1A-DR3-induced necroptosis may be one of the major causes of exacerbation [79]. TL1A inhibitors have demonstrated therapeutic efficacy in inflammatory bowel disease patients, highlighting the potential significance of necroptosis in disease pathogenesis [80].
Death Receptor 6 (DR6) is extensively represented in a diverse range of cells, including neuronal cells and endothelial cells [81]. The binding of amyloid precursor protein (APP) to DR6 elicits necroptosis [82] (Figure 1A). Tumor cells can induce endothelial cell necroptosis by secreting APP, which in turn facilitates their metastasis [83]. Thus, it is apparent that all DRs are inextricably intertwined with necrotic apoptosis.

2.4. TLR3/4 and ZBP1

TLR3 is stimulated by polyinosine–polycytidylic acid (poly(I:C)), dsRNA, and vRNA, whereas TLR4 is upregulated by lipopolysaccharide (LPS), CD14, or myeloid differentiation primary response 88 (MyD88) [84,85]. TLR3 and TLR4 initiate necroptotic signaling by engaging RIPK1 and RIPK3 through Toll/IL-1 receptor domain-containing adaptor inducing interferon-beta (TRIF), which interacts with the RHIM domain. This assembly further recruits TRADD, TRAF2, cIAP1/2, LUBAC, and the FADD-caspase-8 complex, ultimately activating the NF-κB and MAPK pathways [86]. When ubiquitination and phosphorylation are inhibited, this complex can trigger apoptosis, and then caspase-8 activity, when further inhibited, can initiate necrosome formation and necroptosis [87] (Figure 1B).
ZBP1 (also known as DNA-dependent activator of IFN regulatory factors (DAI)) possesses a RHIM structural domain that binds to RIPK3, inducing the phosphorylation of RIPK3 and subsequent phosphorylation of MLKL, leading to necroptosis, independently of RIPK1 [88] (Figure 1B). ZBP1 further triggers cellular inflammation via RIPK1–FADD–caspase-8-induced apoptosis [89]. Notably, ZBP1 mediates nucleus-to-cytoplasm necroptosis. Influenza A virus (IAV) replication produces z-RNAs that activate ZBP1 within the nucleus of infected cells. This activation triggers RIPK3-dependent MLKL phosphorylation, resulting in nuclear membrane disruption, cytosolic DNA release, and subsequent necroptosis [90]. However, it has been demonstrated that the RHIM of RIPK1 inhibits ZBP1 from binding and the activation of RIPK3, further inhibiting ZBP1-induced necroptosis [91]. Recent investigations have demonstrated that RIPK1-DD inhibits ZBP1- and TRIF-induced necroptosis, which is mediated by the cleavage of RIPK1 and RIPK3 by FADD-caspase-8, which is recruited by RIPK1-DD, inhibiting its activation [92].

2.5. ROS

ROS are ubiquitous in the development of necroptosis and are primarily generated by the NOX family and mitochondria. ROS have been implicated in facilitating RIPK1 autophosphorylation, thereby promoting RIPK1 and RIPK3 to generate the necrosome for necroptosis [93,94,95,96,97]. The potential mechanism involves ROS facilitating the oxidation of the RIPK1 C257, C268, and C586 cysteines to generate disulfide bonds to facilitate RIPK1 polymerization, which would promote RIPK1 S161 autophosphorylation and enhanced affinity for RIPK3 [94]. ROS can also facilitate RIPK3-MLKL production [98] or p-MLKL oligomerization independently [99]. ROS can also activate RIPK3-dependent necroptosis production through the c-Jun N-terminal kinase (JNK) pathway [100,101]. Conversely, ROS production is modulated by the RIPK1-RIPK3 complex, which facilitates ROS production [102]. RIPK3 can also facilitate ROS production through the JNK pathway [103]. TRADD has been demonstrated to bind specifically to RIPK3 to trigger ROS production and necroptosis, a process that is independent of RIPK1 [104]. Emerging evidence suggests that mitochondrial ROS (mtROS) play a pivotal role in the transition from pyroptosis to necroptosis. In macrophages carrying the leucine-rich repeat kinase 2 (LRRK2G2019S) mutation, bacterial infection triggers inflammasome activation, resulting in the caspase-1-mediated cleavage of GSDMD and pro-IL-1β. Unlike typical pyroptosis, GSDMD pores localize to mitochondria rather than the plasma membrane, causing mitochondrial damage. This aberrant localization promotes mtDAMP release, mtROS production, and subsequent RIPK1/RIPK3 activation, thereby shifting the cell death modality toward necroptosis [105] (Figure 1A,B).

2.6. IFN

Type I and II interferons (IFNs) promote necrosome assembly under two distinct conditions: (1) inhibition or loss of FADD (via S194 phosphorylation) or caspase-8 activity, and (2) activation of the viral RNA-sensing kinase (PKR, protein kinase R), which interacts with RIPK1 to facilitate necrosome formation [106]. Additionally, cytoplasmic nucleic acid sensors such as RIG-I and STING induce IFN-I and TNFα production, creating an autocrine loop that amplifies necroptotic signaling [107,108]. IFN-I activates the ISGF3 complex (STAT1-STAT2-IRF9) through IFNAR1 binding, leading to transcription-dependent necrosome activation [109]. Similarly, the TNF-IRF1 and LPS-IRF3/7 signaling pathways stimulate IFNβ production, further reinforcing this autocrine feedback mechanism (Figure 1B) [109].
Overall, the conditions under which necroptosis occurs are more demanding, and caspase inactivation is generally observed after viral infection, so we can envisage that in the natural situation, necroptosis is a complementary death pathway in response to viral infection, in which caspase-8 mediates the conversion of apoptosis to necroptosis.

3. Post-Translational Modifications in Necroptosis

3.1. Ubiquitination

Ubiquitination forms a central regulatory framework underlying necroptotic pathways. This post-translational modification involves the covalent attachment of ubiquitin (Ub) molecules to substrate proteins, typically linking the C-terminal glycine residue (Gly76) of ubiquitin to lysine side chains on target proteins. The ubiquitination cascade proceeds through a three-enzyme relay: E1 (ubiquitin-activating enzyme) initiates the process through ATP-dependent ubiquitin activation, E2 (ubiquitin-conjugating enzyme) carries the activated ubiquitin, and E3 (ubiquitin ligase) facilitates the final transfer to specific substrate proteins. This sophisticated enzymatic machinery enables the precise control over necroptotic signaling components, influencing their stability, activity, and interaction dynamics [110]. The removal of ubiquitin molecules is mediated by the DeUBiquitinating enzyme (DUB) family. The linkage of ubiquitin can be classified as monoubiquitination and polyubiquitination. Monoubiquitination can be subdivided into monoubiquitination and multi-monoubiquitination, which refers to the attachment of a single Ub molecule to a target protein or the attachment of multiple Ub molecules to multiple lysines (K) of a target protein. Polyubiquitination means that Ub can polymerize to form a chain or branching structures. Ub contains seven lysines (K6, K11, K27, K29, K33, K48, K63) and an N-terminal methionine (M1), and when a specific residue links all the Ub monomers together, the chain is named after that residue, such as M1-linked chains. Chains composed of different Ub residues produce different spatial conformations, which in turn affect their functions. The chains connected by M1- and K63- adopt a linear chain-like structure, whereas the chains connected by K48- have a zig-zag spherical structure. Different ubiquitin linkage types orchestrate distinct cellular fates. K63- and M1-linked chains primarily facilitate protein complex assembly and signaling platform formation. In contrast, K48-linked chains serve as canonical degradation signals, marking proteins for recognition and proteolysis by the ubiquitin proteasome system (UPS) [111]. It is noteworthy that Ub molecules allow the formation of mixed chains, which refers to the fact that, for example, they can form K63-linked chains before being modified by M1-linked chains, thereby altering their biological function [112].

3.1.1. RIPK1

The modification of RIPK1 by ubiquitination regulates cell fate, determining whether it survives, whether pro-inflammatory signaling pathways are activated, or whether it dies. RIPK1 comprises multiple structural domains. These include the middle RHIM domain, the C-terminal DD, and the N-terminal serine/threonine kinase domain [69] (Figure 2A,B). The RHIM structural domain is responsible for its interaction with other proteins that possess RHIM, while the serine/threonine kinase domain possesses kinase activity that phosphorylates other proteins, and through the DD, stimulated by various signals, it can interact with other proteins possessing a DD such as TRADD, FADD, and caspase-8, to form Complex I.
The E3 ligases cIAP1 and cIAP2 serve as pivotal molecular regulators by catalyzing the K11-, K48-, and K63-linked polyubiquitination of RIPK1 and facilitating its recruitment to Complex I via TRADD-TRAF2 interactions. Specifically, these molecules modify RIPK1 at residue K377 through K63-linked ubiquitination, enabling subsequent LUBAC binding and NF-κB pathway activation [42]. This K63-linked modification is crucial for Complex I stabilization. SMAC mimetics, which inhibit cIAP1 and cIAP2, predictably promote the transition from Complex I to Complex IIa/b. Beyond K63-linked modifications, K48-linked polyubiquitination also regulates RIPK1 function. This regulation depends on cIAP1’s ubiquitin-associated (UBA) domain, which simultaneously suppresses RIPK1 kinase activity and promotes its K48-linked polyubiquitination-mediated degradation. These mechanisms effectively prevent the progression from Complex I to Complex II and block subsequent death signaling pathways [113]. In mice, K376 of RIPK1 (corresponding to K377 in humans) seems to be the primary site targeted by cIAP1/2 for ubiquitination. When RIPK1 has a K376R mutation, it changes the formation of the TNFR1 complex. This mutation also decreases the K11, K63, and linear ubiquitination of RIPK1, ultimately resulting in the death of mouse embryos [114]. LUBAC, another E3 ligase, consists of HOIL-1L, HOIP, and SHARPIN and is linked to Complex I via a K63-linked polyubiquitination chain added by cIAP1/2 [44,45,46,47]. LUBAC interacts with RIPK1 and NEMO via a linear polyubiquitination chain and can provide M1-linked chains for TRADD, TNFR1, and RIPK1, which provide attachment sites for the IKK complex, TAK1 complex, and TBK1, among others, which can further phosphorylate RIPK1 and stabilize Complex I [115,116]. Notably, PP6 is a member of the PPP family of serine/threonine protein phosphatases and has been suggested as a possible member of the novel Complex I that promotes necroptosis by negatively regulating LUBAC-mediated M1-linked polyubiquitination, facilitating RIPK1 activation and cFLIP degradation [117]. There is also an E3 ubiquitin ligase, Mind Bomb-2 (MIB2), which modifies the RIPK1 C-terminal portions K377 and K634 by adding K11-, K48-, and K63-linked polyubiquitination, disrupting RIPK1 oligomerization and the RIPK1-FADD association and inhibiting death signaling [118]. Poly ADP-ribosylation (PARylation) and PARylation-dependent ubiquitination (PARdU)-mediated ubiquitination occur on mouse RIPK1 K376 in response to necroptosis signaling stimulation. PARdU of RIPK1 is mediated by poly(ADP-ribose) polymerase 5A (PARP5A) and the E3 ubiquitin ligase RING finger protein 146 (RNF146), which interact multivalently to form a fluid cohesive complex that is recruited to activated RIPK1 by Tax1-binding protein 1 (TAX1BP1) to promote the proteasomal degradation of RIPK1 and inhibit necroptosis [119]. Mitsugumin 53 (MG53), an E3 ubiquitin ligase, attaches multiple ubiquitin chains to RIPK1 at residues K316, K604, and K627. This leads to the proteasome-mediated degradation of RIPK1, thereby inhibiting necroptosis [120].
Conversely, DeUBiquitinating enzymes can remove ubiquitination modifications and thereby regulate cell survival or death, including A20 (and its binding protein ABIN1), CYLD, USP21, and OTULIN. They can remove K63- or M1-linked polyubiquitination on human RIPK1 K377(mouse K376), leading to Complex I destabilization that triggers death signaling [62,63,64,65,66,67]. Zinc finger protein 91 (ZFP91) also induces RIPK1 deubiquitination to stabilize RIPK1 to promote necroptosis, but its specific site is unknown [68].
Ubiquitination can not only inhibit death signaling, but it can also promote death signaling. The E3 ubiquitin ligase Pellino 1 (PELI1) modifies K63-linked polyubiquitination chains on K115, which mediates RIPK1 activity and promotes the binding of activated RIPK1 to RIPK3 and MLKL to form a necrosome, leading to necroptosis [121]. However, the K115 mutation (present in both humans and mice) has no impact on RIPK1 ubiquitination. It also does not affect the TNF-stimulated NF-κB and MAPK signaling pathways [114]. Another E3 ligase, c-Cbl, modifies RIPK1 at K158 using K63-linked chains. When TAK1 is inhibited, LRRK2, anaphase-promoting complex subunit 11 (APC11), and c-Cbl are more strongly recruited to Complex I. This leads to the ubiquitination of RIPK1, forming a large, insoluble RIPK1 complex (iuRIPK1). This complex is an intermediate between Complex I and IIa/b and can trigger apoptosis or necroptosis [122]. The ubiquitination of RIPK1 in the cytoplasm but not in Complex I can also determine cell fate.
Under normal physiological circumstances, the E3 ligase carboxyl terminus of Hsp70-interacting protein (CHIP) binds to RIPK1 in the cytoplasm. This interaction leads to the ubiquitination of RIPK1 at K571, K604, and K627. Subsequently, the ubiquitin-proteasome system degrades RIPK1, thereby regulating its level of stability [123].
Overall, RIPK1 is degraded intracellularly in four ways at the same sites as mouse RIPK1 K376 or human RIPK1 K377: (1) UBA structural domain of cIAP1-mediated polyubiquitination [113]; (2) PARdU-mediated polyubiquitination [119]; (3) MG53-mediated polyubiquitination [120]; and (4) CHIP-mediated polyubiquitination [123]. Furthermore, in addition to PELI1- and c-Cbl-mediated ubiquitination, the ubiquitination of RIPK1 generally inhibits necroptosis by affecting its kinase activity and causing its proteasomal degradation (Figure 1A and Table 1).

3.1.2. RIPK3

RIPK3 contains the RHIM structural domain and can interact with other proteins possessing the RHIM structural domain such as RIPK1, ZBP1, and TRIF to form a complex [14] (Figure 2A,B). The ubiquitination of RIPK3 at K5 stabilizes the RIPK1-RIPK3 complex. In contrast, the deubiquitinating enzyme A20 removes K63-linked polyubiquitination chains from RIPK3, which in turn inhibits necroptosis [139]. A20 binding and inhibitor of NF-kB 3 (ABIN3) can recruit A20 to enhance its inhibitory effect [140]. Mouse RIPK3-K469 ubiquitination restricts necroptosis and apoptosis by preventing upstream ubiquitination of K359, suggesting that the ubiquitination of K359 promotes necroptosis, but the enzymes that ubiquitinate it are not known [141]. The RIPK3 residues K158, K287, and K307 limit cell death in a way comparable to K469 [141].
Unlike A20, ubiquitin-specific peptidase 22 (USP22) is a DUB that deubiquitinates RIPK3 on K42, K351, and K518, but USP22 specifically promotes necroptosis through the deubiquitination of RIPK3 K518 [142].
In contrast, the following E3 ligases inhibit the proteasomal degradation of RIPK3 by adding K48-linked polyubiquitination that mediates its function. In the cytoplasm, RIPK3 K55 and K363 can be modified by the addition of ubiquitination by the aforementioned E3 ligase CHIP, leading to their lysosomal degradation and the negative regulation of necroptosis [123]. PELI1 acts as an E3 ligase that can add a K48-linked polyubiquitination to RIPK3 K363, whereas the phosphorylation of RIPK3 T182 leads to its interaction with PELI1, which, above all, results in the degradation of the kinase-active RIPK3 [143]. The E3 ligase TRIM25 directly interacts with RIPK3 via its SPRY domain. This interaction enables TRIM25 to mediate the addition of K48-linked polyubiquitination chains to RIPK3 at K501, which in turn promotes RIPK3 degradation [144]. RIPK3 K264 can be modified by K48-linked polyubiquitination and degraded by the UPS, but its E3 ligase is unknown [145].
Parkin is an E3 ligase. RIPK1/3 activates AMPK, which then phosphorylates Parkin at S9. This phosphorylation causes the polyubiquitination of RIPK3. Specifically, K33-linked polyubiquitination occurs at RIPK3 residues K197, K302, and K364. As a result, RIPK3 changes its conformation, inhibiting necrosome formation [146] (Figure 1A and Table 2).

3.1.3. MLKL

MLKL consists of an N-terminal 4HB domain, a two-helix “brace” region, and a C-terminal pseudokinase domain. These three structural domains are crucial for MLKL’s killing effect [165] (Figure 2A,B). The ubiquitination of endogenous MLKL occurs on K51, K77, K172, and K219. RIPK3 phosphorylation of MLKL followed by a K63-linked polyubiquitination on K219 (human K230) prior to its translocation to the cell membrane can promote its oligomerization, ultimately leading to necroptosis, in which the sequence of phosphorylation and ubiquitination of MLKL is critical [166]. It has also been demonstrated that the ubiquitination of MLKL inhibits necroptosis. In the 4HB structural domain of MLKL, residues K9, K51, K69, and K77 are subject to ubiquitination modifications. Monoubiquitination of MLKL does not lead to cell death but instead promotes its proteasome- and lysosome-mediated degradation. In contrast, USP21, a DUB that removes all ubiquitination modifications on all of MLKL, enhances MLKL activation when it produces MLKL-USP21 fusion proteins with MLKL, and it can be assumed that the monoubiquitination of MLKL inhibits necroptosis [167]. Ubiquitination of MLKL aids in clearing intracellular bacteria. In human MLKL, K50, and in mouse MLKL, K50/51 can be modified by E3 ligase ITCH-mediated K63-linked polyubiquitination. This modification enables MLKL to bind to endosomal membranes. Subsequently, MLKL is exocytosed from the cell via extracellular vesicles. As a result, it enhances the disruption of intracellular bacteria by promoting the translocation of bacteria from the endosome to the lysosome [168]. Moreover, the E3 ligase S-phase kinase-associated protein 2 (Skp2) participates in the K48-linked polyubiquitination of MLKL, leading to its degradation. Yet, the precise modification site is still unclear [169].
Interestingly, the LUBAC-mediated M1-linked polyubiquitin modification does not act directly on MLKL, but rather, sways cell death by modulating the subcellular distribution of active MLKL [170] (Figure 1A and Table 3).

3.1.4. FADD

FADD consists of a DD and death effector domain (DED), which plays a key role in constituting the DISC and interacting with caspase-8 [71]. The E3 ligase CHIP mediates the K6-linked polyubiquitination of FADD at K149 and K153. This process is crucial for preventing the formation of the DISC and thus inhibiting cell death [181]. In contrast, a distinct E3 ligase, Makorin ring finger protein 1 (MKRN1), promotes the proteasomal degradation of FADD. This action inhibits the formation of the DISC and death signaling. Nevertheless, the specific site of its action requires further study [182] (Table 4).

3.1.5. Caspase-8

Procaspase-8 has two isoforms, procaspase-8a and procaspase-8b. These are activated mainly when two of their N-terminal death effector domains (DED 1 and DED 2) are recruited to the DISC and dimerized at the DED filaments formed there [214]. The two DEDs of procaspase-8a/b contain the large catalytic domain p18 and the small catalytic domain p10, respectively [214]. The E3 ligase Cullin 3 (CUL3) mediates the K48- and K63-linked polyubiquitination of K461 on the p10 subunit. This stabilizes the active caspase-8 heterotetramer. The process involves p62 binding to CUL3-ubiquitinated caspase-8 and forming an aggregated structure, enhancing procaspase-8 activity [192]. During endoplasmic reticulum (ER) stress, the E3 ligase TRIM13 promotes caspase-8 activity through K63-linked ubiquitination, but the specific ubiquitination site on caspase-8 is unknown [193].
Conversely, homologous to the E6-AP Carboxyl Terminus 3 (HECTD3) causes K63-linked polyubiquitination at caspase-8 K215. This does not lead to caspase-8 degradation but reduces its activation, promoting cell survival [194]. At the DISC, TRAF2 interacts with caspase-8, resulting in polyubiquitination at K224, K229, and K231 of the p18 structural domain. This leads to the proteasomal degradation of active caspase-8 and inhibits death signaling [195]. Similarly, the DR5-Cbl-b/c-Cbl-TRAF2 complex interacts with TRAF2. TRAF2 then mediates the K48-linked polyubiquitination of caspase-8, promoting its proteasomal degradation [215] (Table 4).

3.1.6. cFLIP

The cFLIP proteins contain one long isoform called cFLIPL and two short isoforms called cFLIPS and cFLIPR [216]. The cFLIP protein has two DED structural domains at the N-terminus, whereas cFLIPL also has a catalytically inactivating cysteine asparaginase-like structural domain in its C-terminal region (p20 and p12) [216]. All types of cFLIP are widely recognized as antiapoptotic proteins, and when expressed at high levels, they compete with procaspase-8 for recruitment to the DISC via their DED domains. When expressed at low levels, cFLIPL exerts its pro-apoptotic function via the formation of a procaspase-8/cFLIPL heterodimer. In this heterodimer, cFLIPL stabilizes the active center of procaspase-8, thus enhancing caspase-8 activity. In contrast, cFLIPS and cFLIPR inhibit caspase-8 activation. They do so by forming an inactive heterodimer with procaspase-8, which prevents caspase-8 from being activated [217]. The HOIP catalytic subunit of the E3 ligase LUBAC forms M1-linked polyubiquitinated chains at K351 and K353 of cFLIP, which can inhibit K48-linked polyubiquitination, stabilize cFLIP, inhibit its degradation, and suppress the death process [205]. MIB2 may add K48- and K36-linked polyubiquitination at nine sites, namely, K351, K353, K381, K386, K389, K460, K462, K473, and K474, in cFLIPL, stabilizing cFLIP and inhibiting RIPK1 kinase activity and Complex II production [206]. Both the Skp1-Cullin-1-F-box (SCF) Cullin-Ring E3 ubiquitin ligase complex (SCFSkp2) with Skp2 and the E3 ubiquitin ligase ITCH can promote the degradation of cFLIP through ubiquitination. However, the precise ubiquitination site remains unclear [209,210]. The deubiquitinating enzyme (DUB) Usp27x inhibits cell death related to cFLIP. It does this by removing the K48-linked polyubiquitination added to cFLIP by the E3 ligase TRIM28. As a result, it blocks the proteasomal degradation of cFLIP [211] (Table 4).

3.1.7. TNFR1

Upon activation, TNFR1 on the cell surface quickly recruits the E3 ligase RNF8. RNF8, along with the E2 ubiquitin-conjugating enzyme Ubc13, binds to TNFR1. RNF8 then modifies TNFR1 by adding K63-linked polyubiquitination to the receptor. This modification triggers the internalization of activated TNFR1, an essential step for cell death. Nevertheless, the precise site of this modification remains unconfirmed [186] (Table 4).

3.2. Phosphorylation

Protein phosphorylation, the addition of phosphate groups to proteins, is one of the most common and crucial post-translational modifications (PTMs). It is invariably involved in the necroptosis process [218]. Nine amino acid residues can be phosphorylated: serine, threonine, tyrosine, histidine, lysine, arginine, aspartic acid, glutamic acid, and cysteine. Depending on the amino acid residue, four types of phosphorylation can occur. Serine, threonine, or tyrosine hydroxyl groups undergo O-phosphorylation (forming a P-O bond). The nitrogen-containing side chains of histidine, lysine, or arginine experience N-phosphorylation (P-N bond). Aspartic acid or glutamic acid carboxyl groups are subject to A-phosphorylation (P-OCO bond), and cysteine sulphonyl groups participate in S-phosphorylation (P-S bond). Like ubiquitination, protein phosphorylation is regulated by phosphorylases and dephosphorylases. This regulation impacts cellular activities such as growth, development, death, and senescence [219].

3.2.1. RIPK1

RIPK1 is more prone to undergoing autophosphorylation to stimulate downstream signaling, possibly because of its own weaker kinase activity [125]. RIPK1 S166 autophosphorylation has been extensively and widely studied and is well known [125,126]. In addition to S166, autophosphorylation can also occur in human S14, S15, S20, and S161 and mouse S14, S15, S161, and T169 [127]. S166 autophosphorylation may not cause cell death alone but enhances RIPK1 kinase activity, promotes autophosphorylation at other sites, and may serve as a reliable biomarker of RIPK1 kinase-dependent cell death [126]. The role regarding the phosphorylation of RIPK1 S161 is not as consistent as that of S166. It has been suggested that S161 phosphorylation has little effect on RIPK1 kinase activity [129]. In contrast, it has also been suggested that the autophosphorylation that occurs on S161 recruits more RIPK3 and therefore promotes necrosome formation and necroptosis [94]. Thus, it is not difficult to see that autophosphorylation is an indispensable factor for RIPK1 to act as a cell fate determinant.
In addition to autophosphorylation, there are many other sites where phosphorylation can affect RIPK1. In Complex I, the IKK complex phosphorylates RIPK1 and protects cells from RIPK1 kinase-dependent death [220]. A follow-up study found that RIPK1 S25 appears to be the site of action of the IKK complex [127]. Given that S25 lies within the kinase structural domain of RIPK1, phosphorylation of RIPK1 at S25 by IKKα/β directly curbs RIPK1 kinase activity and S166 autophosphorylation. This action also averts TNF-mediated, RIPK1 kinase-dependent cell death [127]. The IKK complex-mediated phosphorylation of RIPK1 S25 also promotes T cell survival [221]. MAPKAPK2 has been shown to be a key kinase in limiting RIPK1 kinase activity and inhibiting Complex I to Complex II conversion [49,50,51]. Human S320 and mouse S321 and S336 are sites where MK2 acts on RIPK1 [49,50,51]. MK2 phosphorylates RIPK1 directly at S321. This phosphorylation inhibits RIPK1’s binding to FADD/caspase-8 and its induction of RIPK1 kinase-dependent apoptosis and necroptosis [50]. The tyrosine kinases JAK1 and SRC mediate tyrosine phosphorylation of RIPK1 at Y383 (Y384 in humans). This phosphorylation inhibits RIPK1 kinase activity. Y383 tyrosine phosphorylation is crucial for MK2 binding to RIPK1 and for MK2’s further activation, though the precise mechanism remains unclear [52]. TAK1, recruited into Complex I, phosphorylates S321 in the RIPK1 intermediate domain. This phosphorylation inhibits RIPK1’s interaction with FADD/caspase-8 and suppresses cell death [128]. Similarly, the autophagy-initiating kinase ULK1 phosphorylates S357 within the RIPK1 intermediate domain, TBK1 in Complex I phosphorylates RIPK1 T147 or T189/190 (human/mouse), and AMPK phosphorylates RIPK1 S415, which can inhibit RIPK1-dependent death [53,54,55].
Interestingly, RIPK1 S89 phosphorylation restricts its kinase activity, and the RIPK1 S89A mutation up-regulates RIPK1 activity, but the RIPK1 S89D mutation down-regulates RIPK1 activity, an interesting phenomenon [129].
ROS are also involved in the phosphorylation of RIPK1, which was found to accumulate in mitochondria after ROS stimulation and to undergo S166 and S321 phosphorylation, accompanied by ubiquitination. However, how this phosphorylation occurs needs to be further investigated [222].
Of course, dephosphorylation may also regulate RIPK1, but this has not been extensively studied. Protein phosphatase 1 regulatory subunit 3G (PPP1R3G) is involved in a key process. Its catalytic subunit, protein phosphatase 1 γ (PP1γ), is recruited to Complex I. There, it removes the inhibitory phosphorylation of RIPK1 at S25. This action promotes RIPK1-dependent cell death [130].
Protein 2 containing SET and MYND structural domains (SMYD2) catalyzes the lysine methylation of histones and non-histone proteins, but new studies have found that SMYD2 can regulate the phosphorylation of RIPK1 and thus inhibit necroptosis; however, the exact phosphorylation site is unknown [131] (Figure 1A, Figure 2A,B and Table 1).

3.2.2. RIPK3

RIPK3 has a similar structure to RIPK1, but without the DD domain. The difference is that TRIF and ZBP1 can interact with RIPK3 alone without RIPK1. During the activation of RIPK3, RIPK1 acts more as a scaffold rather than as its upstream kinase [85,223]. As with RIPK1, we first discuss the autophosphorylation of RIPK3. The following autophosphorylation sites have been identified, including S199, S211, S215, S227, and T182 in human RIPK3 and S204, S232, and T231 in mouse RIPK3. RIPK3 S227 is the key site for its interaction with MLKL [147,148,149,150]. The phosphorylation of human RIPK3 at T224 and S227 acts synergistically to boost its interaction with MLKL. Under normal resting conditions, the autophosphorylation of RIPK3 at S227 depends on MLKL. However, during necroptosis, this autophosphorylation of S227 becomes independent of MLKL [150]. T224 is not conserved in vivo in mice, possibly implying species-specific differences in RIPK3-MLKL interactions. Similarly, mouse T231/S232 and human S227 are critical for recognition of their homologous MLKL immediate homologs, which is also species-specific [171]. RIPK3 S232 also mediates necroptosis-induced periodontitis in mice [224]. RIPK3 S227 can not only undergo autophosphorylation but can also be modified by other kinases. CK1α, CK1δ, and CK1ε in the casein kinase 1 (CK1) family can bind to RIPK3 during necrosome formation, phosphorylate S227, and further activate necroptosis [153]. However, additional studies have shown that the CK1 family member casein kinase 1G2 (CSNK1G2) inhibits the activation of RIPK3 dimerization and inhibits necroptosis by facilitating the binding of RIPK3 to RIPK3 monomers through RIPK3 S211/T215 autophosphorylation [151]. Thus, the effect of the CK1 family on RIPK3 and necroptosis is not absolute. Next, we discuss other phosphorylation sites. T182 is thought to be the initiating event for S227 autophosphorylation, and the substitution of T182 by alanine (T182A) abrogates S227 phosphorylation and prevents TNFα-induced necroptosis. Interestingly, when RIPK3 is phosphorylated at T182, it interacts with the forkhead-associated (FHA) domain of PELI1. This interaction leads to the PELI1-mediated polyubiquitination of RIPK3 at K363 with K48-linkages. Subsequently, RIPK3 undergoes proteasomal cleavage, thereby inhibiting necroptosis [143]. Substituting S204 (S199A in human RIPK3 and S204A in mouse RIPK3) abolishes the in vitro kinase activity of RIPK3. In contrast, aspartate-substitution mutants (S204D in mouse RIPK3 and S199D in human RIPK3) maintain this in vitro activity. In necroptosis, RIPK1’s key role is to promote the phosphorylation of RIPK3 at S204, either directly or indirectly [129].
The phosphorylation of RIPK3 also mediates a switch to apoptosis. When the levels of the RIPK3 chaperone heat shock protein 90 (Hsp90) and its co-chaperone Cdc37 (CDC37) are low, RIPK3 gets phosphorylated at the S165/T166 site. This phosphorylation inactivates RIPK3 kinase activity and its capacity to recruit RIPK1, FADD, and caspase-8 for forming a cytoplasmic caspase-activated complex. As a result, apoptosis occurs without necroptosis [152].
Under specific conditions, Complex IIb can assemble into structures called ripoptosomes. In the G2/M phase, RIPK3 is considered an extra component of the ripoptosome. Procaspase-8 within the ripoptosome can cleave RIPK3. Polo-like kinase 1 (PLK1) directly binds to RIPK3 and phosphorylates it at S369. This phosphorylation blocks the ripoptosome-mediated cleavage of RIPK3, preserving its pro-death activity. As a result, when mitotic errors happen, an alternative cell death pathway is available [225].
Dephosphatases also regulate RIPK3 activity. Protein phosphatase 1B (Ppm1b) prevents RIPK3 auto-activation and negatively regulates TNF-induced necrotic apoptosis in quiescent cells and is not dependent on the NF-κB pathway. Possible sites of action are T231 and S232, which can undergo autophosphorylation [154] (Figure 1A and Figure 2A,B and Table 2).

3.2.3. MLKL

MLKL, a pseudokinase, lacks the catalytic activity for phosphoryl transfer seen in RIPK1 and RIPK3. This is due to the loss of two key catalytic residues related to phosphoryl transfer. One is from the catalytic loop’s “His-Arg-Asp (HRD)” motif, and the other is from the metal-cofactor binding “Asp-Phe-Gly (DFG)” motif [165].
There seems to be a consensus that MLKL, being downstream of RIPK3, is phosphorylated by it and thus leads to cell death. Phosphorylated MLKL oligomerizes and binds to phosphatidylinositol lipids and cardiolipin. This enables its translocation from the cytoplasm to the plasma and intracellular membranes, where it directly impairs membrane integrity, triggering necroptosis [172]. During this, RIPK3 can phosphorylate T357 and S358 in human MLKL and S345 in mouse MLKL. The phosphorylation of T357 and S358 “opens” the folded MLKL structure, linking the 4HB domain to the C-terminal pseudokinase domain. Following this, phosphorylated MLKL undergoes a conformational change that aids oligomerization [165,171,172]. Calcium/calmodulin-dependent protein kinase II (CAMK2/CaMKII), rather than RIPK3, phosphorylates MLKL at the same site and can promote autophagy [173]. Interestingly, both pharmacological and alcoholic liver injury were accompanied by MLKL aggregation and S358 phosphorylation, suggesting that MLKL S358 phosphorylation may be one of the biomarkers of liver injury [172,226].
In addition to S345, mouse S158, S228, S248, S347, and T349 can be phosphorylated by RIPK3 but have different functions [174,175]. S347 has an auxiliary role to S345, as additional mutations in this residue render MLKL completely ineffective against necroptosis. Phosphorylation of S158, S228, and S248 also regulates necroptosis, and the function of the phosphorylation of T349 is unknown [174,175]. MLKL S441 phosphorylation promotes myelin degradation [176].
In addition to RIPK3, receptor tyrosine kinases from the TAM (Tyro3, Axl, and Mer) family phosphorylate MLKL at Y376. This phosphorylation controls MLKL oligomerization, not its membrane translocation or RIPK3 phosphorylation, and promotes necroptosis [177].
Phosphorylation not only promotes cell death but also inhibits necroptosis and autoinflammation. In humans, phosphorylation of MLKL at S83, or in mice at S82, curbs MLKL activity following RIPK3-mediated MLKL activation [178]. This indicates that MLKL phosphorylation has a dual-sided role in regulating necroptosis (Figure 1A and Figure 2A,B and Table 3).

3.2.4. FADD

FADD was first reported to have two phosphorylated forms called CAP1 and CAP2 [227]. FADD can be phosphorylated at S194 (phosphorylated by casein kinase I alpha (CKIα)) [187,188], S200 (phosphorylated by the anti-apoptotic kinase CK2) [189], S203 (phosphorylated by the mitotic kinases Aurora-A (Aur-A) and PLK1) [190] and is associated with cellular sublocalization. It follows that FADD phosphorylation is less closely associated with cell death (Table 4).

3.2.5. Caspase-8

Phosphorylation of Procaspase-8 T265 is essential for the promotion of necroptosis [228]. Ribosomal protein S6 kinase A1 (p90 RSK) can be activated by 3-phosphatidylinositol-dependent protein kinase 1 (PDK 1) through a non-classical mechanism. Subsequently, p90 RSK phosphorylates procaspase-8 T265. The phosphorylation of procaspase-8 T265 maintains necrosome integrity [196].
There are many other phosphorylation sites on caspase-8, but they have been shown to be more associated with apoptosis, e.g., Y380, Y448, S347, S387, T263, and T273 [197,199,200,201,202,229,230,231,232] (Table 4).

3.2.6. cFLIP

To date, three cFLIP phosphorylation sites have been identified: T166, S193, and S273. Interestingly, T166 and S193 mediate the regulatory crosstalk between cFLIP phosphorylation and ubiquitination. cFLIPL T166 has been shown to be required for K167 ubiquitination [208,209], thereby signaling to promote proteasomal degradation. In contrast, phosphorylation of protein kinase C (PKC) at S193 blocked the ubiquitination of the c-FLIPS and c-FLIPL isoforms at K195 and K192 and inhibited their degradation [207]. In addition, TNFα stimulates the AKT serine/threonine kinase 1 (Akt1)-mediated phosphorylation of cFLIPL S273, thereby promoting the proteasomal degradation of cFLIPL [212]. The degradation of cFLIPL has a negative effect on the inhibition of caspase-8 activity, and its propagation of death signaling requires further validation.

3.3. Glycosylation

Protein glycosylation can be classified into O-linked and N-linked glycosylation, depending on how glycosidic bonds are formed. O-GlcNAcylation, catalyzed by O-GlcNAc transferase (OGT) and reversed by O-GlcNAcase (OGA), plays a role in regulating processes such as host immune responses and signal transduction during pathogen infection. O-GlcNAcylation negatively regulates necroptosis. Upon necroptosis stimulation of erythrocytes, the O-GlcNAcylation of RIPK1 S331 is decreased and the phosphorylation of RIPK1 S166 is enhanced, promoting the formation of the necrosome. The OGA inhibitor Thiamet-G (TMG) reverses the decrease in the O-GlcNAcylation of RIPK1 and promotes necroptosis [132] (Figure 1A, Figure 2A,B and Table 1). The product of the hexosamine biosynthetic pathway (HBP), O-conjugated β-N-acetylglucosamine (O-GlcNAc), modifies RIPK3 via OGT. O-GlcNAcylation of RIPK3 on T467 inhibits its RHIM function and therefore inhibits necroptosis [155]. Sevoflurane (SEVO) increases the O-GlcNAcylation of RIPK3 and inhibits RIPK3 binding to MLKL, inhibiting necroptosis induced by myocardial ischemia–reperfusion injury (MIRI) [156]. OGT glycosylates RIPK3 and reduces the stability of the RIPK3 protein and therefore inhibits necroptosis, so that the loss of O-GlcNAc leads to liver fibrosis and inflammation due to unregulated necroptosis [157]. In the brains of Alzheimer’s disease (AD) patients, elevated O-GlcNAcylation of RIPK3 hampers RIPK3 phosphorylation and its interaction with RIPK1. This leads to a reduction in necroptosis, potentially offering protection against AD [158]. The traditional Chinese medicine Wu-Mei-Wan (WMW) alleviates colitis in mice. It does so by inhibiting necroptosis through enhancing RIPK3 O-GlcNAcylation [159] (Figure 1A and Figure 2A,B and Table 2).
Enteropathogenic Escherichia coli (EPEC) and Salmonella typhimurium (S. typhimurium) possess a type III secretion system (T3SS) with effectors NleB and SseK1/2/3, respectively. NleB and SseK1/2/3 can modify a number of DD proteins by Arg-GlcNAcylation, such as TRADD, FADD, and RIPK1. NleB can modify TRADD-DD R235 and RIPK1-DD R603, and SseK1/2/3 can modify TNFR1-DD R376, TRADD-DD R233/235/245, and TRAILR-DD R293/359 [133,183,184,185,233,234,235]. This blocks their signaling and prevents necroptosis from occurring (Table 1 and Table 4).

3.4. Methylation

Methylation acts through methyltransferases, demethylases, and methylation-dependent binding proteins, which are three methylation-related enzymes that perform writing, erasing, and recognition functions, respectively [236]. Methylation can occur at the DNA level, the RNA level, and the protein level, which are catalyzed by different enzymes. Protein-level methylation can be categorized into histone and non-histone methylation. Arginine and lysine are common amino acids subject to protein methylation. Arginine methylation is mediated by proteins from the protein arginine methyltransferase (PRMT) family, while lysine methylation is carried out by lysine methyltransferase (KMT) [237]. PRMT1 can methylate human RIPK3 R486 and mouse RIPK3 R479, thereby inhibiting RIPK1-RIPK3 interactions and RIPK3 phosphorylation, respectively, and suppressing necrosome formation [160]. Both PRMT1 and PRMT5 interact with RIPK1. They mediate symmetric arginine dimethylation of R486 at the C-terminus of human RIPK3. The methylation of RIPK3 inhibits necroptosis. It does this by blocking RIPK1 S166 autophosphorylation and suppressing downstream signaling [161] (Figure 1A, Figure 2A,B, and Table 2). TRAF2, which forms Complex I, can be methylated by SMYD2, promoting NF-κB expression and thus inflammatory signaling [238].

3.5. Acetylation

Acetylation involves adding acetyl groups, sourced from acetyl coenzyme A (acetyl-CoA), to specific residues in proteins [239]. Acetyltransferases, such as histone acetyltransferases (HATs), lysine acetyltransferases (KATs), and Nα-acetyltransferases (NATs), are responsible for adding these acetyl groups. Conversely, deacetylases, including histone deacetylases (HDACs) and NAD+-dependent sirtuins (SIRTs), remove them [240]. Protein acetylation is divided into histone acetylation and non-histone acetylation. It was shown that RIPK1 can form a complex with HAT1 and SIRT1. RIPK1 can be acetylated in the kinase active region at K115; in the DD region at K625, K627, K642, and K648; and next to the DD region at K596 and K599. Cellular RIPK1-caspase-8-dependent apoptosis was enhanced by the use of SIRT1 inhibitors, suggesting that RIPK1 may promote apoptosis upon acetylation, thus affecting the biological behavior of tumor cells and influencing tumor development [134]. It is reasonable to speculate whether RIPK1 acetylation may also promote necroptosis when caspase-8 activity is inhibited (Figure 1A, Figure 2A,B and Table 1). Conversely, STIR2 can bind RIPK3. When TNF-α is stimulated, RIPK1 is activated and binds more easily to the STIR2-RIPK3 complex. This complex then deacetylates RIPK1. The likely acetylation site of RIPK1 is near K530 in the RHIM structural domain. This deacetylation process promotes the start of necroptosis (Figure 1A, Figure 2A,B, and Table 2). In acute oxalate nephropathy, inhibition of HDAC6 reduces necroptosis. In this process, HDAC6 does not appear to act directly with proteins such as RIPK1 and RIPK3 to exert its deacetylation effect, but rather, with microtubule proteins to promote inflammatory signaling, and the exact mechanism needs to be further elucidated [241]. HDAC3 indirectly regulates necrotic apoptosis. Macrophages deficient in HDAC3 have elevated histone acetylation and increased cathepsin B (CTSB), leading to increased degradation of RIPK1 and inhibition of inflammatory signaling or death signaling [242].

3.6. Disulfide Bonds

The formation and breaking of disulfide bonds can also be considered as a type of PTM. ROS promotes the formation of disulfide bonds between RIPK1 C257, C268, and C586; promotes RIPK1 S161 autophosphorylation; and increases the affinity for RIPK3, enhancing death signaling [94] (Figure 1A and Figure 2A,B and Table 1). Hypothiocyanic acid (HOSCN) acts to inhibit caspase-8 activity and promote necroptosis. It does this by catalyzing the formation of a disulfide bond. This bond links dimers between C360 in the large catalytic subunit and C409 in the small catalytic subunit of caspase-8 [243] (Table 4). MLKL forms disulfide-dependent amyloid polymers via disulfide bonds. This might occur because RIPK1/3 polymers phosphorylate MLKL, causing MLKL molecules to be spaced apart. Then, intramolecular and intermolecular disulfide bonds form through cysteines. This further promotes MLKL tetramer formation and the start of necroptosis. Inhibiting disulfide bond formation at human MLKL C86 can block MLKL polymer formation and subsequent cell death [179]. The thiol oxidoreductase thioredoxin-1 (Trx1) has C32 that cross-links with C86 of human MLKL. This keeps MLKL in a reduced state, inhibits its disulfide bonding, and blocks MLKL polymer formation. Thus, Trx1 can be regarded as an inhibitor of necroptosis [180] (Figure 1A and Figure 2A,B, and Table 3).

3.7. Caspase Cleavage

Protein cleavage mediated by the caspase family can lead to changes in protein structure and is also a type of PTM. Caspase-8 can be inhibited by the cleavage of RIPK1 D324 and RIPK3 D328 [135,162] (Table 1 and Table 2). Not only can caspase-8 cleave RIPK1, but so can caspase-6 and caspase-10, and the cleavage sites appear to be consistent [135]. RIPK1 can be activated in a DNA double-strand break (DSB)-induced signaling platform called the ripoptosome. When caspase is active, caspase-6-mediated cleavage inactivates RIPK1 completely, resulting in apoptosis. However, when caspase is inactive, double-strand breaks (DSBs) enhance NF-κB signaling and the generation of pro-inflammatory cytokines. If caspase-8 is not activated simultaneously, TNF-α signaling stimulation promotes necroptosis [136]. In addition to its cysteine enzyme activity, caspase-6 promotes necroptosis through binding directly to RIPK3. The binding of RIPK3 to the RHIM of ZBP1 to promote necroptosis is not required for its caspase activity [244]. In primary biliary cholangitis (PBC), macrophage caspase-10 has a greater cleavage capacity than caspase-8, with an increased ability to form complexes with RIPK1 and FADD, which can better cleave RIPK1 and inhibit necroptosis, and caspase-10 knockout macrophages are more likely to trigger pyroptosis and necroptosis [137] (Figure 1A and Figure 2A,B).

3.8. Nitrosylation

In cerebral ischemia–reperfusion, inward calcium currents mediated by NMDA receptors activate neuronal nitric oxide synthase, which in turn induces NO production. The C119 residue of RIPK3, when nitrosylated by NO S-nitrosylation, strengthens its kinase activity and contributes to the development of apoptosis and necroptosis [163,164] (Table 2). Caspase-8 can be S-nitrosylated by nitric oxide, which inhibits its activity and interrupts its apoptotic signaling, and whether it promotes necroptosis signaling and its specific modification sites need to be further investigated [203]. The nitrosylation of cFLIPL at C254 and C259 blocks its ubiquitination and proteasomal degradation. This leads to an increase in the cFLIPL concentration. Ordinarily, elevated cFLIPL would promote the formation of the cFLIPL-caspase-8 heterodimer, enhancing caspase-8 activity. However, nitrosylation of the cFLIPL p20 subunit weakens its capacity to stabilize the active site of the cFLIPL-caspase-8 heterodimer. As a result, it functions as an inhibitor of caspase-8 activity [213] (Figure 1A and Figure 2A,B, and Table 4).

3.9. SUMOylation

The small ubiquitin-like modifier (SUMO) can be attached to the FADD DD at K120, K125, and K149 by the E3 SUMO protein ligase PIAS3. This SUMOylation may be linked to a form of mitochondrial fission and caspase-10-related cell death [191]. Caspase-8 can be SUMOylated at K156. This modification is associated with the nuclear localization of caspase-8 but does not impact its activation [204] (Table 4). RIPK1 is SUMOylated at K550, and the SUMOylation of RIPK1 promotes its activation, which is reversed by SUMO-specific protease 1 (SENP1), inhibiting cell death and the development of non-alcoholic steatohepatitis [138] (Figure 1A and Figure 2A,B and Table 1).

4. Conclusions

Necroptosis activation stems from diverse stimuli and undergoes regulation via a myriad of signaling mediators and their post-translational modifications. These elements form sophisticated interdependent networks that fundamentally modulate necroptotic progression. Beyond the aforementioned molecular components, numerous undiscovered entities pertinent to necroptotic pathways warrant comprehensive investigation in forthcoming research endeavors. FK506-binding protein 12 (FKBP12) regulates protein folding and conformational changes. It is closely linked to the autophosphorylation of RIPK1 and RIPK3, as well as the formation of the necrosome [245]. ZFP91 induces the deubiquitination of RIPK1 to stabilize RIPK1 to promote cell death, promotes RIPK1-RIPK3 interactions to stabilize RIPK1 and RIPK3 proteins, and promotes necrotic apoptosis but also promotes mitochondrial ROS production [68]. Under normoglycemia, Cannabinoid receptor 2 (CB2R) represses the expression of RIPK1, RIPK3, and MLKL at the transcriptional level. However, under hyperglycemia, MLKL phosphorylates CB2R to mediate its ubiquitination degradation and promote necroptosis and diabetic heart dysfunction [246]. Interferon-inducible 2′-5′ oligoadenylate synthetase-like (OASL) provides a scaffold for ZBP1-RIPK3-MLKL binding in virus-induced necroptosis and thus has potent antiviral activity [247]. Molecules such as these are increasingly being discovered and studied, suggesting that necroptosis has additional undiscovered mechanisms.
Necroptosis is linked to numerous human diseases. For instance, in infectious diseases such as various viral and bacterial infections, it is associated with the activation of TLRs or ZBP1 [248,249]; it is also associated with tissue injury, including the quantity and type of ischemia–reperfusion injury [250]; organ damage such as liver, pulmonary, myocardial and interstitial fibrosis [251,252]; cardiovascular diseases such as atherosclerosis; central nervous system diseases; liver diseases such as chronic inflammation and fibrosis in aging liver [251,253]; intestinal diseases such as inflammatory bowel diseases; and autoimmune diseases, all of which are associated with human inflammatory diseases [254]. Moreover, necroptosis and its post-translational modifications (PTMs) play crucial roles in tumor development and treatment. This has significant implications for understanding oncological pathophysiological mechanisms and advancing therapeutic research [255,256]. Therefore, by gaining a deep understanding of the mechanisms of necroptosis and the means of its PTMs, we can use a wide range of agonists and inhibitors to promote or prevent the onset of necroptosis and to act as a therapeutic treatment for disease. For instance, different kinase inhibitors can be used to block the kinase activity of RIPK1. Ubiquitination agonists can also be applied to modify RIPK1 further and inhibit necroptosis. This may prevent, reduce, or delay the start and development of some diseases. Necroptosis inhibitors such as Nec-1s and GSK’872 can effectively suppress the expression of p-MLKL. As a result, they significantly inhibit fat necrosis and subsequent fibrosis in fat grafts [257]. In addition, the RIPK1 inhibitor GSK2982772 (compound 5) has the potential to become an effective treatment for psoriasis, rheumatoid arthritis, and ulcerative colitis [258]. In addition, epigenetic silencing of RIPK3 in hepatocytes is a potential target for novel drugs that have been shown to inhibit MLKL-mediated necroptosis, which induces various liver pathologies [259]. There are a number of other drugs that can exert a therapeutic effect by inhibiting necroptosis, such as ursolic acid in intestinal ischemia–reperfusion injury via STAT3 signaling [260] and rapamycin in doxorubicin-induced cardiomyopathy via the RIPK3-MLKL pathway, whose PTMs need to be explored and identified [261]. Conversely, drugs and therapies designed to promote tissue necroptosis by regulating its PTMs can be used to treat several cancers. For example, a combination of SHK and Chi-Ag NPs is able to induce effective ICD in triple-negative breast cancer tissues by synergistically inducing tumor cell necroptosis through the upregulation of RIPK3, pRIPK3, and tetrameric MLKL expression [262]. The novel isobavachalcone (compound 16) also has a therapeutic effect in NSCLC by upregulating RIPK3 and MLKL to mediate necroptosis [263]. The novel small molecule VDX-111 induces necroptosis. It does so by regulating RIPK1 expression, thereby inhibiting the progression of ovarian cancer [264]. Table 5 summarizes part of these agents, including well-characterized inhibitors such as GSK’872 (RIPK3) and GSK2982772 (RIPK1), as well as emerging candidates under preclinical and clinical evaluation. This table serves as a valuable resource for understanding the therapeutic landscape of necroptosis-targeting drugs and their potential applications in disease management.
In addition to the targeted agents mentioned above, for which specific effects are known, there are other potential targets for further investigation. Glycosylation has been shown to modify proteins on RIPK1, RIPK3, and the DISC complex in necroptosis, and all appear to negatively regulate them; for example, the O-GlcNAcylation of human RIPK1 S331 can inhibit the formation of the RIPK1-RIPK3 complex [132], OGT mediates O-GlcNAcylation on RIPK3 T467, which can inhibit necroptosis to treat septic inflammation [155], and glycosylation has been linked to metabolism as well [133,155,157,158,159,183,184,185,233,234]. This may be a trend for future research. The effects of methylation and acetylation on the modulation of necroptosis are unambiguous and may therefore also be targets for drug design [134,160,161]. Additionally, phosphorylation also plays a significant role in the regulation of necroptosis, and proteins such as IKKα/β [220], TBK1 [54], MK2 [50], Pellino 1 [143], CK1 [272], and TAM [177] may be targets for novel drugs. Dephosphorylated proteins such as Ppm1b [154] and TRAF2 [57] have also been implicated in necroptosis. In addition, more kinases and ubiquitination E3 ligases have been identified, such as JAK1, SRC [52], ULK1 [53], PARdU [119], MG53 [120], CHIP [123], cIAP1/cIAP2 [113], LUBAC [115,116], MIB2 [118], A20, CYLD, OTULIN [66,273], Pellino 1, c-Cbl [121,122], and so on, which could be the targets for pharmacological intervention. An increasing number of modification sites are also being found on molecules such as RIPK1, RIPK3, and MLKL, which merit further investigation by researchers. This may drive the development of new targeted drugs to augment this mechanism. Such drugs could potentially delay or treat a broad spectrum of diseases. These include inflammatory, injury-related, immunological, or oncological conditions such as psoriasis, severe rheumatoid arthritis, and Alzheimer’s disease [41].
Secondly, we need to discover more means of PTMs, more modification sites, and their crosstalk in necroptosis. For example, it may be possible to correlate necroptosis with other PTMs, such as neddylation, palmitoylation, and lactylation, which is now a hot research topic. With the continuous development of protein resolution techniques and mass spectrometry, we can easily resolve the structure of the molecules in necroptosis and their PTM sites, so more modification sites are worth discovering and verifying. Interestingly, different PTMs can exist at the same site on the same molecule, for example, RIPK1 K115 is modified by both the E3 ligase PELI1 and the methylation-associated enzymes SIRT1, SIRT2, HAT1, and HAT4 [121,134], and we cannot help but wonder if the two PTMs compete with each other. If so, which PTM dominates during necroptosis? There are also complex relationships between different PTMs at different sites on the same molecule or between different molecules. For example, the O-GlcNAcylation of RIPK1 S331 promotes autophosphorylation of RIPK1 S166 and promotes necroptosis [132]. Disulfide bond formation of RIPK1 C257, C268, and C586 promotes RIPK1 S161 autophosphorylation and enhances its kinase activity [94]. PRMT5 exerts its methylation by interacting directly with RIPK1 and reducing its physical distance from RIPK3 [161]. In addition, we can not only look at PTMs but can also study more at the level of DNA and RNA regulation, which also allows us to understand the mechanism of their occurrence. There is an abundance of studies that have indicated that lncRNAs play an essential role in the initiation and development of necroptosis through protein–protein interactions, transcriptional control, and PTMs undoubtedly. lncRNAs have complicated roles in the regulation of necroptosis pathways, and in the treatment of cancer by necroptosis, the use of lncRNA-based drug technologies is expected to inhibit the progression, spread, and metastasis of cancer by regulating necroptosis [274,275].
Finally, we can explore whether crosstalk is possible with other modes of death via PTMs, such as cellular pyroptosis, ferroptosis, autophagy, pan-apoptosis, etc. Termination of one type of death under certain conditions will result in the transformation of another type of death, and by studying the key molecules involved and their PTMs, this will also facilitate our development of clinical drug targets. For example, phosphorylation of the necroptosis-related pathway molecules RIPK1, RIPK3, and MLKL is increased in acidic pH environments, inhibiting necroptosis and promoting its conversion to apoptosis [276]. Moreover, cell death pathways including apoptosis, necroptosis, pyroptosis, and ferroptosis can have continuous and interconnected roles in specific pathologies and pathophysiological processes. These pathways may be regulated by post-translational modifications (PTMs) [3]. All of these questions concerning necroptosis and its PTMs deserve a considerable amount of further research and demonstration.

Author Contributions

H.X. and Z.H. designed and wrote the manuscript and also generated all the tables and figures; M.X. and X.G. collected and organized the literature; Y.Y., S.Q. and N.R. reviewed and revised the article; C.Z. supervised the entire study, wrote the manuscript, and revised the manuscript. All authors reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by grants from the National Natural Science Foundation of China (82472816 and 82103521 to C Zhou, 82472804 and 82073208 to N Ren, 82072672 to Y Yi, and 82100669 to M Xu), the Sino-German Mobility Program (M-0603 to N Ren), and the Shanghai “Rising Stars of Medical Talents” Youth Development Program (SHWSRS(2024)_070).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

4HBD4-Helical bundle domain
ABINA20 binding and inhibitor of NF-kB
ABIN3A20 binding and inhibitor of NF-kB 3
ACDAccidental cell death
Acetyl-CoAAcetyl coenzyme A
ADAlzheimer’s disease
Akt1AKT serine/threonine kinase 1
AMPKAdenylate-activated protein kinase
AMPKAMP-activated protein kinase
APC11Anaphase-promoting complex subunit 11
APPAmyloid precursor protein
Aur-AAurora-A
CAMK2/CaMKIICalcium/calmodulin-dependent protein kinase II
CB2RCannabinoid receptor 2
CHIPHsp70-interacting protein
cIAP1/2Cellular inhibitor of apoptosis 1 and 2
CK1Casein kinase 1
CKIαCasein kinase I alpha
CSNK1G2Casein kinase 1G2
CTSBCathepsin B
CUL3Cullin 3
CYLDCylindromatosis
DAIDNA-dependent activator of IFN regulatory factors
DAMPsDanger-associated molecular patterns
DDDeath domain
DEDDeath effector domain
DFGAsp-Phe-Gly
DISCDeath-inducing signaling complex
DRDeath receptor
DR6Death Receptor 6
DRP1Dynamin-related protein 1
DSBDouble-strand break
DUBDeubiquitinating enzyme
EPECEnteropathogenic Escherichia coli
EREndoplasmic reticulum
FADDFAS associated via death domain
FHAForkhead-associated
FKBP12FK506-binding protein 12
O-GlcNAcO-Conjugated β-N-acetylglucosamine
HATsHistone acetyltransferases
HBPHexosamine biosynthetic pathway
HDACsHistone deacetylases
HECTD3Homologous to the E6-AP Carboxyl Terminus 3
HOIL-1LHeme-oxidized IRP2 ubiquitin ligase-1
HOIPHOIL-1-interacting protein
HOSCNHypothiocyanic acid
HRDHis-Arg-Asp
Hsp90Heat shock protein 90
IAVInfluenza A virus
IFN-IInterferon type I
IKKIkB kinase
IRF1TNF-IFN regulatory factor 1
iuRIPK1Insoluble RIPK1 complex
KLysine
KATsLysine acetyltransferases
KMTLysine methyltransferase
LPSLipopolysaccharide
LRRK2Leucine-rich repeat kinase 2
LUBACLinear ubiquitin chain assembly complex
M1Methionine
MAPKMitogen-activated protein kinase
MAPKAPK2, MK2MAPK activated protein kinase 2
MG53Mitsugumin 53
MIB2Mind Bomb-2
MIRIMyocardial ischemia–reperfusion injury
MKRN1Makorin ring finger protein 1
MLKLMixed lineage kinase domain-like protein
mtROSMitochondria ROS
MyD88Myeloid differentiation primary response 88
NATsNα-acetyltransferases
NEMONF-κB essential modulator
OASLOligoadenylate synthetase-like
OGAO-GlcNAcase
OGTO-GlcNAc transferase
OTULINOTU deubiquitinase with linear linkage specificity
PARdUPARylation-dependent ubiquitination
PARP5APoly(ADP-ribose) polymerase 5A
PARylationPoly ADP-ribosylation
PBCPrimary biliary cholangitis
PDK 13-Phosphatidylinositol-dependent protein kinase 1
PELI1Pellino E3 ubiquitin protein ligase 1
PGAM5phosphoglycerate mutase family member 5
PIPsPhosphatidylinositol phosphates
PKCProtein kinase C
PKRProtein kinase R
PLK1Polo-like kinase 1
poly(I:C)Polyinosine–polycytidylic acid
PP1γProtein phosphatase 1 γ
Ppm1bProtein phosphatase 1B
PPP1R3GProtein phosphatase 1 regulatory subunit 3G
PRMTProtein arginine methyltransferase
PRRPattern recognition receptor
PTMsPost-translational modifications
RCDRegulated cell death
RDARIPK1 kinase activity-dependent apoptosis
cFLIPLLong isoform of cellular FADD-like interleukin (IL)-1β-converting enzyme (FLICE)-inhibitory protein
RHIMsRIP homotypic interaction motifs
RIPK1Receptor-interacting protein kinase 1
RIPK3Receptor-interacting protein kinase 3
RNF146RING finger protein 146
ROSReactive oxygen species
RSKRibosomal protein S6 kinase A1
SCFSkp1-Cullin-1-F-box
SCFSkp2SCF Cullin-Ring E3 ubiquitin ligase complex
SENP1SUMO-specific protease 1
SEVOSevoflurane
SHARPINSHANK-associated RH domain-interacting protein
SIRTsNAD+-dependent sirtuins
Skp2S-phase kinase-associated protein 2
SMYD2Protein 2 containing SET and MYND structural domains
SPATA2Spermatogenesis associated 2
T3SSType III secretion system
TAB2/3TAK1 Binding Protein 2/3
TAK1TGF activating kinase 1
TAMTyro3, Axl, and Mer
TAX1BP1Tax1-binding protein 1
TBK1TANK binding kinase 1
TMGThiamet-G
TNFTumor necrosis factor
TNFRSFTNF receptor superfamily
TRADDTNFRSF1A associated via death domain
TRAF2TNF receptor associated factor 2
TRIFToll/IL-1 receptor domain-containing adaptor inducing interferon-beta
Trx1Thioredoxin-1
S. typhimuriumSalmonella typhimurium
UbUbiquitin
UBAUbiquitin-associated
UPSUbiquitin proteasome system
USP22Ubiquitin-specific peptidase 22
ZBP1Z-DNA binding protein 1
ZFP91Zinc finger protein 91

References

  1. Steller, H. Mechanisms and genes of cellular suicide. Science 1995, 267, 1445–1449. [Google Scholar] [CrossRef] [PubMed]
  2. Galluzzi, L.; Vitale, I.; Aaronson, S.A.; Abrams, J.M.; Adam, D.; Agostinis, P.; Alnemri, E.S.; Altucci, L.; Amelio, I.; Andrews, D.W.; et al. Molecular mechanisms of cell death: Recommendations of the Nomenclature Committee on Cell Death 2018. Cell Death Differ. 2018, 25, 486–541. [Google Scholar] [CrossRef]
  3. Ai, Y.; Meng, Y.; Yan, B.; Zhou, Q.; Wang, X. The biochemical pathways of apoptotic, necroptotic, pyroptotic, and ferroptotic cell death. Mol. Cell 2024, 84, 170–179. [Google Scholar] [CrossRef]
  4. Newton, K.; Strasser, A.; Kayagaki, N.; Dixit, V.M. Cell death. Cell 2024, 187, 235–256. [Google Scholar] [CrossRef]
  5. Yuan, J.; Ofengeim, D. A guide to cell death pathways. Nat. Rev. Mol. Cell. Biol. 2024, 25, 379–395. [Google Scholar] [CrossRef]
  6. Gao, W.; Wang, X.; Zhou, Y.; Wang, X.; Yu, Y. Autophagy, ferroptosis, pyroptosis, and necroptosis in tumor immunotherapy. Signal Transduct. Target. Ther. 2022, 7, 196. [Google Scholar] [CrossRef]
  7. Xavier, R.J.; Podolsky, D.K. Unravelling the pathogenesis of inflammatory bowel disease. Nature 2007, 448, 427–434. [Google Scholar] [CrossRef] [PubMed]
  8. Iannielli, A.; Bido, S.; Folladori, L.; Segnali, A.; Cancellieri, C.; Maresca, A.; Massimino, L.; Rubio, A.; Morabito, G.; Caporali, L.; et al. Pharmacological Inhibition of Necroptosis Protects from Dopaminergic Neuronal Cell Death in Parkinson’s Disease Models. Cell Rep. 2018, 22, 2066–2079. [Google Scholar] [CrossRef] [PubMed]
  9. Degterev, A.; Huang, Z.; Boyce, M.; Li, Y.; Jagtap, P.; Mizushima, N.; Cuny, G.D.; Mitchison, T.J.; Moskowitz, M.A.; Yuan, J. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for ischemic brain injury. Nat. Chem. Biol. 2005, 1, 112–119. [Google Scholar] [CrossRef]
  10. Luedde, M.; Lutz, M.; Carter, N.; Sosna, J.; Jacoby, C.; Vucur, M.; Gautheron, J.; Roderburg, C.; Borg, N.; Reisinger, F.; et al. RIP3, a kinase promoting necroptotic cell death, mediates adverse remodelling after myocardial infarction. Cardiovasc. Res. 2014, 103, 206–216. [Google Scholar] [CrossRef]
  11. Wang, Q.; Liu, Z.; Ren, J.; Morgan, S.; Assa, C.; Liu, B. Receptor-interacting protein kinase 3 contributes to abdominal aortic aneurysms via smooth muscle cell necrosis and inflammation. Circ. Res. 2015, 116, 600–611. [Google Scholar] [CrossRef] [PubMed]
  12. Matthay, M.A.; Zemans, R.L. The acute respiratory distress syndrome: Pathogenesis and treatment. Annu. Rev. Pathol. 2011, 6, 147–163. [Google Scholar] [CrossRef]
  13. Linkermann, A.; Green, D.R. Necroptosis. N. Engl. J. Med. 2014, 370, 455–465. [Google Scholar] [CrossRef]
  14. Seo, J.; Kim, M.W.; Bae, K.H.; Lee, S.C.; Song, J.; Lee, E.W. The roles of ubiquitination in extrinsic cell death pathways and its implications for therapeutics. Biochem. Pharmacol. 2019, 162, 21–40. [Google Scholar] [CrossRef]
  15. Li, W.; Li, F.; Zhang, X.; Lin, H.K.; Xu, C. Insights into the post-translational modification and its emerging role in shaping the tumor microenvironment. Signal Transduct. Target. Ther. 2021, 6, 422. [Google Scholar] [CrossRef]
  16. Frank, D.; Vince, J.E. Pyroptosis versus necroptosis: Similarities, differences, and crosstalk. Cell Death Differ. 2019, 26, 99–114. [Google Scholar] [CrossRef] [PubMed]
  17. Jiang, X.; Stockwell, B.R.; Conrad, M. Ferroptosis: Mechanisms, biology and role in disease. Nat. Rev. Mol. Cell. Biol. 2021, 22, 266–282. [Google Scholar] [CrossRef] [PubMed]
  18. Weinlich, R.; Oberst, A.; Beere, H.M.; Green, D.R. Necroptosis in development, inflammation and disease. Nat. Rev. Mol. Cell. Biol. 2017, 18, 127–136. [Google Scholar] [CrossRef]
  19. Meier, P.; Legrand, A.J.; Adam, D.; Silke, J. Immunogenic cell death in cancer: Targeting necroptosis to induce antitumour immunity. Nat. Rev. Cancer 2024, 24, 299–315. [Google Scholar] [CrossRef]
  20. Yu, P.; Zhang, X.; Liu, N.; Tang, L.; Peng, C.; Chen, X. Pyroptosis: Mechanisms and diseases. Signal Transduct. Target. Ther. 2021, 6, 128. [Google Scholar] [CrossRef]
  21. Gong, Y.; Fan, Z.; Luo, G.; Yang, C.; Huang, Q.; Fan, K.; Cheng, H.; Jin, K.; Ni, Q.; Yu, X.; et al. The role of necroptosis in cancer biology and therapy. Mol. Cancer 2019, 18, 100. [Google Scholar] [CrossRef] [PubMed]
  22. Stockwell, B.R.; Friedmann Angeli, J.P.; Bayir, H.; Bush, A.I.; Conrad, M.; Dixon, S.J.; Fulda, S.; Gascón, S.; Hatzios, S.K.; Kagan, V.E.; et al. Ferroptosis: A Regulated Cell Death Nexus Linking Metabolism, Redox Biology, and Disease. Cell 2017, 171, 273–285. [Google Scholar] [CrossRef] [PubMed]
  23. Wilson, N.S.; Dixit, V.; Ashkenazi, A. Death receptor signal transducers: Nodes of coordination in immune signaling networks. Nat. Immunol. 2009, 10, 348–355. [Google Scholar] [CrossRef]
  24. Akira, S.; Uematsu, S.; Takeuchi, O. Pathogen recognition and innate immunity. Cell 2006, 124, 783–801. [Google Scholar] [CrossRef]
  25. Meng, Y.; Garnish, S.E.; Davies, K.A.; Black, K.A.; Leis, A.P.; Horne, C.R.; Hildebrand, J.M.; Hoblos, H.; Fitzgibbon, C.; Young, S.N.; et al. Phosphorylation-dependent pseudokinase domain dimerization drives full-length MLKL oligomerization. Nat. Commun. 2023, 14, 6804. [Google Scholar] [CrossRef]
  26. Cai, Z.; Jitkaew, S.; Zhao, J.; Chiang, H.C.; Choksi, S.; Liu, J.; Ward, Y.; Wu, L.G.; Liu, Z.G. Plasma membrane translocation of trimerized MLKL protein is required for TNF-induced necroptosis. Nat. Cell Biol. 2014, 16, 55–65. [Google Scholar] [CrossRef] [PubMed]
  27. Kaczmarek, A.; Vandenabeele, P.; Krysko, D.V. Necroptosis: The release of damage-associated molecular patterns and its physiological relevance. Immunity 2013, 38, 209–223. [Google Scholar] [CrossRef]
  28. Galluzzi, L.; Kepp, O.; Kroemer, G. MLKL regulates necrotic plasma membrane permeabilization. Cell Res. 2014, 24, 139–140. [Google Scholar] [CrossRef]
  29. Dondelinger, Y.; Declercq, W.; Montessuit, S.; Roelandt, R.; Goncalves, A.; Bruggeman, I.; Hulpiau, P.; Weber, K.; Sehon, C.A.; Marquis, R.W.; et al. MLKL compromises plasma membrane integrity by binding to phosphatidylinositol phosphates. Cell Rep. 2014, 7, 971–981. [Google Scholar] [CrossRef]
  30. Dovey, C.M.; Diep, J.; Clarke, B.P.; Hale, A.T.; McNamara, D.E.; Guo, H.; Brown, N.W., Jr.; Cao, J.Y.; Grace, C.R.; Gough, P.J.; et al. MLKL Requires the Inositol Phosphate Code to Execute Necroptosis. Mol. Cell 2018, 70, 936–948.e7. [Google Scholar] [CrossRef]
  31. Chen, X.; Li, W.; Ren, J.; Huang, D.; He, W.T.; Song, Y.; Yang, C.; Li, W.; Zheng, X.; Chen, P.; et al. Translocation of mixed lineage kinase domain-like protein to plasma membrane leads to necrotic cell death. Cell Res. 2014, 24, 105–121. [Google Scholar] [CrossRef]
  32. Wang, Z.; Jiang, H.; Chen, S.; Du, F.; Wang, X. The mitochondrial phosphatase PGAM5 functions at the convergence point of multiple necrotic death pathways. Cell 2012, 148, 228–243. [Google Scholar] [CrossRef]
  33. Zhou, Z.; Han, V.; Han, J. New components of the necroptotic pathway. Protein Cell 2012, 3, 811–817. [Google Scholar] [CrossRef] [PubMed]
  34. Kang, J.S.; Cho, N.J.; Lee, S.W.; Lee, J.G.; Lee, J.H.; Yi, J.; Choi, M.S.; Park, S.; Gil, H.W.; Oh, J.C.; et al. RIPK3 causes mitochondrial dysfunction and albuminuria in diabetic podocytopathy through PGAM5-Drp1 signaling. Metabolism 2024, 159, 155982. [Google Scholar] [CrossRef]
  35. Estaquier, J.; Arnoult, D. Inhibiting Drp1-mediated mitochondrial fission selectively prevents the release of cytochrome c during apoptosis. Cell Death Differ. 2007, 14, 1086–1094. [Google Scholar] [CrossRef] [PubMed]
  36. Li, Y.; Chen, H.; Yang, Q.; Wan, L.; Zhao, J.; Wu, Y.; Wang, J.; Yang, Y.; Niu, M.; Liu, H.; et al. Increased Drp1 promotes autophagy and ESCC progression by mtDNA stress mediated cGAS-STING pathway. J. Exp. Clin. Cancer Res. 2022, 41, 76. [Google Scholar] [CrossRef] [PubMed]
  37. Gong, Y.N.; Guy, C.; Olauson, H.; Becker, J.U.; Yang, M.; Fitzgerald, P.; Linkermann, A.; Green, D.R. ESCRT-III Acts Downstream of MLKL to Regulate Necroptotic Cell Death and Its Consequences. Cell 2017, 169, 286–300.e16. [Google Scholar] [CrossRef]
  38. Rothe, J.; Lesslauer, W.; Lötscher, H.; Lang, Y.; Koebel, P.; Köntgen, F.; Althage, A.; Zinkernagel, R.; Steinmetz, M.; Bluethmann, H. Mice lacking the tumour necrosis factor receptor 1 are resistant to TNF-mediated toxicity but highly susceptible to infection by Listeria monocytogenes. Nature 1993, 364, 798–802. [Google Scholar] [CrossRef]
  39. Kontoyiannis, D.; Pasparakis, M.; Pizarro, T.T.; Cominelli, F.; Kollias, G. Impaired on/off regulation of TNF biosynthesis in mice lacking TNF AU-rich elements: Implications for joint and gut-associated immunopathologies. Immunity 1999, 10, 387–398. [Google Scholar] [CrossRef]
  40. Xanthoulea, S.; Pasparakis, M.; Kousteni, S.; Brakebusch, C.; Wallach, D.; Bauer, J.; Lassmann, H.; Kollias, G. Tumor necrosis factor (TNF) receptor shedding controls thresholds of innate immune activation that balance opposing TNF functions in infectious and inflammatory diseases. J. Exp. Med. 2004, 200, 367–376. [Google Scholar] [CrossRef]
  41. Seo, J.; Nam, Y.W.; Kim, S.; Oh, D.B.; Song, J. Necroptosis molecular mechanisms: Recent findings regarding novel necroptosis regulators. Exp. Mol. Med. 2021, 53, 1007–1017. [Google Scholar] [CrossRef]
  42. Bertrand, M.J.; Milutinovic, S.; Dickson, K.M.; Ho, W.C.; Boudreault, A.; Durkin, J.; Gillard, J.W.; Jaquith, J.B.; Morris, S.J.; Barker, P.A. cIAP1 and cIAP2 facilitate cancer cell survival by functioning as E3 ligases that promote RIP1 ubiquitination. Mol. Cell 2008, 30, 689–700. [Google Scholar] [CrossRef] [PubMed]
  43. Peltzer, N.; Darding, M.; Walczak, H. Holding RIPK1 on the Ubiquitin Leash in TNFR1 Signaling. Trends Cell Biol. 2016, 26, 445–461. [Google Scholar] [CrossRef]
  44. Tokunaga, F.; Sakata, S.; Saeki, Y.; Satomi, Y.; Kirisako, T.; Kamei, K.; Nakagawa, T.; Kato, M.; Murata, S.; Yamaoka, S.; et al. Involvement of linear polyubiquitylation of NEMO in NF-kappaB activation. Nat. Cell Biol. 2009, 11, 123–132. [Google Scholar] [CrossRef] [PubMed]
  45. Haas, T.L.; Emmerich, C.H.; Gerlach, B.; Schmukle, A.C.; Cordier, S.M.; Rieser, E.; Feltham, R.; Vince, J.; Warnken, U.; Wenger, T.; et al. Recruitment of the linear ubiquitin chain assembly complex stabilizes the TNF-R1 signaling complex and is required for TNF-mediated gene induction. Mol. Cell 2009, 36, 831–844. [Google Scholar] [CrossRef]
  46. Gerlach, B.; Cordier, S.M.; Schmukle, A.C.; Emmerich, C.H.; Rieser, E.; Haas, T.L.; Webb, A.I.; Rickard, J.A.; Anderton, H.; Wong, W.W.; et al. Linear ubiquitination prevents inflammation and regulates immune signalling. Nature 2011, 471, 591–596. [Google Scholar] [CrossRef] [PubMed]
  47. Ikeda, F.; Deribe, Y.L.; Skånland, S.S.; Stieglitz, B.; Grabbe, C.; Franz-Wachtel, M.; van Wijk, S.J.; Goswami, P.; Nagy, V.; Terzic, J.; et al. SHARPIN forms a linear ubiquitin ligase complex regulating NF-κB activity and apoptosis. Nature 2011, 471, 637–641. [Google Scholar] [CrossRef]
  48. Silke, J.; Brink, R. Regulation of TNFRSF and innate immune signalling complexes by TRAFs and cIAPs. Cell Death Differ. 2010, 17, 35–45. [Google Scholar] [CrossRef]
  49. Dondelinger, Y.; Delanghe, T.; Rojas-Rivera, D.; Priem, D.; Delvaeye, T.; Bruggeman, I.; Van Herreweghe, F.; Vandenabeele, P.; Bertrand, M.J.M. MK2 phosphorylation of RIPK1 regulates TNF-mediated cell death. Nat. Cell Biol. 2017, 19, 1237–1247. [Google Scholar] [CrossRef]
  50. Jaco, I.; Annibaldi, A.; Lalaoui, N.; Wilson, R.; Tenev, T.; Laurien, L.; Kim, C.; Jamal, K.; Wicky John, S.; Liccardi, G.; et al. MK2 Phosphorylates RIPK1 to Prevent TNF-Induced Cell Death. Mol. Cell 2017, 66, 698–710.e5. [Google Scholar] [CrossRef]
  51. Menon, M.B.; Gropengießer, J.; Fischer, J.; Novikova, L.; Deuretzbacher, A.; Lafera, J.; Schimmeck, H.; Czymmeck, N.; Ronkina, N.; Kotlyarov, A.; et al. p38(MAPK)/MK2-dependent phosphorylation controls cytotoxic RIPK1 signalling in inflammation and infection. Nat. Cell Biol. 2017, 19, 1248–1259. [Google Scholar] [CrossRef]
  52. Tu, H.; Xiong, W.; Zhang, J.; Zhao, X.; Lin, X. Tyrosine phosphorylation regulates RIPK1 activity to limit cell death and inflammation. Nat. Commun. 2022, 13, 6603. [Google Scholar] [CrossRef] [PubMed]
  53. Wu, W.; Wang, X.; Berleth, N.; Deitersen, J.; Wallot-Hieke, N.; Böhler, P.; Schlütermann, D.; Stuhldreier, F.; Cox, J.; Schmitz, K.; et al. The Autophagy-Initiating Kinase ULK1 Controls RIPK1-Mediated Cell Death. Cell Rep. 2020, 31, 107547. [Google Scholar] [CrossRef]
  54. Xu, D.; Jin, T.; Zhu, H.; Chen, H.; Ofengeim, D.; Zou, C.; Mifflin, L.; Pan, L.; Amin, P.; Li, W.; et al. TBK1 Suppresses RIPK1-Driven Apoptosis and Inflammation during Development and in Aging. Cell 2018, 174, 1477–1491.e19. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, T.; Xu, D.; Trefts, E.; Lv, M.; Inuzuka, H.; Song, G.; Liu, M.; Lu, J.; Liu, J.; Chu, C.; et al. Metabolic orchestration of cell death by AMPK-mediated phosphorylation of RIPK1. Science 2023, 380, 1372–1380. [Google Scholar] [CrossRef]
  56. Meng, H.; Liu, Z.; Li, X.; Wang, H.; Jin, T.; Wu, G.; Shan, B.; Christofferson, D.E.; Qi, C.; Yu, Q.; et al. Death-domain dimerization-mediated activation of RIPK1 controls necroptosis and RIPK1-dependent apoptosis. Proc. Natl. Acad. Sci. USA 2018, 115, E2001–E2009. [Google Scholar] [CrossRef]
  57. Petersen, S.L.; Chen, T.T.; Lawrence, D.A.; Marsters, S.A.; Gonzalvez, F.; Ashkenazi, A. TRAF2 is a biologically important necroptosis suppressor. Cell Death Differ. 2015, 22, 1846–1857. [Google Scholar] [CrossRef]
  58. Wang, L.; Du, F.; Wang, X. TNF-alpha induces two distinct caspase-8 activation pathways. Cell 2008, 133, 693–703. [Google Scholar] [CrossRef] [PubMed]
  59. Newton, K. Multitasking Kinase RIPK1 Regulates Cell Death and Inflammation. Cold Spring Harb. Perspect. Biol. 2020, 12, a036368. [Google Scholar] [CrossRef]
  60. Priem, D.; Devos, M.; Druwé, S.; Martens, A.; Slowicka, K.; Ting, A.T.; Pasparakis, M.; Declercq, W.; Vandenabeele, P.; van Loo, G.; et al. A20 protects cells from TNF-induced apoptosis through linear ubiquitin-dependent and -independent mechanisms. Cell Death Dis. 2019, 10, 692. [Google Scholar] [CrossRef]
  61. Dziedzic, S.A.; Su, Z.; Jean Barrett, V.; Najafov, A.; Mookhtiar, A.K.; Amin, P.; Pan, H.; Sun, L.; Zhu, H.; Ma, A.; et al. ABIN-1 regulates RIPK1 activation by linking Met1 ubiquitylation with Lys63 deubiquitylation in TNF-RSC. Nat. Cell Biol. 2018, 20, 58–68. [Google Scholar] [CrossRef]
  62. Wertz, I.E.; O’Rourke, K.M.; Zhou, H.; Eby, M.; Aravind, L.; Seshagiri, S.; Wu, P.; Wiesmann, C.; Baker, R.; Boone, D.L.; et al. De-ubiquitination and ubiquitin ligase domains of A20 downregulate NF-kappaB signalling. Nature 2004, 430, 694–699. [Google Scholar] [CrossRef]
  63. Dou, Z.; Bonacci, T.R.; Shou, P.; Landoni, E.; Woodcock, M.G.; Sun, C.; Savoldo, B.; Herring, L.E.; Emanuele, M.J.; Song, F.; et al. 4-1BB-encoding CAR causes cell death via sequestration of the ubiquitin-modifying enzyme A20. Cell. Mol. Immunol. 2024, 21, 905–917. [Google Scholar] [CrossRef] [PubMed]
  64. Li, M.; Liu, Y.; Xu, C.; Zhao, Q.; Liu, J.; Xing, M.; Li, X.; Zhang, H.; Wu, X.; Wang, L.; et al. Ubiquitin-binding domain in ABIN1 is critical for regulating cell death and inflammation during development. Cell Death Differ. 2022, 29, 2034–2045. [Google Scholar] [CrossRef] [PubMed]
  65. Enesa, K.; Zakkar, M.; Chaudhury, H.; Luong, L.A.; Rawlinson, L.; Mason, J.C.; Haskard, D.O.; Dean, J.L.; Evans, P.C. NF-kappaB suppression by the deubiquitinating enzyme Cezanne: A novel negative feedback loop in pro-inflammatory signaling. J. Biol. Chem. 2008, 283, 7036–7045. [Google Scholar] [CrossRef]
  66. Draber, P.; Kupka, S.; Reichert, M.; Draberova, H.; Lafont, E.; de Miguel, D.; Spilgies, L.; Surinova, S.; Taraborrelli, L.; Hartwig, T.; et al. LUBAC-Recruited CYLD and A20 Regulate Gene Activation and Cell Death by Exerting Opposing Effects on Linear Ubiquitin in Signaling Complexes. Cell Rep. 2015, 13, 2258–2272. [Google Scholar] [CrossRef] [PubMed]
  67. Xinyu, W.; Qian, W.; Yanjun, W.; Jingwen, K.; Keying, X.; Jiazheng, J.; Haibing, Z.; Kai, W.; Xiao, X.; Lixing, Z. Polarity protein AF6 functions as a modulator of necroptosis by regulating ubiquitination of RIPK1 in liver diseases. Cell Death Dis. 2023, 14, 673. [Google Scholar] [CrossRef]
  68. Zhong, Y.; Zhang, Z.H.; Wang, J.Y.; Xing, Y.; Ri, M.H.; Jin, H.L.; Zuo, H.X.; Li, M.Y.; Ma, J.; Jin, X. Zinc finger protein 91 mediates necroptosis by initiating RIPK1-RIPK3-MLKL signal transduction in response to TNF receptor 1 ligation. Toxicol. Lett. 2022, 356, 75–88. [Google Scholar] [CrossRef]
  69. He, S.; Wang, X. RIP kinases as modulators of inflammation and immunity. Nat. Immunol. 2018, 19, 912–922. [Google Scholar] [CrossRef]
  70. Micheau, O.; Tschopp, J. Induction of TNF receptor I-mediated apoptosis via two sequential signaling complexes. Cell 2003, 114, 181–190. [Google Scholar] [CrossRef]
  71. Fu, T.M.; Li, Y.; Lu, A.; Li, Z.; Vajjhala, P.R.; Cruz, A.C.; Srivastava, D.B.; DiMaio, F.; Penczek, P.A.; Siegel, R.M.; et al. Cryo-EM Structure of Caspase-8 Tandem DED Filament Reveals Assembly and Regulation Mechanisms of the Death-Inducing Signaling Complex. Mol. Cell 2016, 64, 236–250. [Google Scholar] [CrossRef] [PubMed]
  72. Fox, J.L.; Hughes, M.A.; Meng, X.; Sarnowska, N.A.; Powley, I.R.; Jukes-Jones, R.; Dinsdale, D.; Ragan, T.J.; Fairall, L.; Schwabe, J.W.R.; et al. Cryo-EM structural analysis of FADD:Caspase-8 complexes defines the catalytic dimer architecture for co-ordinated control of cell fate. Nat. Commun. 2021, 12, 819. [Google Scholar] [CrossRef] [PubMed]
  73. Varfolomeev, E.; Maecker, H.; Sharp, D.; Lawrence, D.; Renz, M.; Vucic, D.; Ashkenazi, A. Molecular determinants of kinase pathway activation by Apo2 ligand/tumor necrosis factor-related apoptosis-inducing ligand. J. Biol. Chem. 2005, 280, 40599–40608. [Google Scholar] [CrossRef] [PubMed]
  74. Hartwig, T.; Montinaro, A.; von Karstedt, S.; Sevko, A.; Surinova, S.; Chakravarthy, A.; Taraborrelli, L.; Draber, P.; Lafont, E.; Arce Vargas, F.; et al. The TRAIL-Induced Cancer Secretome Promotes a Tumor-Supportive Immune Microenvironment via CCR2. Mol. Cell 2017, 65, 730–742.e5. [Google Scholar] [CrossRef]
  75. Holler, N.; Zaru, R.; Micheau, O.; Thome, M.; Attinger, A.; Valitutti, S.; Bodmer, J.L.; Schneider, P.; Seed, B.; Tschopp, J. Fas triggers an alternative, caspase-8-independent cell death pathway using the kinase RIP as effector molecule. Nat. Immunol. 2000, 1, 489–495. [Google Scholar] [CrossRef]
  76. Henry, C.M.; Martin, S.J. Caspase-8 Acts in a Non-enzymatic Role as a Scaffold for Assembly of a Pro-inflammatory “FADDosome” Complex upon TRAIL Stimulation. Mol. Cell 2017, 65, 715–729.e5. [Google Scholar] [CrossRef]
  77. Migone, T.S.; Zhang, J.; Luo, X.; Zhuang, L.; Chen, C.; Hu, B.; Hong, J.S.; Perry, J.W.; Chen, S.F.; Zhou, J.X.; et al. TL1A is a TNF-like ligand for DR3 and TR6/DcR3 and functions as a T cell costimulator. Immunity 2002, 16, 479–492. [Google Scholar] [CrossRef]
  78. Bittner, S.; Knoll, G.; Ehrenschwender, M. Death receptor 3 mediates necroptotic cell death. Cell. Mol. Life Sci. 2017, 74, 543–554. [Google Scholar] [CrossRef]
  79. Liu, X.; Zhang, J.; Zhang, D.; Pan, Y.; Zeng, R.; Xu, C.; Shi, S.; Xu, J.; Qi, Q.; Dong, X.; et al. Necroptosis plays a role in TL1A-induced airway inflammation and barrier damage in asthma. Respir. Res. 2024, 25, 271. [Google Scholar] [CrossRef]
  80. Hassan-Zahraee, M.; Ye, Z.; Xi, L.; Baniecki, M.L.; Li, X.; Hyde, C.L.; Zhang, J.; Raha, N.; Karlsson, F.; Quan, J.; et al. Antitumor Necrosis Factor-like Ligand 1A Therapy Targets Tissue Inflammation and Fibrosis Pathways and Reduces Gut Pathobionts in Ulcerative Colitis. Inflamm. Bowel Dis. 2022, 28, 434–446. [Google Scholar] [CrossRef]
  81. Pan, G.; Bauer, J.H.; Haridas, V.; Wang, S.; Liu, D.; Yu, G.; Vincenz, C.; Aggarwal, B.B.; Ni, J.; Dixit, V.M. Identification and functional characterization of DR6, a novel death domain-containing TNF receptor. FEBS Lett. 1998, 431, 351–356. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, L.; Shen, Q.; Liao, H.; Fu, H.; Wang, Q.; Yu, J.; Zhang, W.; Chen, C.; Dong, Y.; Yang, X.; et al. Multi-Arm PEG/Peptidomimetic Conjugate Inhibitors of DR6/APP Interaction Block Hematogenous Tumor Cell Extravasation. Adv. Sci. 2021, 8, e2003558. [Google Scholar] [CrossRef]
  83. Strilic, B.; Yang, L.; Albarrán-Juárez, J.; Wachsmuth, L.; Han, K.; Müller, U.C.; Pasparakis, M.; Offermanns, S. Tumour-cell-induced endothelial cell necroptosis via death receptor 6 promotes metastasis. Nature 2016, 536, 215–218. [Google Scholar] [CrossRef]
  84. He, S.; Liang, Y.; Shao, F.; Wang, X. Toll-like receptors activate programmed necrosis in macrophages through a receptor-interacting kinase-3-mediated pathway. Proc. Natl. Acad. Sci. USA 2011, 108, 20054–20059. [Google Scholar] [CrossRef] [PubMed]
  85. Kaiser, W.J.; Sridharan, H.; Huang, C.; Mandal, P.; Upton, J.W.; Gough, P.J.; Sehon, C.A.; Marquis, R.W.; Bertin, J.; Mocarski, E.S. Toll-like receptor 3-mediated necrosis via TRIF, RIP3, and MLKL. J. Biol. Chem. 2013, 288, 31268–31279. [Google Scholar] [CrossRef] [PubMed]
  86. Grootjans, S.; Vanden Berghe, T.; Vandenabeele, P. Initiation and execution mechanisms of necroptosis: An overview. Cell Death Differ. 2017, 24, 1184–1195. [Google Scholar] [CrossRef]
  87. Polykratis, A.; Hermance, N.; Zelic, M.; Roderick, J.; Kim, C.; Van, T.M.; Lee, T.H.; Chan, F.K.M.; Pasparakis, M.; Kelliher, M.A. Cutting edge: RIPK1 Kinase inactive mice are viable and protected from TNF-induced necroptosis in vivo. J. Immunol. 2014, 193, 1539–1543. [Google Scholar] [CrossRef]
  88. 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 2019, 26, 564. [Google Scholar] [CrossRef]
  89. Koerner, L.; Wachsmuth, L.; Kumari, S.; Schwarzer, R.; Wagner, T.; Jiao, H.; Pasparakis, M. ZBP1 causes inflammation by inducing RIPK3-mediated necroptosis and RIPK1 kinase activity-independent apoptosis. Cell Death Differ. 2024, 31, 938–953. [Google Scholar] [CrossRef]
  90. Zhang, T.; Yin, C.; Boyd, D.F.; Quarato, G.; Ingram, J.P.; Shubina, M.; Ragan, K.B.; Ishizuka, T.; Crawford, J.C.; Tummers, B.; et al. Influenza Virus Z-RNAs Induce ZBP1-Mediated Necroptosis. Cell 2020, 180, 1115–1129.e13. [Google Scholar] [CrossRef]
  91. Lin, J.; Kumari, S.; Kim, C.; Van, T.M.; Wachsmuth, L.; Polykratis, A.; Pasparakis, M. RIPK1 counteracts ZBP1-mediated necroptosis to inhibit inflammation. Nature 2016, 540, 124–128. [Google Scholar] [CrossRef] [PubMed]
  92. Imai, T.; Lin, J.; Kaya, G.G.; Ju, E.; Kondylis, V.; Kelepouras, K.; Liccardi, G.; Kim, C.; Pasparakis, M. The RIPK1 death domain restrains ZBP1- and TRIF-mediated cell death and inflammation. Immunity 2024, 57, 1497–1513.e6. [Google Scholar] [CrossRef]
  93. Byun, H.S.; Ju, E.; Park, K.A.; Sohn, K.C.; Jung, C.S.; Hong, J.H.; Ro, H.; Lee, H.Y.; Quan, K.T.; Park, I.; et al. Rubiarbonol B induces RIPK1-dependent necroptosis via NOX1-derived ROS production. Cell Biol. Toxicol. 2023, 39, 1677–1696. [Google Scholar] [CrossRef]
  94. Zhang, Y.; Su, S.S.; Zhao, S.; Yang, Z.; Zhong, C.Q.; Chen, X.; Cai, Q.; Yang, Z.H.; Huang, D.; Wu, R.; et al. RIP1 autophosphorylation is promoted by mitochondrial ROS and is essential for RIP3 recruitment into necrosome. Nat. Commun. 2017, 8, 14329. [Google Scholar] [CrossRef]
  95. Kim, S.; Lee, H.; Lim, J.W.; Kim, H. Astaxanthin induces NADPH oxidase activation and receptor-interacting protein kinase 1-mediated necroptosis in gastric cancer AGS cells. Mol. Med. Rep. 2021, 24, 837. [Google Scholar] [CrossRef] [PubMed]
  96. Li, L.; Tan, H.; Zou, Z.; Gong, J.; Zhou, J.; Peng, N.; Su, L.; Maegele, M.; Cai, D.; Gu, Z. Preventing necroptosis by scavenging ROS production alleviates heat stress-induced intestinal injury. Int. J. Hyperth. 2020, 37, 517–530. [Google Scholar] [CrossRef] [PubMed]
  97. Zhao, W.; Feng, H.; Sun, W.; Liu, K.; Lu, J.J.; Chen, X. Tert-butyl hydroperoxide (t-BHP) induced apoptosis and necroptosis in endothelial cells: Roles of NOX4 and mitochondrion. Redox Biol. 2017, 11, 524–534. [Google Scholar] [CrossRef]
  98. Fan, Y.; Lu, J.; Yu, Z.; Qu, X.; Guan, S. 1,3-Dichloro-2-propanol-Induced Renal Tubular Cell Necroptosis through the ROS/RIPK3/MLKL Pathway. J. Agric. Food Chem. 2022, 70, 10847–10857. [Google Scholar] [CrossRef]
  99. Huang, Y.T.; Liang, Q.Q.; Zhang, H.R.; Chen, S.Y.; Xu, L.H.; Zeng, B.; Xu, R.; Shi, F.L.; Ouyang, D.Y.; Zha, Q.B.; et al. Baicalin inhibits necroptosis by decreasing oligomerization of phosphorylated MLKL and mitigates caerulein-induced acute pancreatitis in mice. Int. Immunopharmacol. 2022, 108, 108885. [Google Scholar] [CrossRef]
  100. Du, J.; Zhang, X.; Zhang, J.; Huo, S.; Li, B.; Wang, Q.; Song, M.; Shao, B.; Li, Y. Necroptosis and NLPR3 inflammasome activation mediated by ROS/JNK pathway participate in AlCl(3)-induced kidney damage. Food Chem. Toxicol. 2023, 178, 113915. [Google Scholar] [CrossRef]
  101. Wang, K.J.; Meng, X.Y.; Chen, J.F.; Wang, K.Y.; Zhou, C.; Yu, R.; Ma, Q. Emodin Induced Necroptosis and Inhibited Glycolysis in the Renal Cancer Cells by Enhancing ROS. Oxidative Med. Cell. Longev. 2021, 2021, 8840590. [Google Scholar] [CrossRef] [PubMed]
  102. Zhou, Z.; Lu, B.; Wang, C.; Wang, Z.; Luo, T.; Piao, M.; Meng, F.; Chi, G.; Luo, Y.; Ge, P. RIP1 and RIP3 contribute to shikonin-induced DNA double-strand breaks in glioma cells via increase of intracellular reactive oxygen species. Cancer Lett. 2017, 390, 77–90. [Google Scholar] [CrossRef]
  103. Jia, Y.; Wang, F.; Guo, Q.; Li, M.; Wang, L.; Zhang, Z.; Jiang, S.; Jin, H.; Chen, A.; Tan, S.; et al. Curcumol induces RIPK1/RIPK3 complex-dependent necroptosis via JNK1/2-ROS signaling in hepatic stellate cells. Redox Biol. 2018, 19, 375–387. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, L.; Chang, X.; Feng, J.; Yu, J.; Chen, G. TRADD Mediates RIPK1-Independent Necroptosis Induced by Tumor Necrosis Factor. Front. Cell Dev. Biol. 2019, 7, 393. [Google Scholar] [CrossRef] [PubMed]
  105. Weindel, C.G.; Martinez, E.L.; Zhao, X.; Mabry, C.J.; Bell, S.L.; Vail, K.J.; Coleman, A.K.; VanPortfliet, J.J.; Zhao, B.; Wagner, A.R.; et al. Mitochondrial ROS promotes susceptibility to infection via gasdermin D-mediated necroptosis. Cell 2022, 185, 3214–3231.e23. [Google Scholar] [CrossRef]
  106. Thapa, R.J.; Nogusa, S.; Chen, P.; Maki, J.L.; Lerro, A.; Andrake, M.; Rall, G.F.; Degterev, A.; Balachandran, S. Interferon-induced RIP1/RIP3-mediated necrosis requires PKR and is licensed by FADD and caspases. Proc. Natl. Acad. Sci. USA 2013, 110, E3109–E3118. [Google Scholar] [CrossRef]
  107. Schock, S.N.; Chandra, N.V.; Sun, Y.; Irie, T.; Kitagawa, Y.; Gotoh, B.; Coscoy, L.; Winoto, A. Induction of necroptotic cell death by viral activation of the RIG-I or STING pathway. Cell Death Differ. 2017, 24, 615–625. [Google Scholar] [CrossRef]
  108. Yang, Y.; Wu, M.; Cao, D.; Yang, C.; Jin, J.; Wu, L.; Hong, X.; Li, W.; Lu, L.; Li, J.; et al. ZBP1-MLKL necroptotic signaling potentiates radiation-induced antitumor immunity via intratumoral STING pathway activation. Sci. Adv. 2021, 7, eabf6290. [Google Scholar] [CrossRef]
  109. McComb, S.; Cessford, E.; Alturki, N.A.; Joseph, J.; Shutinoski, B.; Startek, J.B.; Gamero, A.M.; Mossman, K.L.; Sad, S. Type-I interferon signaling through ISGF3 complex is required for sustained Rip3 activation and necroptosis in macrophages. Proc. Natl. Acad. Sci. USA 2014, 111, E3206–E3213. [Google Scholar] [CrossRef]
  110. Hershko, A. Lessons from the discovery of the ubiquitin system. Trends Biochem. Sci. 1996, 21, 445–449. [Google Scholar] [CrossRef]
  111. Datta, A.B.; Hura, G.L.; Wolberger, C. The structure and conformation of Lys63-linked tetraubiquitin. J. Mol. Biol. 2009, 392, 1117–1124. [Google Scholar] [CrossRef]
  112. Emmerich, C.H.; Bakshi, S.; Kelsall, I.R.; Ortiz-Guerrero, J.; Shpiro, N.; Cohen, P. Lys63/Met1-hybrid ubiquitin chains are commonly formed during the activation of innate immune signalling. Biochem. Biophys. Res. Commun. 2016, 474, 452–461. [Google Scholar] [CrossRef] [PubMed]
  113. Annibaldi, A.; Wicky John, S.; Vanden Berghe, T.; Swatek, K.N.; Ruan, J.; Liccardi, G.; Bianchi, K.; Elliott, P.R.; Choi, S.M.; Van Coillie, S.; et al. Ubiquitin-Mediated Regulation of RIPK1 Kinase Activity Independent of IKK and MK2. Mol. Cell 2018, 69, 566–580.e5. [Google Scholar] [CrossRef]
  114. Kist, M.; Kőműves, L.G.; Goncharov, T.; Dugger, D.L.; Yu, C.; Roose-Girma, M.; Newton, K.; Webster, J.D.; Vucic, D. Impaired RIPK1 ubiquitination sensitizes mice to TNF toxicity and inflammatory cell death. Cell Death Differ. 2021, 28, 985–1000. [Google Scholar] [CrossRef] [PubMed]
  115. Taraborrelli, L.; Peltzer, N.; Montinaro, A.; Kupka, S.; Rieser, E.; Hartwig, T.; Sarr, A.; Darding, M.; Draber, P.; Haas, T.L.; et al. LUBAC prevents lethal dermatitis by inhibiting cell death induced by TNF, TRAIL and CD95L. Nat. Commun. 2018, 9, 3910. [Google Scholar] [CrossRef]
  116. Peltzer, N.; Darding, M.; Montinaro, A.; Draber, P.; Draberova, H.; Kupka, S.; Rieser, E.; Fisher, A.; Hutchinson, C.; Taraborrelli, L.; et al. LUBAC is essential for embryogenesis by preventing cell death and enabling haematopoiesis. Nature 2018, 557, 112–117. [Google Scholar] [CrossRef] [PubMed]
  117. Wu, G.; Li, D.; Liang, W.; Sun, W.; Xie, X.; Tong, Y.; Shan, B.; Zhang, M.; Lu, X.; Yuan, J.; et al. PP6 negatively modulates LUBAC-mediated M1-ubiquitination of RIPK1 and c-FLIP(L) to promote TNFα-mediated cell death. Cell Death Dis. 2022, 13, 773. [Google Scholar] [CrossRef]
  118. Feltham, R.; Jamal, K.; Tenev, T.; Liccardi, G.; Jaco, I.; Domingues, C.M.; Morris, O.; John, S.W.; Annibaldi, A.; Widya, M.; et al. Mind Bomb Regulates Cell Death during TNF Signaling by Suppressing RIPK1’s Cytotoxic Potential. Cell Rep. 2018, 23, 470–484. [Google Scholar] [CrossRef]
  119. Hou, S.; Zhang, J.; Jiang, X.; Yang, Y.; Shan, B.; Zhang, M.; Liu, C.; Yuan, J.; Xu, D. PARP5A and RNF146 phase separation restrains RIPK1-dependent necroptosis. Mol. Cell 2024, 84, 938–954.e8. [Google Scholar] [CrossRef]
  120. Wang, Q.; Park, K.H.; Geng, B.; Chen, P.; Yang, C.; Jiang, Q.; Yi, F.; Tan, T.; Zhou, X.; Bian, Z.; et al. MG53 Inhibits Necroptosis Through Ubiquitination-Dependent RIPK1 Degradation for Cardiac Protection Following Ischemia/Reperfusion Injury. Front. Cardiovasc. Med. 2022, 9, 868632. [Google Scholar] [CrossRef]
  121. Wang, H.; Meng, H.; Li, X.; Zhu, K.; Dong, K.; Mookhtiar, A.K.; Wei, H.; Li, Y.; Sun, S.C.; Yuan, J. PELI1 functions as a dual modulator of necroptosis and apoptosis by regulating ubiquitination of RIPK1 and mRNA levels of c-FLIP. Proc. Natl. Acad. Sci. USA 2017, 114, 11944–11949. [Google Scholar] [CrossRef] [PubMed]
  122. Amin, P.; Florez, M.; Najafov, A.; Pan, H.; Geng, J.; Ofengeim, D.; Dziedzic, S.A.; Wang, H.; Barrett, V.J.; Ito, Y.; et al. Regulation of a distinct activated RIPK1 intermediate bridging complex I and complex II in TNFα-mediated apoptosis. Proc. Natl. Acad. Sci. USA 2018, 115, E5944–E5953. [Google Scholar] [CrossRef]
  123. Seo, J.; Lee, E.W.; Sung, H.; Seong, D.; Dondelinger, Y.; Shin, J.; Jeong, M.; Lee, H.K.; Kim, J.H.; Han, S.Y.; et al. CHIP controls necroptosis through ubiquitylation- and lysosome-dependent degradation of RIPK3. Nat. Cell Biol. 2016, 18, 291–302. [Google Scholar] [CrossRef]
  124. Li, X.; Zhang, M.; Huang, X.; Liang, W.; Li, G.; Lu, X.; Li, Y.; Pan, H.; Shi, L.; Zhu, H.; et al. Ubiquitination of RIPK1 regulates its activation mediated by TNFR1 and TLRs signaling in distinct manners. Nat. Commun. 2020, 11, 6364. [Google Scholar] [CrossRef] [PubMed]
  125. Degterev, A.; Hitomi, J.; Germscheid, M.; Ch’en, I.L.; Korkina, O.; Teng, X.; Abbott, D.; Cuny, G.D.; Yuan, C.; Wagner, G.; et al. Identification of RIP1 kinase as a specific cellular target of necrostatins. Nat. Chem. Biol. 2008, 4, 313–321. [Google Scholar] [CrossRef]
  126. Laurien, L.; Nagata, M.; Schünke, H.; Delanghe, T.; Wiederstein, J.L.; Kumari, S.; Schwarzer, R.; Corona, T.; Krüger, M.; Bertrand, M.J.M.; et al. Autophosphorylation at serine 166 regulates RIP kinase 1-mediated cell death and inflammation. Nat. Commun. 2020, 11, 1747. [Google Scholar] [CrossRef] [PubMed]
  127. Dondelinger, Y.; Delanghe, T.; Priem, D.; Wynosky-Dolfi, M.A.; Sorobetea, D.; Rojas-Rivera, D.; Giansanti, P.; Roelandt, R.; Gropengiesser, J.; Ruckdeschel, K.; et al. Serine 25 phosphorylation inhibits RIPK1 kinase-dependent cell death in models of infection and inflammation. Nat. Commun. 2019, 10, 1729. [Google Scholar] [CrossRef]
  128. Geng, J.; Ito, Y.; Shi, L.; Amin, P.; Chu, J.; Ouchida, A.T.; Mookhtiar, A.K.; Zhao, H.; Xu, D.; Shan, B.; et al. Regulation of RIPK1 activation by TAK1-mediated phosphorylation dictates apoptosis and necroptosis. Nat. Commun. 2017, 8, 359. [Google Scholar] [CrossRef]
  129. McQuade, T.; Cho, Y.; Chan, F.K. Positive and negative phosphorylation regulates RIP1- and RIP3-induced programmed necrosis. Biochem. J. 2013, 456, 409–415. [Google Scholar] [CrossRef]
  130. Du, J.; Xiang, Y.; Liu, H.; Liu, S.; Kumar, A.; Xing, C.; Wang, Z. RIPK1 dephosphorylation and kinase activation by PPP1R3G/PP1γ promote apoptosis and necroptosis. Nat. Commun. 2021, 12, 7067. [Google Scholar] [CrossRef]
  131. Yu, Y.Q.; Thonn, V.; Patankar, J.V.; Thoma, O.M.; Waldner, M.; Zielinska, M.; Bao, L.L.; Gonzalez-Acera, M.; Wallmüller, S.; Engel, F.B.; et al. SMYD2 targets RIPK1 and restricts TNF-induced apoptosis and necroptosis to support colon tumor growth. Cell Death Dis. 2022, 13, 52. [Google Scholar] [CrossRef] [PubMed]
  132. Seo, J.; Kim, Y.; Ji, S.; Kim, H.B.; Jung, H.; Yi, E.C.; Lee, Y.H.; Shin, I.; Yang, W.H.; Cho, J.W. O-GlcNAcylation of RIPK1 rescues red blood cells from necroptosis. Front. Immunol. 2023, 14, 1160490. [Google Scholar] [CrossRef] [PubMed]
  133. Giogha, C.; Scott, N.E.; Wong Fok Lung, T.; Pollock, G.L.; Harper, M.; Goddard-Borger, E.D.; Pearson, J.S.; Hartland, E.L. NleB2 from enteropathogenic Escherichia coli is a novel arginine-glucose transferase effector. PLoS Pathog. 2021, 17, e1009658. [Google Scholar] [CrossRef]
  134. Carafa, V.; Nebbioso, A.; Cuomo, F.; Rotili, D.; Cobellis, G.; Bontempo, P.; Baldi, A.; Spugnini, E.P.; Citro, G.; Chambery, A.; et al. RIP1-HAT1-SIRT Complex Identification and Targeting in Treatment and Prevention of Cancer. Clin. Cancer Res.. 2018, 24, 2886–2900. [Google Scholar] [CrossRef] [PubMed]
  135. Lalaoui, N.; Boyden, S.E.; Oda, H.; Wood, G.M.; Stone, D.L.; Chau, D.; Liu, L.; Stoffels, M.; Kratina, T.; Lawlor, K.E.; et al. Mutations that prevent caspase cleavage of RIPK1 cause autoinflammatory disease. Nature 2020, 577, 103–108. [Google Scholar] [CrossRef]
  136. van Raam, B.J.; Ehrnhoefer, D.E.; Hayden, M.R.; Salvesen, G.S. Intrinsic cleavage of receptor-interacting protein kinase-1 by caspase-6. Cell Death Differ. 2013, 20, 86–96. [Google Scholar] [CrossRef]
  137. Cho, M.; Dho, S.H.; Shin, S.; Lee, Y.; Kim, Y.; Lee, J.; Yu, S.J.; Park, S.H.; Lee, K.A.; Kim, L.K. Caspase-10 affects the pathogenesis of primary biliary cholangitis by regulating inflammatory cell death. J. Autoimmun. 2022, 133, 102940. [Google Scholar] [CrossRef]
  138. Yan, L.; Zhang, T.; Wang, K.; Chen, Z.; Yang, Y.; Shan, B.; Sun, Q.; Zhang, M.; Zhang, Y.; Zhong, Y.; et al. SENP1 prevents steatohepatitis by suppressing RIPK1-driven apoptosis and inflammation. Nat. Commun. 2022, 13, 7153. [Google Scholar] [CrossRef]
  139. Onizawa, M.; Oshima, S.; Schulze-Topphoff, U.; Oses-Prieto, J.A.; Lu, T.; Tavares, R.; Prodhomme, T.; Duong, B.; Whang, M.I.; Advincula, R.; et al. The ubiquitin-modifying enzyme A20 restricts ubiquitination of the kinase RIPK3 and protects cells from necroptosis. Nat. Immunol. 2015, 16, 618–627. [Google Scholar] [CrossRef]
  140. Zhou, M.; He, J.; Shi, Y.; Liu, X.; Luo, S.; Cheng, C.; Ge, W.; Qu, C.; Du, P.; Chen, Y. ABIN3 Negatively Regulates Necroptosis-induced Intestinal Inflammation Through Recruiting A20 and Restricting the Ubiquitination of RIPK3 in Inflammatory Bowel Disease. J. Crohns Colitis 2021, 15, 99–114. [Google Scholar] [CrossRef]
  141. Frank, D.; Garnish, S.E.; Sandow, J.J.; Weir, A.; Liu, L.; Clayer, E.; Meza, L.; Rashidi, M.; Cobbold, S.A.; Scutts, S.R.; et al. Ubiquitylation of RIPK3 beyond-the-RHIM can limit RIPK3 activity and cell death. iScience 2022, 25, 104632. [Google Scholar] [CrossRef] [PubMed]
  142. Roedig, J.; Kowald, L.; Juretschke, T.; Karlowitz, R.; Ahangarian Abhari, B.; Roedig, H.; Fulda, S.; Beli, P.; van Wijk, S.J. USP22 controls necroptosis by regulating receptor-interacting protein kinase 3 ubiquitination. EMBO Rep. 2021, 22, e50163. [Google Scholar] [CrossRef]
  143. Choi, S.W.; Park, H.H.; Kim, S.; Chung, J.M.; Noh, H.J.; Kim, S.K.; Song, H.K.; Lee, C.W.; Morgan, M.J.; Kang, H.C.; et al. PELI1 Selectively Targets Kinase-Active RIP3 for Ubiquitylation-Dependent Proteasomal Degradation. Mol. Cell 2018, 70, 920–935.e7. [Google Scholar] [CrossRef] [PubMed]
  144. Mei, P.; Xie, F.; Pan, J.; Wang, S.; Gao, W.; Ge, R.; Gao, B.; Gao, S.; Chen, X.; Wang, Y.; et al. E3 ligase TRIM25 ubiquitinates RIP3 to inhibit TNF induced cell necrosis. Cell Death Differ. 2021, 28, 2888–2899. [Google Scholar] [CrossRef] [PubMed]
  145. Moriwaki, K.; Chan, F.K. Regulation of RIPK3- and RHIM-dependent Necroptosis by the Proteasome. J. Biol. Chem. 2016, 291, 5948–5959. [Google Scholar] [CrossRef]
  146. Lee, S.B.; Kim, J.J.; Han, S.A.; Fan, Y.; Guo, L.S.; Aziz, K.; Nowsheen, S.; Kim, S.S.; Park, S.Y.; Luo, Q.; et al. The AMPK-Parkin axis negatively regulates necroptosis and tumorigenesis by inhibiting the necrosome. Nat. Cell Biol. 2019, 21, 940–951. [Google Scholar] [CrossRef]
  147. Chen, W.; Zhou, Z.; Li, L.; Zhong, C.Q.; Zheng, X.; Wu, X.; Zhang, Y.; Ma, H.; Huang, D.; Li, W.; et al. Diverse sequence determinants control human and mouse receptor interacting protein 3 (RIP3) and mixed lineage kinase domain-like (MLKL) interaction in necroptotic signaling. J. Biol. Chem. 2013, 288, 16247–16261. [Google Scholar] [CrossRef]
  148. Meng, Y.; Davies, K.A.; Fitzgibbon, C.; Young, S.N.; Garnish, S.E.; Horne, C.R.; Luo, C.; Garnier, J.M.; Liang, L.Y.; Cowan, A.D.; et al. Human RIPK3 maintains MLKL in an inactive conformation prior to cell death by necroptosis. Nat. Commun. 2021, 12, 6783. [Google Scholar] [CrossRef]
  149. Sun, L.; Wang, H.; Wang, Z.; He, S.; Chen, S.; Liao, D.; Wang, L.; Yan, J.; Liu, W.; Lei, X.; et al. Mixed lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell 2012, 148, 213–227. [Google Scholar] [CrossRef]
  150. Meng, Y.; Horne, C.R.; Samson, A.L.; Dagley, L.F.; Young, S.N.; Sandow, J.J.; Czabotar, P.E.; Murphy, J.M. Human RIPK3 C-lobe phosphorylation is essential for necroptotic signaling. Cell Death Dis. 2022, 13, 565. [Google Scholar] [CrossRef]
  151. Li, D.; Ai, Y.; Guo, J.; Dong, B.; Li, L.; Cai, G.; Chen, S.; Xu, D.; Wang, F.; Wang, X. Casein kinase 1G2 suppresses necroptosis-promoted testis aging by inhibiting receptor-interacting kinase 3. Elife 2020, 9, e61564. [Google Scholar] [CrossRef]
  152. Li, D.; Chen, J.; Guo, J.; Li, L.; Cai, G.; Chen, S.; Huang, J.; Yang, H.; Zhuang, Y.; Wang, F.; et al. A phosphorylation of RIPK3 kinase initiates an intracellular apoptotic pathway that promotes prostaglandin(2α)-induced corpus luteum regression. Elife 2021, 10, e67409. [Google Scholar] [CrossRef]
  153. Hanna-Addams, S.; Liu, S.; Liu, H.; Chen, S.; Wang, Z. CK1α, CK1δ, and CK1ε are necrosome components which phosphorylate serine 227 of human RIPK3 to activate necroptosis. Proc. Natl. Acad. Sci. USA 2020, 117, 1962–1970. [Google Scholar] [CrossRef] [PubMed]
  154. Chen, W.; Wu, J.; Li, L.; Zhang, Z.; Ren, J.; Liang, Y.; Chen, F.; Yang, C.; Zhou, Z.; Su, S.S.; et al. Ppm1b negatively regulates necroptosis through dephosphorylating Rip3. Nat. Cell Biol. 2015, 17, 434–444. [Google Scholar] [CrossRef]
  155. Li, X.; Gong, W.; Wang, H.; Li, T.; Attri, K.S.; Lewis, R.E.; Kalil, A.C.; Bhinderwala, F.; Powers, R.; Yin, G.; et al. O-GlcNAc Transferase Suppresses Inflammation and Necroptosis by Targeting Receptor-Interacting Serine/Threonine-Protein Kinase 3. Immunity 2019, 50, 576–590.e6. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, J.; Yu, P.; Hua, F.; Hu, Y.; Xiao, F.; Liu, Q.; Huang, D.; Deng, F.; Wei, G.; Deng, W.; et al. Sevoflurane postconditioning reduces myocardial ischemia reperfusion injury-induced necroptosis by up-regulation of OGT-mediated O-GlcNAcylated RIPK3. Aging 2020, 12, 25452–25468. [Google Scholar] [CrossRef]
  157. Zhang, B.; Li, M.D.; Yin, R.; Liu, Y.; Yang, Y.; Mitchell-Richards, K.A.; Nam, J.H.; Li, R.; Wang, L.; Iwakiri, Y.; et al. O-GlcNAc transferase suppresses necroptosis and liver fibrosis. JCI Insight 2019, 4, e127709. [Google Scholar] [CrossRef] [PubMed]
  158. Park, J.; Ha, H.J.; Chung, E.S.; Baek, S.H.; Cho, Y.; Kim, H.K.; Han, J.; Sul, J.H.; Lee, J.; Kim, E.; et al. O-GlcNAcylation ameliorates the pathological manifestations of Alzheimer’s disease by inhibiting necroptosis. Sci. Adv. 2021, 7, eabd3207. [Google Scholar] [CrossRef]
  159. Wu, F.; Shao, Q.; Cheng, Z.; Xiong, X.; Fang, K.; Zhao, Y.; Dong, R.; Xu, L.; Lu, F.; Chen, G. Traditional herbal formula Wu-Mei-Wan alleviates TNBS-induced colitis in mice by inhibiting necroptosis through increasing RIPK3 O-GlcNAcylation. Chin. Med. 2021, 16, 78. [Google Scholar] [CrossRef]
  160. Zhang, L.; He, Y.; Jiang, Y.; Wu, Q.; Liu, Y.; Xie, Q.; Zou, Y.; Wu, J.; Zhang, C.; Zhou, Z.; et al. PRMT1 reverts the immune escape of necroptotic colon cancer through RIP3 methylation. Cell Death Dis. 2023, 14, 233. [Google Scholar] [CrossRef]
  161. Chauhan, C.; Martinez-Val, A.; Niedenthal, R.; Olsen, J.V.; Kotlyarov, A.; Bekker-Jensen, S.; Gaestel, M.; Menon, M.B. PRMT5-mediated regulatory arginine methylation of RIPK3. Cell Death Discov. 2023, 9, 14. [Google Scholar] [CrossRef] [PubMed]
  162. Feng, S.; Yang, Y.; Mei, Y.; Ma, L.; Zhu, D.E.; Hoti, N.; Castanares, M.; Wu, M. Cleavage of RIP3 inactivates its caspase-independent apoptosis pathway by removal of kinase domain. Cell. Signal. 2007, 19, 2056–2067. [Google Scholar] [CrossRef] [PubMed]
  163. Miao, W.; Qu, Z.; Shi, K.; Zhang, D.; Zong, Y.; Zhang, G.; Zhang, G.; Hu, S. RIP3 S-nitrosylation contributes to cerebral ischemic neuronal injury. Brain Res. 2015, 1627, 165–176. [Google Scholar] [CrossRef]
  164. Zhong, Y.; Peng, P.; Zhang, M.; Han, D.; Yang, H.; Yan, X.; Hu, S. Effect of S-Nitrosylation of RIP3 Induced by Cerebral Ischemia on its Downstream Signaling Pathway. J. Stroke Cerebrovasc. Dis. 2022, 31, 106516. [Google Scholar] [CrossRef]
  165. Murphy, J.M.; Czabotar, P.E.; Hildebrand, J.M.; Lucet, I.S.; Zhang, J.G.; Alvarez-Diaz, S.; Lewis, R.; Lalaoui, N.; Metcalf, D.; Webb, A.I.; et al. The pseudokinase MLKL mediates necroptosis via a molecular switch mechanism. Immunity 2013, 39, 443–453. [Google Scholar] [CrossRef]
  166. Garcia, L.R.; Tenev, T.; Newman, R.; Haich, R.O.; Liccardi, G.; John, S.W.; Annibaldi, A.; Yu, L.; Pardo, M.; Young, S.N.; et al. Ubiquitylation of MLKL at lysine 219 positively regulates necroptosis-induced tissue injury and pathogen clearance. Nat. Commun. 2021, 12, 3364. [Google Scholar] [CrossRef]
  167. Liu, Z.; Dagley, L.F.; Shield-Artin, K.; Young, S.N.; Bankovacki, A.; Wang, X.; Tang, M.; Howitt, J.; Stafford, C.A.; Nachbur, U.; et al. Oligomerization-driven MLKL ubiquitylation antagonizes necroptosis. EMBO J. 2021, 40, e103718. [Google Scholar] [CrossRef]
  168. Yoon, S.; Bogdanov, K.; Wallach, D. Site-specific ubiquitination of MLKL targets it to endosomes and targets Listeria and Yersinia to the lysosomes. Cell Death Differ. 2022, 29, 306–322. [Google Scholar] [CrossRef] [PubMed]
  169. Zhou, H.; Zhou, L.; Guan, Q.; Hou, X.; Wang, C.; Liu, L.; Wang, J.; Yu, X.; Li, W.; Liu, H. Skp2-mediated MLKL degradation confers cisplatin-resistant in non-small cell lung cancer cells. Commun. Biol. 2023, 6, 805. [Google Scholar] [CrossRef]
  170. Weinelt, N.; Wächtershäuser, K.N.; Celik, G.; Jeiler, B.; Gollin, I.; Zein, L.; Smith, S.; Andrieux, G.; Das, T.; Roedig, J.; et al. LUBAC-mediated M1 Ub regulates necroptosis by segregating the cellular distribution of active MLKL. Cell Death Dis. 2024, 15, 77. [Google Scholar] [CrossRef]
  171. Davies, K.A.; Fitzgibbon, C.; Young, S.N.; Garnish, S.E.; Yeung, W.; Coursier, D.; Birkinshaw, R.W.; Sandow, J.J.; Lehmann, W.I.L.; Liang, L.Y.; et al. Distinct pseudokinase domain conformations underlie divergent activation mechanisms among vertebrate MLKL orthologues. Nat. Commun. 2020, 11, 3060. [Google Scholar] [CrossRef] [PubMed]
  172. Wang, H.; Sun, L.; Su, L.; Rizo, J.; Liu, L.; Wang, L.F.; Wang, F.S.; Wang, X. Mixed lineage kinase domain-like protein MLKL causes necrotic membrane disruption upon phosphorylation by RIP3. Mol. Cell 2014, 54, 133–146. [Google Scholar] [CrossRef]
  173. Zhan, Q.; Jeon, J.; Li, Y.; Huang, Y.; Xiong, J.; Wang, Q.; Xu, T.L.; Li, Y.; Ji, F.H.; Du, G.; et al. CAMK2/CaMKII activates MLKL in short-term starvation to facilitate autophagic flux. Autophagy 2022, 18, 726–744. [Google Scholar] [CrossRef]
  174. Rodriguez, D.A.; Weinlich, R.; Brown, S.; Guy, C.; Fitzgerald, P.; Dillon, C.P.; Oberst, A.; Quarato, G.; Low, J.; Cripps, J.G.; et al. Characterization of RIPK3-mediated phosphorylation of the activation loop of MLKL during necroptosis. Cell Death Differ. 2016, 23, 76–88. [Google Scholar] [CrossRef]
  175. Tanzer, M.C.; Tripaydonis, A.; Webb, A.I.; Young, S.N.; Varghese, L.N.; Hall, C.; Alexander, W.S.; Hildebrand, J.M.; Silke, J.; Murphy, J.M. Necroptosis signalling is tuned by phosphorylation of MLKL residues outside the pseudokinase domain activation loop. Biochem. J. 2015, 471, 255–265. [Google Scholar] [CrossRef]
  176. Ying, Z.; Pan, C.; Shao, T.; Liu, L.; Li, L.; Guo, D.; Zhang, S.; Yuan, T.; Cao, R.; Jiang, Z.; et al. Mixed Lineage Kinase Domain-like Protein MLKL Breaks Down Myelin following Nerve Injury. Mol. Cell 2018, 72, 457–468.e5. [Google Scholar] [CrossRef] [PubMed]
  177. Najafov, A.; Mookhtiar, A.K.; Luu, H.S.; Ordureau, A.; Pan, H.; Amin, P.P.; Li, Y.; Lu, Q.; Yuan, J. TAM Kinases Promote Necroptosis by Regulating Oligomerization of MLKL. Mol. Cell 2019, 75, 457–468.e4. [Google Scholar] [CrossRef]
  178. Zhu, X.; Yang, N.; Yang, Y.; Yuan, F.; Yu, D.; Zhang, Y.; Shu, Z.; Nan, N.; Hu, H.; Liu, X.; et al. Spontaneous necroptosis and autoinflammation are blocked by an inhibitory phosphorylation on MLKL during neonatal development. Cell Res. 2022, 32, 407–410. [Google Scholar] [CrossRef] [PubMed]
  179. Liu, S.; Liu, H.; Johnston, A.; Hanna-Addams, S.; Reynoso, E.; Xiang, Y.; Wang, Z. MLKL forms disulfide bond-dependent amyloid-like polymers to induce necroptosis. Proc. Natl. Acad. Sci. USA 2017, 114, E7450–E7459. [Google Scholar] [CrossRef]
  180. Reynoso, E.; Liu, H.; Li, L.; Yuan, A.L.; Chen, S.; Wang, Z. Thioredoxin-1 actively maintains the pseudokinase MLKL in a reduced state to suppress disulfide bond-dependent MLKL polymer formation and necroptosis. J. Biol. Chem. 2017, 292, 17514–17524. [Google Scholar] [CrossRef]
  181. Seo, J.; Lee, E.W.; Shin, J.; Seong, D.; Nam, Y.W.; Jeong, M.; Lee, S.H.; Lee, C.; Song, J. K6 linked polyubiquitylation of FADD by CHIP prevents death inducing signaling complex formation suppressing cell death. Oncogene 2018, 37, 4994–5006. [Google Scholar] [CrossRef] [PubMed]
  182. Lee, E.W.; Kim, J.H.; Ahn, Y.H.; Seo, J.; Ko, A.; Jeong, M.; Kim, S.J.; Ro, J.Y.; Park, K.M.; Lee, H.W.; et al. Ubiquitination and degradation of the FADD adaptor protein regulate death receptor-mediated apoptosis and necroptosis. Nat. Commun. 2012, 3, 978. [Google Scholar] [CrossRef] [PubMed]
  183. Newson, J.P.M.; Scott, N.E.; Yeuk Wah Chung, I.; Wong Fok Lung, T.; Giogha, C.; Gan, J.; Wang, N.; Strugnell, R.A.; Brown, N.F.; Cygler, M.; et al. Salmonella Effectors SseK1 and SseK3 Target Death Domain Proteins in the TNF and TRAIL Signaling Pathways. Mol. Cell. Proteom. 2019, 18, 1138–1156. [Google Scholar] [CrossRef]
  184. Xue, J.; Hu, S.; Huang, Y.; Zhang, Q.; Yi, X.; Pan, X.; Li, S. Arg-GlcNAcylation on TRADD by NleB and SseK1 Is Crucial for Bacterial Pathogenesis. Front. Cell Dev. Biol. 2020, 8, 641. [Google Scholar] [CrossRef] [PubMed]
  185. Li, S.; Zhang, L.; Yao, Q.; Li, L.; Dong, N.; Rong, J.; Gao, W.; Ding, X.; Sun, L.; Chen, X.; et al. Pathogen blocks host death receptor signalling by arginine GlcNAcylation of death domains. Nature 2013, 501, 242–246. [Google Scholar] [CrossRef]
  186. Fritsch, J.; Stephan, M.; Tchikov, V.; Winoto-Morbach, S.; Gubkina, S.; Kabelitz, D.; Schütze, S. Cell fate decisions regulated by K63 ubiquitination of tumor necrosis factor receptor 1. Mol. Cell. Biol. 2014, 34, 3214–3228. [Google Scholar] [CrossRef]
  187. Marín-Rubio, J.L.; Pérez-Gómez, E.; Fernández-Piqueras, J.; Villa-Morales, M. S194-P-FADD as a marker of aggressiveness and poor prognosis in human T-cell lymphoblastic lymphoma. Carcinogenesis 2019, 40, 1260–1268. [Google Scholar] [CrossRef]
  188. Alappat, E.C.; Feig, C.; Boyerinas, B.; Volkland, J.; Samuels, M.; Murmann, A.E.; Thorburn, A.; Kidd, V.J.; Slaughter, C.A.; Osborn, S.L.; et al. Phosphorylation of FADD at serine 194 by CKIalpha regulates its nonapoptotic activities. Mol. Cell 2005, 19, 321–332. [Google Scholar] [CrossRef]
  189. Vilmont, V.; Filhol, O.; Hesse, A.M.; Couté, Y.; Hue, C.; Rémy-Tourneur, L.; Mistou, S.; Cochet, C.; Chiocchia, G. Modulatory role of the anti-apoptotic protein kinase CK2 in the sub-cellular localization of Fas associated death domain protein (FADD). Biochim. Biophys. Acta 2015, 1853, 2885–2896. [Google Scholar] [CrossRef]
  190. Jang, M.S.; Lee, S.J.; Kang, N.S.; Kim, E. Cooperative phosphorylation of FADD by Aur-A and Plk1 in response to taxol triggers both apoptotic and necrotic cell death. Cancer Res. 2011, 71, 7207–7215. [Google Scholar] [CrossRef]
  191. Choi, S.G.; Kim, H.; Jeong, E.I.; Lee, H.J.; Park, S.; Lee, S.Y.; Lee, H.J.; Lee, S.W.; Chung, C.H.; Jung, Y.K. SUMO-Modified FADD Recruits Cytosolic Drp1 and Caspase-10 to Mitochondria for Regulated Necrosis. Mol. Cell. Biol. 2017, 37, e00254-16. [Google Scholar] [CrossRef] [PubMed]
  192. Jin, Z.; Li, Y.; Pitti, R.; Lawrence, D.; Pham, V.C.; Lill, J.R.; Ashkenazi, A. Cullin3-based polyubiquitination and p62-dependent aggregation of caspase-8 mediate extrinsic apoptosis signaling. Cell 2009, 137, 721–735. [Google Scholar] [CrossRef]
  193. Tomar, D.; Prajapati, P.; Sripada, L.; Singh, K.; Singh, R.; Singh, A.K.; Singh, R. TRIM13 regulates caspase-8 ubiquitination, translocation to autophagosomes and activation during ER stress induced cell death. Biochim. Biophys. Acta 2013, 1833, 3134–3144. [Google Scholar] [CrossRef]
  194. Li, Y.; Kong, Y.; Zhou, Z.; Chen, H.; Wang, Z.; Hsieh, Y.C.; Zhao, D.; Zhi, X.; Huang, J.; Zhang, J.; et al. The HECTD3 E3 ubiquitin ligase facilitates cancer cell survival by promoting K63-linked polyubiquitination of caspase-8. Cell Death Dis. 2013, 4, e935. [Google Scholar] [CrossRef]
  195. Gonzalvez, F.; Lawrence, D.; Yang, B.; Yee, S.; Pitti, R.; Marsters, S.; Pham, V.C.; Stephan, J.P.; Lill, J.; Ashkenazi, A. TRAF2 Sets a threshold for extrinsic apoptosis by tagging caspase-8 with a ubiquitin shutoff timer. Mol. Cell 2012, 48, 888–899. [Google Scholar] [CrossRef]
  196. Yang, Z.H.; Wu, X.N.; He, P.; Wang, X.; Wu, J.; Ai, T.; Zhong, C.Q.; Wu, X.; Cong, Y.; Zhu, R.; et al. A Non-canonical PDK1-RSK Signal Diminishes Pro-caspase-8-Mediated Necroptosis Blockade. Mol. Cell 2020, 80, 296–310.e6. [Google Scholar] [CrossRef]
  197. Peng, C.; Cho, Y.Y.; Zhu, F.; Zhang, J.; Wen, W.; Xu, Y.; Yao, K.; Ma, W.Y.; Bode, A.M.; Dong, Z. Phosphorylation of caspase-8 (Thr-263) by ribosomal S6 kinase 2 (RSK2) mediates caspase-8 ubiquitination and stability. J. Biol. Chem. 2011, 286, 6946–6954. [Google Scholar] [CrossRef] [PubMed]
  198. Helmke, C.; Raab, M.; Rödel, F.; Matthess, Y.; Oellerich, T.; Mandal, R.; Sanhaji, M.; Urlaub, H.; Rödel, C.; Becker, S.; et al. Ligand stimulation of CD95 induces activation of Plk3 followed by phosphorylation of caspase-8. Cell Res. 2016, 26, 914–934. [Google Scholar] [CrossRef] [PubMed]
  199. Alvarado-Kristensson, M.; Melander, F.; Leandersson, K.; Rönnstrand, L.; Wernstedt, C.; Andersson, T. p38-MAPK signals survival by phosphorylation of caspase-8 and caspase-3 in human neutrophils. J. Exp. Med. 2004, 199, 449–458. [Google Scholar] [CrossRef]
  200. Cursi, S.; Rufini, A.; Stagni, V.; Condò, I.; Matafora, V.; Bachi, A.; Bonifazi, A.P.; Coppola, L.; Superti-Furga, G.; Testi, R.; et al. Src kinase phosphorylates Caspase-8 on Tyr380: A novel mechanism of apoptosis suppression. EMBO J. 2006, 25, 1895–1905. [Google Scholar] [CrossRef]
  201. Matthess, Y.; Raab, M.; Sanhaji, M.; Lavrik, I.N.; Strebhardt, K. Cdk1/cyclin B1 controls Fas-mediated apoptosis by regulating caspase-8 activity. Mol. Cell. Biol. 2010, 30, 5726–5740. [Google Scholar] [CrossRef] [PubMed]
  202. Parrish, A.B.; Freel, C.D.; Kornbluth, S. Cellular mechanisms controlling caspase activation and function. Cold Spring Harb. Perspect. Biol. 2013, 5, a008672. [Google Scholar] [CrossRef] [PubMed]
  203. Kim, Y.M.; Kim, T.H.; Chung, H.T.; Talanian, R.V.; Yin, X.M.; Billiar, T.R. Nitric oxide prevents tumor necrosis factor alpha-induced rat hepatocyte apoptosis by the interruption of mitochondrial apoptotic signaling through S-nitrosylation of caspase-8. Hepatology 2000, 32, 770–778. [Google Scholar] [CrossRef]
  204. Besnault-Mascard, L.; Leprince, C.; Auffredou, M.T.; Meunier, B.; Bourgeade, M.F.; Camonis, J.; Lorenzo, H.K.; Vazquez, A. Caspase-8 sumoylation is associated with nuclear localization. Oncogene 2005, 24, 3268–3273. [Google Scholar] [CrossRef]
  205. Tang, Y.; Joo, D.; Liu, G.; Tu, H.; You, J.; Jin, J.; Zhao, X.; Hung, M.C.; Lin, X. Linear ubiquitination of cFLIP induced by LUBAC contributes to TNFα-induced apoptosis. J. Biol. Chem. 2018, 293, 20062–20072. [Google Scholar] [CrossRef]
  206. Nakabayashi, O.; Takahashi, H.; Moriwaki, K.; Komazawa-Sakon, S.; Ohtake, F.; Murai, S.; Tsuchiya, Y.; Koyahara, Y.; Saeki, Y.; Yoshida, Y.; et al. MIND bomb 2 prevents RIPK1 kinase activity-dependent and -independent apoptosis through ubiquitylation of cFLIP(L). Commun. Biol. 2021, 4, 80. [Google Scholar] [CrossRef] [PubMed]
  207. Poukkula, M.; Kaunisto, A.; Hietakangas, V.; Denessiouk, K.; Katajamäki, T.; Johnson, M.S.; Sistonen, L.; Eriksson, J.E. Rapid turnover of c-FLIPshort is determined by its unique C-terminal tail. J. Biol. Chem. 2005, 280, 27345–27355. [Google Scholar] [CrossRef]
  208. Wilkie-Grantham, R.P.; Matsuzawa, S.; Reed, J.C. Novel phosphorylation and ubiquitination sites regulate reactive oxygen species-dependent degradation of anti-apoptotic c-FLIP protein. J. Biol. Chem. 2013, 288, 12777–12790. [Google Scholar] [CrossRef]
  209. Chang, L.; Kamata, H.; Solinas, G.; Luo, J.L.; Maeda, S.; Venuprasad, K.; Liu, Y.C.; Karin, M. The E3 ubiquitin ligase itch couples JNK activation to TNFalpha-induced cell death by inducing c-FLIP(L) turnover. Cell 2006, 124, 601–613. [Google Scholar] [CrossRef]
  210. Roberts, J.Z.; Holohan, C.; Sessler, T.; Fox, J.; Crawford, N.; Riley, J.S.; Khawaja, H.; Majkut, J.; Evergren, E.; Humphreys, L.M.; et al. The SCF(Skp2) ubiquitin ligase complex modulates TRAIL-R2-induced apoptosis by regulating FLIP(L). Cell Death Differ. 2020, 27, 2726–2741. [Google Scholar] [CrossRef]
  211. Dold, M.N.; Ng, X.; Alber, C.; Gentle, I.E.; Häcker, G.; Weber, A. The deubiquitinase Usp27x as a novel regulator of cFLIP(L) protein expression and sensitizer to death-receptor-induced apoptosis. Apoptosis 2022, 27, 112–132. [Google Scholar] [CrossRef]
  212. Shi, B.; Tran, T.; Sobkoviak, R.; Pope, R.M. Activation-induced degradation of FLIP(L) is mediated via the phosphatidylinositol 3-kinase/Akt signaling pathway in macrophages. J. Biol. Chem. 2009, 284, 14513–14523. [Google Scholar] [CrossRef] [PubMed]
  213. Chanvorachote, P.; Nimmannit, U.; Wang, L.; Stehlik, C.; Lu, B.; Azad, N.; Rojanasakul, Y. Nitric oxide negatively regulates Fas CD95-induced apoptosis through inhibition of ubiquitin-proteasome-mediated degradation of FLICE inhibitory protein. J. Biol. Chem. 2005, 280, 42044–42050. [Google Scholar] [CrossRef]
  214. Scaffidi, C.; Medema, J.P.; Krammer, P.H.; Peter, M.E. FLICE is predominantly expressed as two functionally active isoforms, caspase-8/a and caspase-8/b. J. Biol. Chem. 1997, 272, 26953–26958. [Google Scholar] [CrossRef]
  215. Xu, L.; Zhang, Y.; Qu, X.; Che, X.; Guo, T.; Li, C.; Ma, R.; Fan, Y.; Ma, Y.; Hou, K.; et al. DR5-Cbl-b/c-Cbl-TRAF2 complex inhibits TRAIL-induced apoptosis by promoting TRAF2-mediated polyubiquitination of caspase-8 in gastric cancer cells. Mol. Oncol. 2017, 11, 1733–1751. [Google Scholar] [CrossRef] [PubMed]
  216. Oztürk, S.; Schleich, K.; Lavrik, I.N. Cellular FLICE-like inhibitory proteins (c-FLIPs): Fine-tuners of life and death decisions. Exp Cell Res. 2012, 318, 1324–1331. [Google Scholar] [CrossRef] [PubMed]
  217. Yu, J.W.; Jeffrey, P.D.; Shi, Y. Mechanism of procaspase-8 activation by c-FLIPL. Proc. Natl. Acad. Sci. USA 2009, 106, 8169–8174. [Google Scholar] [CrossRef]
  218. Li, X.; Wilmanns, M.; Thornton, J.; Köhn, M. Elucidating human phosphatase-substrate networks. Sci. Signal. 2013, 6, rs10. [Google Scholar] [CrossRef]
  219. Ardito, F.; Giuliani, M.; Perrone, D.; Troiano, G.; Lo Muzio, L. The crucial role of protein phosphorylation in cell signaling and its use as targeted therapy (Review). Int. J. Mol. Med. 2017, 40, 271–280. [Google Scholar] [CrossRef] [PubMed]
  220. Dondelinger, Y.; Jouan-Lanhouet, S.; Divert, T.; Theatre, E.; Bertin, J.; Gough, P.J.; Giansanti, P.; Heck, A.J.; Dejardin, E.; Vandenabeele, P.; et al. NF-κB-Independent Role of IKKα/IKKβ in Preventing RIPK1 Kinase-Dependent Apoptotic and Necroptotic Cell Death during TNF Signaling. Mol. Cell 2015, 60, 63–76. [Google Scholar] [CrossRef]
  221. Blanchett, S.; Dondelinger, Y.; Barbarulo, A.; Bertrand, M.J.M.; Seddon, B. Phosphorylation of RIPK1 serine 25 mediates IKK dependent control of extrinsic cell death in T cells. Front. Immunol. 2022, 13, 1067164. [Google Scholar] [CrossRef] [PubMed]
  222. Wang, B.; Fu, J.; Chai, Y.; Liu, Y.; Chen, Y.; Yin, J.; Pu, Y.; Chen, C.; Wang, F.; Liu, Z.; et al. Accumulation of RIPK1 into mitochondria is requisite for oxidative stress-mediated necroptosis and proliferation in Rat Schwann cells. Int. J. Med. Sci. 2022, 19, 1965–1976. [Google Scholar] [CrossRef]
  223. Wu, X.N.; Yang, Z.H.; Wang, X.K.; Zhang, Y.; Wan, H.; Song, Y.; Chen, X.; Shao, J.; Han, J. Distinct roles of RIP1-RIP3 hetero- and RIP3-RIP3 homo-interaction in mediating necroptosis. Cell Death Differ. 2014, 21, 1709–1720. [Google Scholar] [CrossRef] [PubMed]
  224. Tan, L.; Chan, W.; Zhang, J.; Wang, J.; Wang, Z.; Liu, J.; Li, J.; Liu, X.; Wang, M.; Hao, L.; et al. Regulation of RIP1-Mediated necroptosis via necrostatin-1 in periodontitis. J. Periodontal Res. 2023, 58, 919–931. [Google Scholar] [CrossRef] [PubMed]
  225. Gupta, K.; Liu, B. PLK1-mediated S369 phosphorylation of RIPK3 during G2 and M phases enables its ripoptosome incorporation and activity. iScience 2021, 24, 102320. [Google Scholar] [CrossRef]
  226. Wu, X.; Fan, X.; McMullen, M.R.; Miyata, T.; Kim, A.; Pathak, V.; Wu, J.; Day, L.Z.; Hardesty, J.E.; Welch, N.; et al. Macrophage-derived MLKL in alcohol-associated liver disease: Regulation of phagocytosis. Hepatology 2023, 77, 902–919. [Google Scholar] [CrossRef]
  227. Muzio, M.; Chinnaiyan, A.M.; Kischkel, F.C.; O’Rourke, K.; Shevchenko, A.; Ni, J.; Scaffidi, C.; Bretz, J.D.; Zhang, M.; Gentz, R.; et al. FLICE, a novel FADD-homologous ICE/CED-3-like protease, is recruited to the CD95 (Fas/APO-1) death--inducing signaling complex. Cell 1996, 85, 817–827. [Google Scholar] [CrossRef]
  228. Newton, K.; Wickliffe, K.E.; Dugger, D.L.; Maltzman, A.; Roose-Girma, M.; Dohse, M.; Kőműves, L.; Webster, J.D.; Dixit, V.M. Cleavage of RIPK1 by caspase-8 is crucial for limiting apoptosis and necroptosis. Nature 2019, 574, 428–431. [Google Scholar] [CrossRef]
  229. Powley, I.R.; Hughes, M.A.; Cain, K.; MacFarlane, M. Caspase-8 tyrosine-380 phosphorylation inhibits CD95 DISC function by preventing procaspase-8 maturation and cycling within the complex. Oncogene 2016, 35, 5629–5640. [Google Scholar] [CrossRef]
  230. Barbero, S.; Barilà, D.; Mielgo, A.; Stagni, V.; Clair, K.; Stupack, D. Identification of a critical tyrosine residue in caspase 8 that promotes cell migration. J. Biol. Chem. 2008, 283, 13031–13034. [Google Scholar] [CrossRef]
  231. Senft, J.; Helfer, B.; Frisch, S.M. Caspase-8 interacts with the p85 subunit of phosphatidylinositol 3-kinase to regulate cell adhesion and motility. Cancer Res. 2007, 67, 11505–11509. [Google Scholar] [CrossRef]
  232. Jia, S.H.; Parodo, J.; Kapus, A.; Rotstein, O.D.; Marshall, J.C. Dynamic regulation of neutrophil survival through tyrosine phosphorylation or dephosphorylation of caspase-8. J. Biol. Chem. 2008, 283, 5402–5413. [Google Scholar] [CrossRef] [PubMed]
  233. Günster, R.A.; Matthews, S.A.; Holden, D.W.; Thurston, T.L.M. SseK1 and SseK3 Type III Secretion System Effectors Inhibit NF-κB Signaling and Necroptotic Cell Death in Salmonella-Infected Macrophages. Infect. Immun. 2017, 85, e00010-17. [Google Scholar] [CrossRef]
  234. Pan, X.; Luo, J.; Li, S. Bacteria-Catalyzed Arginine Glycosylation in Pathogens and Host. Front. Cell. Infect. Microbiol. 2020, 10, 185. [Google Scholar] [CrossRef] [PubMed]
  235. Ding, J.; Pan, X.; Du, L.; Yao, Q.; Xue, J.; Yao, H.; Wang, D.C.; Li, S.; Shao, F. Structural and Functional Insights into Host Death Domains Inactivation by the Bacterial Arginine GlcNAcyltransferase Effector. Mol. Cell 2019, 74, 922–935.e6. [Google Scholar] [CrossRef]
  236. Lake, A.N.; Bedford, M.T. Protein methylation and DNA repair. Mutat. Res. 2007, 618, 91–101. [Google Scholar] [CrossRef] [PubMed]
  237. Smith, B.C.; Denu, J.M. Chemical mechanisms of histone lysine and arginine modifications. Biochim. Biophys. Acta 2009, 1789, 45–57. [Google Scholar] [CrossRef]
  238. Wu, W.; Wang, J.; Xiao, C.; Su, Z.; Su, H.; Zhong, W.; Mao, J.; Liu, X.; Zhu, Y.Z. SMYD2-mediated TRAF2 methylation promotes the NF-κB signaling pathways in inflammatory diseases. Clin. Transl. Med. 2021, 11, e591. [Google Scholar] [CrossRef]
  239. Allfrey, V.G.; Faulkner, R.; Mirsky, A.E. Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. Proc. Natl. Acad. Sci. USA 1964, 51, 786–794. [Google Scholar] [CrossRef]
  240. Seto, E.; Yoshida, M. Erasers of histone acetylation: The histone deacetylase enzymes. Cold Spring Harb. Perspect. Biol. 2014, 6, a018713. [Google Scholar] [CrossRef]
  241. Sedmaki, K.; Karnam, K.; Sharma, P.; Mahale, A.; Routholla, G.; Ghosh, B.; Prakash Kulkarni, O. HDAC6 inhibition attenuates renal injury by reducing IL-1β secretion and RIP kinase mediated necroptosis in acute oxalate nephropathy. Int. Immunopharmacol. 2022, 110, 108919. [Google Scholar] [CrossRef]
  242. Yang, L.; Chen, S.; Xia, J.; Zhou, Y.; Peng, L.; Fan, H.; Han, Y.; Duan, L.; Cheng, G.; Yang, H.; et al. Histone deacetylase 3 facilitates TNFα-mediated NF-κB activation through suppressing CTSB induced RIP1 degradation and is required for host defense against bacterial infection. Cell Biosci. 2022, 12, 81. [Google Scholar] [CrossRef]
  243. Bozonet, S.M.; Magon, N.J.; Schwartfeger, A.J.; Konigstorfer, A.; Heath, S.G.; Vissers, M.C.M.; Morris, V.K.; Göbl, C.; Murphy, J.M.; Salvesen, G.S.; et al. Oxidation of caspase-8 by hypothiocyanous acid enables TNF-mediated necroptosis. J. Biol. Chem. 2023, 299, 104792. [Google Scholar] [CrossRef]
  244. Zheng, M.; Karki, R.; Vogel, P.; Kanneganti, T.D. Caspase-6 Is a Key Regulator of Innate Immunity, Inflammasome Activation, and Host Defense. Cell 2020, 181, 674–687.e13. [Google Scholar] [CrossRef]
  245. Wang, Z.; Feng, J.; Yu, J.; Chen, G. FKBP12 mediates necroptosis by initiating RIPK1-RIPK3-MLKL signal transduction in response to TNF receptor 1 ligation. J. Cell Sci. 2019, 132, jcs227777. [Google Scholar] [CrossRef]
  246. Gao, P.; Cao, M.; Jiang, X.; Wang, X.; Zhang, G.; Tang, X.; Yang, C.; Komuro, I.; Ge, J.; Li, L.; et al. Cannabinoid Receptor 2-Centric Molecular Feedback Loop Drives Necroptosis in Diabetic Heart Injuries. Circulation 2023, 147, 158–174. [Google Scholar] [CrossRef] [PubMed]
  247. Lee, S.A.; Chang, L.C.; Jung, W.; Bowman, J.W.; Kim, D.; Chen, W.; Foo, S.S.; Choi, Y.J.; Choi, U.Y.; Bowling, A.; et al. OASL phase condensation induces amyloid-like fibrillation of RIPK3 to promote virus-induced necroptosis. Nat. Cell Biol. 2023, 25, 92–107. [Google Scholar] [CrossRef] [PubMed]
  248. Galluzzi, L.; Kepp, O.; Chan, F.K.; Kroemer, G. Necroptosis: Mechanisms and Relevance to Disease. Annu. Rev. Pathol. 2017, 12, 103–130. [Google Scholar] [CrossRef]
  249. Xia, X.; Lei, L.; Wang, S.; Hu, J.; Zhang, G. Necroptosis and its role in infectious diseases. Apoptosis 2020, 25, 169–178. [Google Scholar] [CrossRef]
  250. Choi, M.E.; Price, D.R.; Ryter, S.W.; Choi, A.M.K. Necroptosis: A crucial pathogenic mediator of human disease. JCI Insight 2019, 4, e128834. [Google Scholar] [CrossRef] [PubMed]
  251. Mohammed, S.; Thadathil, N.; Selvarani, R.; Nicklas, E.H.; Wang, D.; Miller, B.F.; Richardson, A.; Deepa, S.S. Necroptosis contributes to chronic inflammation and fibrosis in aging liver. Aging Cell 2021, 20, e13512. [Google Scholar] [CrossRef] [PubMed]
  252. Tao, H.; Zhao, H.; Ge, D.; Liao, J.; Shao, L.; Mo, A.; Hu, L.; Xu, K.; Wu, J.; Mu, M.; et al. Necroptosis in pulmonary macrophages promotes silica-induced inflammation and interstitial fibrosis in mice. Toxicol. Lett. 2022, 355, 150–159. [Google Scholar] [CrossRef]
  253. Saeed, W.K.; Jun, D.W.; Jang, K.; Koh, D.H. Necroptosis signaling in liver diseases: An update. Pharmacol. Res. 2019, 148, 104439. [Google Scholar] [CrossRef] [PubMed]
  254. Khoury, M.K.; Gupta, K.; Franco, S.R.; Liu, B. Necroptosis in the Pathophysiology of Disease. Am. J. Pathol. 2020, 190, 272–285. [Google Scholar] [CrossRef]
  255. Della Torre, L.; Nebbioso, A.; Stunnenberg, H.G.; Martens, J.H.A.; Carafa, V.; Altucci, L. The Role of Necroptosis: Biological Relevance and Its Involvement in Cancer. Cancers 2021, 13, 684. [Google Scholar] [CrossRef]
  256. Lalaoui, N.; Brumatti, G. Relevance of necroptosis in cancer. Immunol. Cell Biol. 2017, 95, 137–145. [Google Scholar] [CrossRef]
  257. Liu, X.; Yao, Y.; Zhu, Y.; Lu, F.; Chen, X. Inhibition of Adipocyte Necroptosis Alleviates Fat Necrosis and Fibrosis After Grafting in a Murine Model. Aesthetic Surg. J. 2024, 44, NP585–NP605. [Google Scholar] [CrossRef]
  258. Harris, P.A.; Berger, S.B.; Jeong, J.U.; Nagilla, R.; Bandyopadhyay, D.; Campobasso, N.; Capriotti, C.A.; Cox, J.A.; Dare, L.; Dong, X.; et al. Discovery of a First-in-Class Receptor Interacting Protein 1 (RIP1) Kinase Specific Clinical Candidate (GSK2982772) for the Treatment of Inflammatory Diseases. J. Med. Chem. 2017, 60, 1247–1261. [Google Scholar] [CrossRef]
  259. Preston, S.P.; Stutz, M.D.; Allison, C.C.; Nachbur, U.; Gouil, Q.; Tran, B.M.; Duvivier, V.; Arandjelovic, P.; Cooney, J.P.; Mackiewicz, L.; et al. Epigenetic Silencing of RIPK3 in Hepatocytes Prevents MLKL-mediated Necroptosis From Contributing to Liver Pathologies. Gastroenterology 2022, 163, 1643–1657.e14. [Google Scholar] [CrossRef]
  260. Shi, Y.; Liu, J.; Hou, M.; Tan, Z.; Chen, F.; Zhang, J.; Liu, Y.; Leng, Y. Ursolic acid improves necroptosis via STAT3 signaling in intestinal ischemia/reperfusion injury. Int. Immunopharmacol. 2024, 138, 112463. [Google Scholar] [CrossRef] [PubMed]
  261. Shimizu, M.; Ohwada, W.; Yano, T.; Kouzu, H.; Sato, T.; Ogawa, T.; Osanami, A.; Toda, Y.; Nagahama, H.; Tanno, M.; et al. Contribution of MLKL to the development of doxorubicin-induced cardiomyopathy and its amelioration by rapamycin. J. Pharmacol. Sci. 2024, 156, 9–18. [Google Scholar] [CrossRef]
  262. Liang, J.; Tian, X.; Zhou, M.; Yan, F.; Fan, J.; Qin, Y.; Chen, B.; Huo, X.; Yu, Z.; Tian, Y.; et al. Shikonin and chitosan-silver nanoparticles synergize against triple-negative breast cancer through RIPK3-triggered necroptotic immunogenic cell death. Biomaterials 2024, 309, 122608. [Google Scholar] [CrossRef] [PubMed]
  263. Chen, J.; Zhao, L.; Xu, M.F.; Huang, D.; Sun, X.L.; Zhang, Y.X.; Li, H.M.; Wu, C.Z. Novel isobavachalcone derivatives induce apoptosis and necroptosis in human non-small cell lung cancer H1975 cells. J. Enzym. Inhib. Med. Chem. 2024, 39, 2292006. [Google Scholar] [CrossRef]
  264. Persenaire, C.; Babbs, B.; Yamamoto, T.M.; Nebbia, M.; Jordan, K.R.; Adams, S.; Lambert, J.R.; Bitler, B.G. VDX-111, a novel small molecule, induces necroptosis to inhibit ovarian cancer progression. Mol. Carcinog. 2024, 63, 1248–1259. [Google Scholar] [CrossRef]
  265. Vissers, M.; Heuberger, J.; Groeneveld, G.J.; Oude Nijhuis, J.; De Deyn, P.P.; Hadi, S.; Harris, J.; Tsai, R.M.; Cruz-Herranz, A.; Huang, F.; et al. Safety, pharmacokinetics and target engagement of novel RIPK1 inhibitor SAR443060 (DNL747) for neurodegenerative disorders: Randomized, placebo-controlled, double-blind phase I/Ib studies in healthy subjects and patients. Clin. Transl. Sci. 2022, 15, 2010–2023. [Google Scholar] [CrossRef] [PubMed]
  266. Majdi, A.; Aoudjehane, L.; Ratziu, V.; Islam, T.; Afonso, M.B.; Conti, F.; Mestiri, T.; Lagouge, M.; Foufelle, F.; Ballenghien, F.; et al. Inhibition of receptor-interacting protein kinase 1 improves experimental non-alcoholic fatty liver disease. J. Hepatol. 2020, 72, 627–635. [Google Scholar] [CrossRef]
  267. Cao, L.; Mu, W. Necrostatin-1 and necroptosis inhibition: Pathophysiology and therapeutic implications. Pharmacol. Res. 2021, 163, 105297. [Google Scholar] [CrossRef]
  268. Hofmans, S.; Devisscher, L.; Martens, S.; Van Rompaey, D.; Goossens, K.; Divert, T.; Nerinckx, W.; Takahashi, N.; De Winter, H.; Van Der Veken, P.; et al. Tozasertib Analogues as Inhibitors of Necroptotic Cell Death. J. Med. Chem. 2018, 61, 1895–1920. [Google Scholar] [CrossRef]
  269. Harris, P.A.; Marinis, J.M.; Lich, J.D.; Berger, S.B.; Chirala, A.; Cox, J.A.; Eidam, P.M.; Finger, J.N.; Gough, P.J.; Jeong, J.U.; et al. Identification of a RIP1 Kinase Inhibitor Clinical Candidate (GSK3145095) for the Treatment of Pancreatic Cancer. ACS Med. Chem. Lett. 2019, 10, 857–862. [Google Scholar] [CrossRef]
  270. Duan, X.; Liu, X.; Liu, N.; Huang, Y.; Jin, Z.; Zhang, S.; Ming, Z.; Chen, H. Inhibition of keratinocyte necroptosis mediated by RIPK1/RIPK3/MLKL provides a protective effect against psoriatic inflammation. Cell Death Dis. 2020, 11, 134. [Google Scholar] [CrossRef] [PubMed]
  271. Xia, K.; Zhu, F.; Yang, C.; Wu, S.; Lin, Y.; Ma, H.; Yu, X.; Zhao, C.; Ji, Y.; Ge, W.; et al. Discovery of a Potent RIPK3 Inhibitor for the Amelioration of Necroptosis-Associated Inflammatory Injury. Front. Cell Dev. Biol. 2020, 8, 606119. [Google Scholar] [CrossRef] [PubMed]
  272. Lee, S.Y.; Kim, H.; Li, C.M.; Kang, J.; Najafov, A.; Jung, M.; Kang, S.; Wang, S.; Yuan, J.; Jung, Y.K. Casein kinase-1γ1 and 3 stimulate tumor necrosis factor-induced necroptosis through RIPK3. Cell Death Dis. 2019, 10, 923. [Google Scholar] [CrossRef]
  273. Keusekotten, K.; Elliott, P.R.; Glockner, L.; Fiil, B.K.; Damgaard, R.B.; Kulathu, Y.; Wauer, T.; Hospenthal, M.K.; Gyrd-Hansen, M.; Krappmann, D.; et al. OTULIN antagonizes LUBAC signaling by specifically hydrolyzing Met1-linked polyubiquitin. Cell 2013, 153, 1312–1326. [Google Scholar] [CrossRef]
  274. Syed, R.U.; Afsar, S.; Aboshouk, N.A.M.; Salem Alanzi, S.; Abdalla, R.A.H.; Khalifa, A.A.S.; Enrera, J.A.; Elafandy, N.M.; Abdalla, R.A.H.; Ali, O.H.H.; et al. LncRNAs in necroptosis: Deciphering their role in cancer pathogenesis and therapy. Pathol. Res. Pract. 2024, 256, 155252. [Google Scholar] [CrossRef] [PubMed]
  275. Balusu, S.; Horré, K.; Thrupp, N.; Craessaerts, K.; Snellinx, A.; Serneels, L.; T’Syen, D.; Chrysidou, I.; Arranz, A.M.; Sierksma, A.; et al. MEG3 activates necroptosis in human neuron xenografts modeling Alzheimer’s disease. Science 2023, 381, 1176–1182. [Google Scholar] [CrossRef]
  276. Lu, H.; Xu, L.; Steriopoulos, J.; McLeod, P.; Huang, X.; Min, J.; Peng, T.; Jevnikar, A.M.; Zhang, Z.X. An acidic pH environment converts necroptosis to apoptosis. Biochem. Biophys. Res. Commun. 2024, 725, 150215. [Google Scholar] [CrossRef]
Figure 1. The signaling pathway of necroptosis and its PTMs. (A) Binding of cell surface TNFR1, DR3, DR6, and their respective ligands TNF, TLA1, and APP recruits the formation of Complex I, which promotes the activation of the NF-κB and MAPK pathways, the expression of inflammatory genes, and cell survival. After Complex I is destabilized, it can form Complex II, including Complex IIa and IIb, which induce apoptosis and RIPK1-dependent apoptosis, respectively. When caspase-8 activity is inhibited, a ripoptosome can be formed in the cell, which further forms a necrosome and induces necroptosis. The PTMs of RIPK1, RIPK3, and MLKL and their corresponding modified proteins are also labeled in the figure. (B) On the left side of the figure, FAS and TRAILR1/2 bind to their respective ligands FASL and TRAIL1/2 to first form the death-inducing signaling complex (DISC), which can induce apoptosis and necroptosis when caspase-8 is inhibited. After separation from the receptor, a complex similar to Complex I can be formed in the cytoplasm as the FADDosome, which can promote the expression of the NF-κB and MAPK pathways and cell survival. On the right side of the figure, only caspase-8 inhibition is shown. When caspase-8 is not inhibited, TLR3, TLR4, and ZBP1 can also form Complex I-like complexes in the cytoplasm upon binding to their corresponding ligands, promoting cell survival. When caspase-8 is inhibited, TLR3 and TLR4 promote its autophosphorylation and the recruitment of MLKL through direct binding of TRIF to RIPK3, while ZBP1 binds to RIPK3 directly, which promotes necroptosis. Interferon (IFN) can promote necroptosis through multiple pathways.
Figure 1. The signaling pathway of necroptosis and its PTMs. (A) Binding of cell surface TNFR1, DR3, DR6, and their respective ligands TNF, TLA1, and APP recruits the formation of Complex I, which promotes the activation of the NF-κB and MAPK pathways, the expression of inflammatory genes, and cell survival. After Complex I is destabilized, it can form Complex II, including Complex IIa and IIb, which induce apoptosis and RIPK1-dependent apoptosis, respectively. When caspase-8 activity is inhibited, a ripoptosome can be formed in the cell, which further forms a necrosome and induces necroptosis. The PTMs of RIPK1, RIPK3, and MLKL and their corresponding modified proteins are also labeled in the figure. (B) On the left side of the figure, FAS and TRAILR1/2 bind to their respective ligands FASL and TRAIL1/2 to first form the death-inducing signaling complex (DISC), which can induce apoptosis and necroptosis when caspase-8 is inhibited. After separation from the receptor, a complex similar to Complex I can be formed in the cytoplasm as the FADDosome, which can promote the expression of the NF-κB and MAPK pathways and cell survival. On the right side of the figure, only caspase-8 inhibition is shown. When caspase-8 is not inhibited, TLR3, TLR4, and ZBP1 can also form Complex I-like complexes in the cytoplasm upon binding to their corresponding ligands, promoting cell survival. When caspase-8 is inhibited, TLR3 and TLR4 promote its autophosphorylation and the recruitment of MLKL through direct binding of TRIF to RIPK3, while ZBP1 binds to RIPK3 directly, which promotes necroptosis. Interferon (IFN) can promote necroptosis through multiple pathways.
Biomolecules 15 00549 g001aBiomolecules 15 00549 g001b
Figure 2. The structure of RIPK1, RIPK3, and MLKL of humans and mice and their respective PTM sites and their contribution to necroptosis. (A) PTMs of human RIPK1, RIPK3 and MLKL. (B) PTMs of mouse RIPK1, RIPK3 and MLKL. PTMs at different structural domains with different sites are labeled in the figure. Red boxes indicate proteins with added PTMs, blue boxes indicate proteins with removed PTMs, and arrow color and direction indicate contribution to necroptosis.
Figure 2. The structure of RIPK1, RIPK3, and MLKL of humans and mice and their respective PTM sites and their contribution to necroptosis. (A) PTMs of human RIPK1, RIPK3 and MLKL. (B) PTMs of mouse RIPK1, RIPK3 and MLKL. PTMs at different structural domains with different sites are labeled in the figure. Red boxes indicate proteins with added PTMs, blue boxes indicate proteins with removed PTMs, and arrow color and direction indicate contribution to necroptosis.
Biomolecules 15 00549 g002aBiomolecules 15 00549 g002b
Table 1. PTMs of RIPK1.
Table 1. PTMs of RIPK1.
SpeciesSiteDomainProteinNecroptosisReferences
Ubiquitination
HumanK377IntermediatecIAP1, cIAP2[42]
[114]
MouseK376IntermediatecIAP1, cIAP2[42]
[114]
Human and MouseUnknownUnknowncIAP1(UBA domain)[113]
Human and MouseUnknownUnknownLUBAC[44]
[45]
[46]
[47]
HumanK377IntermediateMIB2[118]
HumanK634Death domainMIB2[118]
MouseK376IntermediatePARdU[119]
HumanK377IntermediatePARdU[119]
HumanK316IntermediateMG53[120]
HumanK604, K627Death domainMG53[120]
HumanK377IntermediateA20, OTULIN CYLD, USP21, [62]
[63]
[64]
[65]
[66]
[67]
MouseK376IntermediateA20, OTULIN CYLD, USP21, [62]
[63]
[64]
[65]
[66]
[67]
Human and MouseUnknownUnknownZFP91[68]
HumanK115Kinase domainPELI1[121]
MouseK115Kinase domainPELI1[121]
HumanK158Kinase domainc-Cbl[122]
HumanK571IntermediateCHIP[123]
HumanK604, K627Death domainCHIP[123]
MouseK612Death domainUnknown[124]
HumanK13N-terminal (prior to Kinase domain)UnknownUnknown[124]
HumanK302, K305, K306, K396, K530, K565, K585IntermediateUnknownUnknown[124]
HumanK30, K45, K49, K65, K77, K97, K105, K132, K137, K140, K153, K163, K167, K184, K185, K204, K265, K284Kinase domainUnknownUnknown[124]
HumanK596, K599, K642, K648Death domainUnknownUnknown[124]
MouseK13, K20N-terminal (prior to Kinase domain)UnknownUnknown[124]
MouseK306, K307, K392, K395, K429, K519, K550, K584IntermediateUnknownUnknown[124]
MouseK30, K45, K46, K65, K77, K105, K137, K140, K153, K163, K167, K205Kinase domainUnknownUnknown[124]
MouseK589, K619, K627, K633Death domainUnknownUnknown[124]
Phosphorylation
HumanS166Kinase domainRIPK1[125]
[126]
HumanS161Kinase domainRIPK1[94]
HumanS14, S15, S20N-terminal (prior to Kinase domain)RIPK1No effect[127]
MouseS14, S15N-terminal (prior to Kinase domain)RIPK1No effect[127]
MouseT169Kinase domainRIPK1Unknown[127]
HumanS25Kinase domainIKKα/β[127]
MouseS25Kinase domainIKKα/β[127]
HumanS320IntermediateMK2[49]
[50]
[51]
MouseS321, S336IntermediateMK2[49]
[50]
[51]
MouseS321IntermediateTAK1[128]
HumanY384IntermediateJAK1, SRC[52]
MouseY383IntermediateJAK1, SRC[52]
HumanS357IntermediateULK1[53]
HumanT147Kinase domainTBK1[54]
MouseS189, S190Kinase domainTBK1[54]
HumanS416IntermediateAMPK[55]
MouseS415IntermediateAMPK[55]
HumanS89Kinase domainUnknown[129]
Mouse, rat, xenopus, zebrafishS89Kinase domainUnknown[129]
HumanS25Kinase domainPPP1R3G[130]
MouseS25Kinase domainPPP1R3G[130]
HumanUnknownUnknownSMYD2[131]
Glycosylation
HumanS331IntermediateOGT[132]
MouseS332IntermediateOGT[132]
HumanR603Death domainNIeB2[133]
Acetylation
HumanK115Kinase domainSIRT1, SIRT2[134]
HumanK115Kinase domainHAT1, HAT4[134]
HumanK625, K627, K642, K648Death domainSIRT1, SIRT2[134]
HumanK625, K627, K642, K648Death domainHAT1, HAT4[134]
HumanK596, K599IntermediateSIRT1, SIRT2[134]
HumanK596, K599IntermediateHAT1, HAT4[134]
Disulfide bonds
HumanC257, C268Kinase domainROS[94]
HumanC586IntermediateROS[94]
Caspase cleavage
HumanD324IntermediateCaspase-6, -8, -10[135]
[136]
[137]
MouseD325IntermediateCaspase-6, -8, -10[135]
[136]
[137]
SUMOylation
HumanK550IntermediateSENP1[138]
Downward-pointing arrows mean that necroptosis is inhibited, and upward-pointing arrows mean that necroptosis is promoted.
Table 2. PTMs of RIPK3.
Table 2. PTMs of RIPK3.
SpeciesSiteDomainProteinNecroptosisReferences
Ubiquitination
Human and MouseK5N-terminal (prior to Kinase domain)A20[139]
[140]
MouseK158, K287Kinase domainUnknown[139]
MouseK307IntermediateUnknown[139]
MouseK469C-terminal (after RHIM domain)Unknown[139]
MouseK359IntermediateUnknown[139]
HumanK42Kinase domainUSP22No effect[142]
HumanK351IntermediateUSP22No effect[142]
HumanK518C-terminal (after RHIM domain)USP22[142]
HumanK55Kinase domainCHIP[123]
HumanK363IntermediateCHIP[123]
HumanK363IntermediatePELI1[143]
HumanK501C-terminal (after RHIM domain)TRIM25[144]
HumanK264Kinase domainUnknown[145]
HumanK197Kinase domainParkin[146]
HumanK302, K364IntermediateParkin[146]
Phosphorylation
HumanS199, S211, S215, S227, T182, T215Kinase domainRIPK3[147]
[148]
[149]
[150]
[151]
[150]
MouseS204, S232, T231Kinase domainRIPK3[147]
[148]
[149]
[150]
Human and MouseS164, T165Kinase domainRIPK3[152]
HumanT224Kinase domainUnknown[150]
HumanS227Kinase domainCK1α, CK1δ, CK1ε[153]
MouseS204Kinase domainRIPK1[129]
Human and MouseS164, T165Kinase domainRIPK3-Hsp90/CDC37[152]
MouseS369IntermediatePLK1[127]
MouseS232, T231Kinase domainPpm1b[154]
Glycosylation
HumanT467RHIM domainOGT[155]
[156]
[157]
[158]
[159]
Methylation
HumanR486C-terminal (after RHIM domain)PRMT1, PRMT5[160]
[161]
MouseR479C-terminal (after RHIM domain)PRMT1, PRMT5[160]
[161]
Caspase cleavage
HumanD328IntermediateCaspase-8[162]
MouseD328IntermediateCaspase-8[162]
RatD328IntermediateCaspase-8[162]
Nitrosylation
HumanC119Kinase domainNO[163]
[164]
Downward-pointing arrows mean that necroptosis is inhibited, and upward-pointing arrows mean that necroptosis is promoted.
Table 3. PTMs of MLKL.
Table 3. PTMs of MLKL.
SpeciesSiteDomainProteinNecroptosisReferences
Ubiquitination
MouseK51, K774HBUnknown[166]
MouseK172BraceUnknown[166]
MouseK219PseudokinaseUnknown[166]
HumanK230PseudokinaseUnknown[166]
MouseK9, K51, K694HBUSP21[167]
HumanK504HBITCH[168]
MouseK50, K514HBITCH[168]
Human and MouseUnknownUnknownSkp2[169]
Phosphorylation
HumanT357, S358PseudokinaseRIPK3[165]
[171]
[172]
HumanT357, S358PseudokinaseCAMK2/CaMKII[173]
MouseS345PseudokinaseRIPK3[165]
[171]
[172]
MouseS347PseudokinaseRIPK3No effect[174]
[175]
MouseS158BraceRIPK3[174]
[175]
MouseS228, S248PseudokinaseRIPK3[174]
[175]
MouseT349PseudokinaseRIPK3Unknown[174]
[175]
MouseS441PseudokinaseUnknownUnknown[176]
Human and MouseY376PseudokinaseTAM[177]
HumanS834HBUnknown[178]
MouseS824HBUnknown[178]
Disulfide bonds
HumanC864HBMLKL[179]
HumanC864HBTrx1[180]
Downward-pointing arrows mean that necroptosis is inhibited, and upward-pointing arrows mean that necroptosis is promoted.
Table 4. PTMs of TRADD, TRAILR, TNFR1, FADD, caspase-8, and cFLIP.
Table 4. PTMs of TRADD, TRAILR, TNFR1, FADD, caspase-8, and cFLIP.
SpeciesSiteDomainProteinNecroptosisReferences
TRADD
Glycosylation
HumanR235, R245Death domainSseK1[183]
[184]
MouseR233Death domainSseK1[183]
HumanR235Death domainNIeB[185]
TRAILR
Glycosylation
HumanR359Death domainSseK3[183]
MouseR239Death domainSseK3[183]
TNFR1
Glycosylation
HumanR376Death domainSseK3[183]
MouseR376Death domainSseK3[183]
Ubiquitination
Human and Mouse UnknownUnknownRNF8[186]
FADD
Ubiquitination
HumanK149, K153Death domainCHIP[181]
Human and MouseUnknownUnknownMKRN1[182]
Phosphorylation
HumanS194Death domainCKIαNo effect[187]
[188]
MouseS191Death domainCKIαNo effect[187]
[188]
HumanS200Death domainCK2No effect[189]
HumanS203Death domainAur-A, Plk1No effect[190]
SUMOylation
HumanK120, K125, K149Death domainPIAS3Unknown[191]
Caspase-8
Ubiquitination
HumanK461C-terminal domain (P10 subunit)CUL3[192]
Human and MouseUnknownUnknownTRIM13[193]
HumanK215IntermediateHECTD3[194]
HumanK224, K229, K231C-terminal domain (P18 subunit)TRAF2Unknown[195]
Phosphorylation
MouseT265C-terminal domain (P18 subunit)p90 RSK[196]
HumanT263C-terminal domain (P18 subunit)P90 RSK (RSK2)Unknown[197]
HumanT273C-terminal domain (P18 subunit)Plk3Unknown[198]
HumanS347C-terminal domain (P18 subunit)P38-MAPKUnknown[199]
HumanY380C-terminal domain (Intermediate)SrcUnknown[200]
HumanS387C-terminal domain (Intermediate)CDK1Unknown[201]
HumanY448C-terminal domain (P10 subunit)LynUnknown[202]
Nitrosylation
Human and MouseUnknownUnknownNOUnknown[203]
SUMOylation
HumanK156DED domainUnknownUnknown[204]
cFLIP
Ubiquitination
HumanK351, K353C-terminal domain (P20 subunit)LUBAC[205]
HumanK351, K353, K381, K386, K389, K460, K462, K473, K474C-terminal domainMIB2[206]
HumanK167, K192, K195IntermediateUnknownUnknown[207]
[208]
[209]
Human and MouseUnknownUnknownSCFSkp2Unknown[210]
Human and MouseUnknownUnknownITCHUnknown[209]
Human and MouseUnknownUnknownUsp27xUnknown[211]
Human and MouseUnknownUnknownTRIM28Unknown[211]
Phosphorylation
HumanT166IntermediateUnknownUnknown[207]
[208]
[209]
HumanS193IntermediatePKCUnknown[207]
[208]
[209]
HumanS273C-terminal domain (P20 subunit)Akt1Unknown[212]
Nitrosylation
HumanC254, C259C-terminal domain (P20 subunit)NO[213]
Downward-pointing arrows mean that necroptosis is inhibited, and upward-pointing arrows mean that necroptosis is promoted.
Table 5. Approved and investigational drugs targeting the necroptosis pathway: targets, development stages, and associated diseases/indications.
Table 5. Approved and investigational drugs targeting the necroptosis pathway: targets, development stages, and associated diseases/indications.
Drug NameTargetDevelopment StageDisease/IndicationReferences
GSK2982772RIPK1 inhibitorPhase II Clinical TrialPsoriasis, Ulcerative Colitis, Rheumatoid Arthritis[258]
SAR443060 (DNL758)RIPK1 inhibitorPhase I Clinical TrialNeurodegenerative Disorders[265]
RIPA-56RIPK1 inhibitorPreclinicalNon-Alcoholic Fatty Liver Disease[266]
Necrostatin-1s (Nec-1s)RIPK1 inhibitorPreclinicalNeurodegenerative Disorders, Ischemia–Reperfusion Injury, Cardiovascular Diseases, Renal Diseases, Liver Diseases, etc.[267]
TozasertibRIPK1 inhibitorPhase II Clinical TrialMouse Model of TNF-α-induced systemic inflammatory response syndrome[268]
GSK3145095RIPK1 inhibitorDiscontinued (Phase I)Pancreatic Cancer, Colorectal Cancer[269]
Necrosulfonamide (NSA)MLKL inhibitorPreclinicalCardiac Arrest, Spinal Cord Injury, Intracerebral Hemorrhage[270]
GSK’872RIPK3 inhibitorPreclinicalIschemia–Reperfusion Injury, Sepsis, Neurodegenerative Diseases, Psoriasis, Ulcerative Colitis, Rheumatoid Arthritis[271]
VDX-111RIPK1 agonistPreclinicalOvarian Cancer[264]
SHK+Chi-Ag NPsRIPK3 agonistPreclinicalTriple-Negative Breast Cancer[262]
Seventeen isobavachalcone (IBC) derivatives (1-17)RIPK3, MLKL agonistPreclinicalNon-Small Cell Lung Cancer[263]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xiao, H.; Han, Z.; Xu, M.; Gao, X.; Qiu, S.; Ren, N.; Yi, Y.; Zhou, C. The Role of Post-Translational Modifications in Necroptosis. Biomolecules 2025, 15, 549. https://doi.org/10.3390/biom15040549

AMA Style

Xiao H, Han Z, Xu M, Gao X, Qiu S, Ren N, Yi Y, Zhou C. The Role of Post-Translational Modifications in Necroptosis. Biomolecules. 2025; 15(4):549. https://doi.org/10.3390/biom15040549

Chicago/Turabian Style

Xiao, Hao, Zeping Han, Min Xu, Xukang Gao, Shuangjian Qiu, Ning Ren, Yong Yi, and Chenhao Zhou. 2025. "The Role of Post-Translational Modifications in Necroptosis" Biomolecules 15, no. 4: 549. https://doi.org/10.3390/biom15040549

APA Style

Xiao, H., Han, Z., Xu, M., Gao, X., Qiu, S., Ren, N., Yi, Y., & Zhou, C. (2025). The Role of Post-Translational Modifications in Necroptosis. Biomolecules, 15(4), 549. https://doi.org/10.3390/biom15040549

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop