Plants utilize cell-surface localized and intracellular leucine-rich repeat (LRR) immune receptors to detect pathogens and to activate defense responses, including transcriptional reprogramming and the initiation of a form of programmed cell death of infected cells. Cell death initiation is mainly associated with the activation of nucleotide-binding LRR receptors (NLRs). NLRs recognize the presence or cellular activity of pathogen-derived virulence proteins, so-called effectors. Effector-dependent NLR activation leads to the formation of higher order oligomeric complexes, termed resistosomes. Resistosomes can either form potential calcium-permeable cation channels at cellular membranes and initiate calcium influxes resulting in activation of immunity and cell death or function as NADases whose activity is needed for the activation of downstream immune signaling components, depending on the N-terminal domain of the NLR protein. In this mini-review, the current knowledge on the mechanisms of NLR-mediated cell death and resistance pathways during plant immunity is discussed.

Beside physical barriers, including bark, leaf hairs (trichomes), waxy cuticles and plant cell walls, plant-invading (micro-)organisms have to face a highly sophisticated and heavily interconnected two-tiered, receptor based innate immune system [1]. At the cell surface, plasma membrane spanning pattern-recognition receptors (PRRs) perceive the presence of invading organisms, like insects, fungi, oomycetes, bacteria or parasitic plants, by sensing so-called microbe- or host-derived damage-associated molecular patterns (MAMPs/DAMPs) [2,3]. PRRs are receptor kinases (RKs) or receptor-like proteins (RLPs), and MAMP/DAMP perception leads to a cascade of immune responses, commonly referred to as pattern-triggered immunity (PTI) [3]. PTI induces a rapid calcium (Ca2+) influx, production of reactive oxygen species (ROS), the release of anti-microbial molecules and proteins into the apoplast (the intercellular space), initiation of mitogen-activated protein kinase (MAPK) signaling, transcriptional reprogramming and the up-regulation of defense genes, among others [3,4]. This ‘cocktail’ of induced responses is often sufficient to successfully defend host non-adapted microbes and, to a limited extent, host-adapted pathogens. To interfere with PTI and to counter resistance, thus promote pathogenesis, adapted pathogens evolved effector molecules and proteins that they secrete into host cells or the apoplast [5,6].

However, in resistant plants, effectors can be perceived either directly or indirectly by a second class of immune receptors, the intracellular nucleotide-binding (NB) leucine-rich repeat (LRR) receptor (NLR) family. NLR activation through effector recognition initiates a robust immune response referred to as effector-triggered immunity (ETI) [7,8]. ETI responses are both similar and distinct in their molecular signature to PTI responses [9,10]. Initiation of ETI by activated NLRs triggers a prolonged calcium influx [11]and often results in the induction of the hypersensitive response (HR), a form of programmed cell death at the site of infection that further restricts pathogen proliferation and spread [12]. NLR-induced HR triggers the establishment of systemic acquired resistance (SAR), which primes systemic tissues to faster and more robustly respond to secondary infections [13]. It is important to emphasize that during natural infections ETI is often, if not always, happening in the presence of activated PTI — this is also the case in most experimental set-ups used to study or assay for ETI. Exceptions may be the use of transgenic plants expressing pathogenic effectors in an inducible fashion [9,10], or during viral infections — although more and more evidences are being presented that suggest that there is a PTI(-like) activation during viral infections [14–16].

NLR-like proteins can also be found in animals and fungi, where they convergently evolved as intracellular surveillance proteins to detect different kinds of pathogens and to function in allorecognition, respectively [7,8,17]. The first plant NLR genes were cloned in the early 1990s from tobacco (Nicotiana tabacum; [18]) and Arabidopsis thaliana (hereafter: Arabidopsis; [19,20]), and since then, our knowledge of how these important proteins contribute to immunity has tremendously expanded [21]. Especially research made in the last couple of years has enabled the field to make a huge step forward in understanding NLR activation, function in cell death initiation and immune signaling. Intriguingly, these results demonstrate that the activation of at least some NLRs results in oligomerization, which enables NLRs to form either an active holoenzyme or a potential cation channel at cellular membranes, depending on their N-terminal domain architecture.

This mini-review aims at summarizing some milestones in NLR biology. It further provides an overview of the recent achievements in determining how NLRs might mobilize antimicrobial defenses, while simultaneously it outlines the still unanswered questions of NLR-mediated immunity.

Most plant NLRs, as well as the metazoan NLRs, have a modular and conserved tripartite domain architecture, with a variable N-terminal domain, a central nucleotide-binding and oligomerization domain (NB or NOD), and a C-terminal leucine-rich repeat (LRR) domain. The plant NB domain (also referred to as NB-ARC, nucleotide-binding adaptor shared by apoptotic protease activating factor-1 (Apaf-1), Resistance (R)-protein, and cell death abnormal (CED) protein) is subdivided into three subdomains: the NB domain, an ARC1 or helical domain 1 (HD1) and an ARC2 or winged-helix domain (WHD), and belongs to the AAA+ ATPase superfamily [22,23]. The variable N-terminal domain can either be a Toll/Interleukin-1 receptor/Resistance protein (TIR) domain, a coiled-coil (CC) or a resistance to powdery mildew 8 (RPW8)-like CC (CCR) domain [24]. The different NLR classes are therefore also termed as TNLs (with a TIR domain), CNLs (with a CC domain) and RNLs (harboring a CCR domain). The CC and CCR domains share structural homology with the N-terminal 4-helical bundle of mammalian and plant mixed-linage kinase like (MLKL) proteins and pore-forming fungal HET-s/Helo domain containing proteins [24–29]. TIR domains have been shown to be structurally and functionally conserved — at least to some extent — and can be found in prokaryotic and other eukaryotic organisms [30,31], where TIR domain-containing proteins have also been implicated in immunity and initiation of cell death responses [32–34]. Plant NLR LRR domains have the classical horseshoe shaped structure with parallel beta-strands facing the concave (inner) side and on the convex side with variable secondary structures, including alpha helical regions [35].

Intramolecular interactions between the different domains keep NLR proteins in an inactive, but signaling competent state, in which the NB-ARC domain binds adenosine diphosphate (ADP) via conserved and functionally important motifs. The NB-ARC domain likely functions as a molecular switch, transitioning the NLR from an inactive ADP-bound to an active adenosine triphosphate (ATP)-bound state [23,36,37]. However, ATP-binding to the NB-ARC domain might not always be associated with NLR activity, as recently shown for the Arabidopsis TNL Recognition of Peronospora parasitica 1A (RPP1) [38]. The C-terminal LRR domain often plays an important role in the autoinhibitory process via interactions with the NB-ARC and/or the N-terminal domains [39]. Deletion of the LRR domain can, for example, result in the constitutive activation of NLRs in both plants and animals [40–42]. In addition to this regulatory function, many LRR domains or C-terminal LRR-adjacent domains (C-JID) are also involved in direct binding of effectors [38,43]. Cryogenic electron microscopy (cryo-EM) analyses have revealed that the LRR domain together with C-CJID domain of two TNLs, Recognition of XopQ1 (ROQ1) and RPP1A, is required for direct effector recognition [38,43]. Furthermore, a recent survey of NLR complements of over 60 Arabidopsis ecotypes and more than 50 Brachypodium distachyon lines confirmed that LRR domains of many NLRs show high allelic diversity especially in residues or regions predicted to be specificity determining [44]. This suggests that the LRR modularity functions to generate binding sites to non-self as well as self-molecules, and thus NLRs containing such LRR domains might be a source for new direct and indirect effector-binding/sensing immune receptors [44]. The N-terminal CC/CCR and TIR domains mediate NLR signaling and are required or at least participate in NLR oligomerization, and can therefore be considered as the signaling domains. A number of studies demonstrated that overexpression of isolated CC/CCR or TIR domains was sufficient to induce cell death or auto-immunity [28,39,45–48], further supporting their signaling function.

Within the last two years the publication of cryo-EM structures of three full length plant NLRs, the Arabidopsis CNL ZAR1 [23,49] and TNL RPP1A [38], and the N. benthamiana TNL ROQ1[43], have been released, which enabled us to better understand the molecular events downstream of effector recognition that lead to NLR activation. NLR proteins perceive the presence or activity of effectors either through direct interaction or by sensing modifications of host effector targets or decoys of such [50]. Direct effector recognition is often mediated by the LRR domain, a C-terminal extension termed the C-terminal jelly roll and Ig-like domain (C-JID) or Post-LRR (PL) domain or through a non-canonical ‘integrated domain’ (ID) that can often be found at the C-terminal end, but can also be integrated somewhere else in the protein domain structure [38,43,51–53]. In a recent study, structure-guided engineering was used to expand the recognition profile of the rice NLR Pikp-1 to variants of a rice blast pathogen effector by mutating residues important for effector binding in the ID of Pikp-1 [54]. This clearly demonstrates how structure-function analysis enable us to understand effector recognition and how this knowledge can be directly applied to generate (modified) improved immune receptors for agriculture.

Indirect recognition of effectors is mediated by NLRs that guard effector targets or that interact inducibly with effector-modified host proteins. Often these host proteins have a function in immunity, and one of the best-studied examples is the small Arabidopsis protein RPM1-interacting protein 4 (RIN4), which is guarded by two CNLs RPM1 and RPS2 [55–57]. RIN4 is targeted by numerous bacterial effectors which post translationally modify specific RIN4 residues or proteolytically cleave the protein [55,57–60]. It is believed that these modifications lead to conformational changes in the preformed NLR-guardee complex that than result in NLR activation and downstream signaling. How the different modifications on RIN4 activate RPM1 or RPS2 in detail requires further analysis. Given that activation of the two CNLs require different modifications on RIN4 it can be assumed that the molecular mechanism(s) of activation might vary as well. Thus, how the conformational change allowing the exchange of ADP by ATP and the oligomerization is induced may differ, but the outcome, formation of a functional ‘resistosome’, may be similar.

How effector-recognition can mechanistically induce NLR activation was impressively demonstrated by recent cryo-EM studies that presented the structure of inactive and effector-activated Arabidopsis CNL ZAR1 [23,49,61]. ZAR1 can detect the presence/function of a large number of effectors from different pathogens [62]. Inactive ZAR1 is bound to ADP and exists in a monomeric state, where it can interact with various receptor-like cytoplasmic kinases (RLCKs) of the RLCK-XII subfamily. ZAR1-interacting RLCKs can be considered as ‘adaptor proteins’ in this complex, because they function as the docking-site for another subfamily of RLCKs (subfamily RLCK-VII), which, upon modifications by ZAR1-recognized effectors, bind to the preformed ZAR1–RLCK-XII complex. Interaction of the RLCK-VII member with the ZAR1–RLCK-XII complex removes a structural inhibition on the ZAR1 NB-ARC domain, which was exerted by parts of the RLCK-XII protein, and in turn leads to the exchange of ADP by ATP and eventually to oligomerization of the tertiary complexes into a so-called resistosome [49](Figure 1). Comparisons between the cryo-EM structure of inactive and active ZAR1 uncovered the mechanism of ZAR1 activation in fantastic detail and suggests a dynamic structural change from an inactive, to an intermediate and further into an active, signaling competent pentameric complex [23,61].

NLR mediated immune signaling.

Figure 1.
NLR mediated immune signaling.

Schematic presentation of CNL (Arabidopsis ZAR1 (1)), TNL (Arabidopsis RPP1 (2)) and helper NLR (Arabidopsis RNL NRG1.1 (3)) cell death and resistance signaling. (1) The bacterial effector AvrAc uridylates the receptor-like cytoplasmic kinase (RLCK) PBL2 resulting in PBL2UMP. PBL2UMP binds to the preformed ZAR1–RKS1 complex via interaction with RKS1, which induces a conformational change in the ZAR1–RKS1 complex resulting in nucleotide exchange (ADP to ATP) in the ZAR1 nucleotide-binding domain (NB-ARC) and eventually to the pentamerization of this tetrameric ZAR1–RKS1–PBL2UMP complex into a so-called resistosome. The ZAR1 resistosome relocalizes to the plasma membrane and forms a calcium permeable cation channel via the N-terminal coiled-coil (CC) domains of the ZAR1 proteins [61,63]. (2) RPP1 recognizes and directly binds the oomycete-derived effector ATR1 via its leucine-rich repeat domain and the C-terminal jelly roll and Ig-like domain (C-JID). ATR1 binding induces the tetramerization of four RPP1-ATR1 sub-complexes into a RPP1 resistosome and thus a conformational activation or RPP1. Interactions between the RPP1 NB-ARC domains and the N-terminal TIR domains stabilize the resistosome complex without the necessity of ATP binding. RPP1 tetramerization results in the opening up of the NADase active site, transforming the TIR domains into an active holoenzyme [38]. Hydrolysis of NAD+ results in the production of various potential signaling molecules, such as variant cyclic ADPR (v-cADPR), that are hypothesized to activate and induce interaction of the TNL downstream signaling components EDS1, PAD4, SAG101 and the RNLs (helper NLRs). (3) How the EDS1/PAD4 or EDS1/SAG101 RNL components are activated by TIR domain NADase activity is not clear. Recently, it was shown that the Arabidopsis NRG1.1/NRG1A and ADR1 (both RNLs) can form activation-dependent oligomers and function also as calcium permeable cation channels at the plasma membrane [27,64]. Note: There is no experimental evidence yet that RNL resistosomes form ring-like structures (here a hexamere), which are similar to the pentameric ZAR1 ring-like resistosome. (4) It is known that NLR-activated effector-triggered immunity leads to calcium influx, which is required for cell death and resistance signaling, but how channel or pore formation by ZAR1 or NRG1.1/NRG1A is causing cell death (cell collapse) is not understood. It is possible that the calcium influx activates other cation channels important for immunity (for example the CNGCs) and components required for cell collapse, and that damage-associated molecular patterns (DAMPs) might also be released by the NLR formed pores to create signals enhancing immune responses that may lead to systemic resistance and maybe cell death. Created with BioRender.com.

Figure 1.
NLR mediated immune signaling.

Schematic presentation of CNL (Arabidopsis ZAR1 (1)), TNL (Arabidopsis RPP1 (2)) and helper NLR (Arabidopsis RNL NRG1.1 (3)) cell death and resistance signaling. (1) The bacterial effector AvrAc uridylates the receptor-like cytoplasmic kinase (RLCK) PBL2 resulting in PBL2UMP. PBL2UMP binds to the preformed ZAR1–RKS1 complex via interaction with RKS1, which induces a conformational change in the ZAR1–RKS1 complex resulting in nucleotide exchange (ADP to ATP) in the ZAR1 nucleotide-binding domain (NB-ARC) and eventually to the pentamerization of this tetrameric ZAR1–RKS1–PBL2UMP complex into a so-called resistosome. The ZAR1 resistosome relocalizes to the plasma membrane and forms a calcium permeable cation channel via the N-terminal coiled-coil (CC) domains of the ZAR1 proteins [61,63]. (2) RPP1 recognizes and directly binds the oomycete-derived effector ATR1 via its leucine-rich repeat domain and the C-terminal jelly roll and Ig-like domain (C-JID). ATR1 binding induces the tetramerization of four RPP1-ATR1 sub-complexes into a RPP1 resistosome and thus a conformational activation or RPP1. Interactions between the RPP1 NB-ARC domains and the N-terminal TIR domains stabilize the resistosome complex without the necessity of ATP binding. RPP1 tetramerization results in the opening up of the NADase active site, transforming the TIR domains into an active holoenzyme [38]. Hydrolysis of NAD+ results in the production of various potential signaling molecules, such as variant cyclic ADPR (v-cADPR), that are hypothesized to activate and induce interaction of the TNL downstream signaling components EDS1, PAD4, SAG101 and the RNLs (helper NLRs). (3) How the EDS1/PAD4 or EDS1/SAG101 RNL components are activated by TIR domain NADase activity is not clear. Recently, it was shown that the Arabidopsis NRG1.1/NRG1A and ADR1 (both RNLs) can form activation-dependent oligomers and function also as calcium permeable cation channels at the plasma membrane [27,64]. Note: There is no experimental evidence yet that RNL resistosomes form ring-like structures (here a hexamere), which are similar to the pentameric ZAR1 ring-like resistosome. (4) It is known that NLR-activated effector-triggered immunity leads to calcium influx, which is required for cell death and resistance signaling, but how channel or pore formation by ZAR1 or NRG1.1/NRG1A is causing cell death (cell collapse) is not understood. It is possible that the calcium influx activates other cation channels important for immunity (for example the CNGCs) and components required for cell collapse, and that damage-associated molecular patterns (DAMPs) might also be released by the NLR formed pores to create signals enhancing immune responses that may lead to systemic resistance and maybe cell death. Created with BioRender.com.

Close modal

An activation mechanism similarly depending on effector recognition and oligomerization (specifically tetramerization) was recently postulated for the two plant TNLs RPP1A and ROQ1 [38,43]. Cryo-EM structures of effector-bound ROQ1 and RPP1A defined a tetramerization into a symmetric dimer of two dimers that was essential for cell death activity and immune signaling [38,43] (Figure 1). Complex formation, self-association and also heterooligomerization was shown to be important for the activity of many other NLRs, including CNLs, TNLs and the RNLs [27,28,48,65–69]. Interestingly, many plant NLRs already self-associate in the absence of effector-recognition — thus, in the pre-activation state (El Kasmi lab unpublished results and [48,64,65,68,70]). In these cases, sensing of the effector most likely induces conformational changes in the preformed dimers or oligomers allowing nucleotide exchange and possibly the interaction with downstream signaling components or the formation of resistosome-like complexes. It would be interesting to see whether for such NLRs a shift into high-molecular weight complexes could be observed upon activation, which would suggest a specific recruitment of downstream signaling components or the formation of higher oligomeric complexes.

Plant NLR activation by effectors or by mutations affecting nucleotide binding leads to oligomerization and the formation of large NLR complexes, termed resistosomes [27,36,38,43,61]. Regardless of the type of NLR (TNL, CNL or CCR-NLR) oligomerization is most likely required to bring the N-terminal domains into close proximity so that these can facilitate their signaling function or the recruitment of downstream components. This is further supported by many reports demonstrating that the expression of TIR or CC/CCR domains alone is sufficient to induce cell death and cause (auto-) immunity [28,39,45–48].

The mechanism(s) of initiating immunity is however very different between the two major NLR classes, the TNLs and CNLs. Intriguingly, all tested TNLs require the presence of members of the RNL family, a CC-NLR subfamily, for proper cell death and resistance activation, and therefore RNLs are also considered as helper NLRs [25,71–77]. In solanaceous plants (for example potato, tomato and tobacco) another family of helper NLRs was described, the NLR-required for cell death (NRC) proteins [78]. NRCs are typical CNLs and form a functional NLR network with phylogenetically related CNLs, and this NLR network most likely enhances the robustness and complexity of NLR-mediated immunity [79]. CNL function therefore can depend on helper NLRs but many CNLs are thought to function independent of helper NLR assistance — at least to a certain degree [71,73,76] — and these CNLs are also termed as ‘helper-independent’ NLRs or singletons [80].

The cryo-EM structure of activated full-length Arabidopsis CNL ZAR1 suggested that the CC domains of the five ZAR1 proteins in the resistosome complex form a funnel-like structure that may penetrate the plasma membrane to either build a pore or channel or to disrupt membrane integrity [61]. A recent follow-up study impressively demonstrated that the activated ZAR1 resistosome indeed forms a Ca2+ permeable cation channel at the plant plasma membrane [63]. Formation of this channel requires ZAR1 activation, oligomerization, (plasma) membrane localization and the formation of a pore by the CC domain. Specific negatively charged glutamate (E) residues of the CC domain positioned in the channel pore are required for cation (Ca2+) fluxes and the subsequent initiation of cell death and immunity. How ZAR1 activation exactly triggers cell death during an ETI response is still unclear. However, cell biological time-lapse analyses have shown that ZAR1 activation leads to intracellular calcium influx and reactive oxygen species (ROS) production, changes in morphology of cellular compartments/organelles and ultimately to rupture of the cell, most likely due to loss of plasma membrane integrity. All these responses were lost when the conserved glutamate (E11) was mutated, indicating that channel activity and therefore most likely Ca2+ influxes are upstream of these responses [63].

It was previously shown that the Arabidopsis CNL (RPM1) and the ADR1 RNL-subfamily self-associate and localize to and function at the plasma membrane to induce a cell death response [64,65]. Self-association of the ADR1s seems to be enhanced in an activation-dependent manner [64,81]. Another recent study demonstrates that autoactive forms of the Arabidopsis helper NLRs NRG1.1/NRG1A and ADR1, both members of the RNL subfamily, do also form Ca2+ permeable cation channels at the plasma membrane (Figure 1), and that RNL-mediated cell death and Ca2+ fluxes also required negatively charged residues in their CCR domain [27]. The important negatively charged residues (E11 in ZAR1, D11 in ADR1 and E14 in NRG1.1/NRG1A) are not conserved in all CNLs, and thus there might be other mechanisms in place [82]. It is very likely that NRG1.1/NRG1A oligomerizes and forms high molecular weight complexes in an activation dependent manner — similar to ZAR1- as an NRG1.1/NRG1A P-loop loss-of-function mutant, affected in nucleotide binding, did not. Based on these recent findings and the fact that NLR-triggered ETI leads to prolonged calcium influxes, conceivably, all CNL and RNL triggered cell death responses may require the formation of some kind of channel, most likely cation channels, in cellular membranes. A dynamic, activation-dependent membrane localization was recently reported for Arabidopsis RNLs and a solanaceous NRC member, NRC4 [27,64,83], suggesting that probably all cell-death and resistance executing CNLs, including the RNLs and NRCs, function at a cellular membrane. This assumption is supported by several reports showing the inhibition of CNL-mediated cell death responses by general calcium channel blockers, for example lanthanum or gadolinium [11,27,63,65,84]. However, whether cation (specifically Ca2+) channel activity is also required for CNL and RNL-mediated signaling leading to disease resistance needs to be determined. There are some examples where NLR-mediated disease resistance does not require cell death or where cell death can occur even in the absence of resistance — thus, cell death and resistance are not necessarily coupled [12,85,86]. The observation that inhibition of Ca2+ fluxes by lanthanum blocks the Arabidopsis RPM1 (CNL)-triggered cell death response, but not disease resistance, may indicate that Ca2+ fluxes are, at least in case of RPM1 mediated immunity, not required for disease resistance [11,65]. However, the exact mechanism by which lanthanum is suppressing cell death initiated by NLRs is not known. Similarly, how long the lanthanum effect is lasting in planta is also not really studied, but this would be important to know, since determining resistance to infections in plants is normally a matter of days.

TNL-triggered immune responses require the presence of the helper NLRs (in Arabidopsis the RNL members of the ADR1 and NRG1 subfamilies) and the lipase-like protein family members Enhanced Disease Susceptibility 1 (EDS1), Phytoalexin Deficient 4 (PAD4) and Senescence-Associated Gene 101 (SAG101) [25,71,77,81,87–89] (Figure 1). Effector-recognition by TNLs induces self-association and in case of TNL-pairs, for example the Arabidopsis RPS4 and RRS1 proteins, hetero-association that is required for proper immune signaling [48]. The formation of a tetrameric TNL-resistosome by effector-bound RPP1 or ROQ1 enables their TIR domains to build a twofold symmetric dimer of TIR domain dimers, which is essential for the recently discovered nicotinamide adenine dinucleotide (NAD+) hydrolase (NADase) activity of this domains [38,43,90,91] (Figure 1). Plant TIR domain NADase activity depends on the presence of a highly conserved glutamate residue that can also be found in TIR domains of prokaryotes and other eukaryotes [30,34], for example in the mammalian TIR domain containing sterile alpha and TIR motif containing 1 (SARM1) protein, which mediates axonal cell death via its NADase enzymatic activity [92,93]. NAD+ and its cleavage products are important for many essential cellular functions, including immune signaling and the activation of calcium channels [34]. The enzymatic hydrolysis of NAD+ by SARM1 leads to NAD+ depletion, which is hypothesized to be sufficient for SARM1 induced cell death [93]. However, a detectable NAD+ depletion by plant TIRs could not be observed [90]. Plant TIR domains hydrolyze NAD+ into ADP-Ribose (ADPR), Nicotinamide (NAM) and variant cyclic-ADPR (v-cADPR) [90,91], whereas SARM1 activity generates NAM, ADPR and cADPR [92], thus plant and animal catalytic products differ from each other. Intriguingly, heterologous expression of SARM1 TIR domain in plant cells also induces a cell death response, similar to overexpression of plant TIR domains [90,91]. However, the SARM1 TIR domain cell death activity in plants does not depend on EDS1 [90,91] and most likely also not on the RNL helper NLRs, although both components are absolutely required for plant TIR-induced cell death [77]. This indicates that the mechanism of cell death induction mediated by SARM1 and plant TIR domains is different and/or that v-cADPR produced by plant TIR domains is required for signal initiation upstream of EDS1 and RNLs. It is worth to mention that in planta expression of bacterial TIR domains, which also produce a v-cADPR, was also not sufficient to activate a cell death response [94]. Probably because the v-cADPR produced differs or bacterial TIR domains are not able to interact with other proteins required for plant TIR domain signaling. However, the exact mechanism of how plant TIR domain NADase-activity and the thereby produced signaling molecules initiate or activate immune responses is largely unknown.

CNL-mediated ETI and also TNL-mediated ETI, which is essentially activation of RNLs [71,73–76] (Figure 1), lead to measurable increases in cytosolic calcium concentrations, the initiation of ROS production and activation of MAPK as well as calcium-dependent protein kinase (CDPK) signaling [95,96]. These immune outputs are all responses that are shared with PTI [97–99]. One difference, also this is not exclusively true for all cases, is the initiation of the HR cell death at the final stage of ETI. HR cell death is often associated to be specific for NLR activity, although many studies reported that perception of apoplastic effectors or even MAMPs by PRRs does also induce HR-like cell death responses [100–103]. Furthermore, overexpression of or loss-of-function mutations in important plant PRR co-receptors, like the LRR-RK SOBIR1 or members of the SERK family (for example BAK1 or BKK1), induces severe autoimmune phenotypes with obvious cell death symptoms [104–110]. However, recent findings demonstrated that some of these autoimmune phenotypes induced by a disturbed LRR-RK abundancy are indeed linked to NLR activity [111,112]. Similarly, resistance and cell death triggered by perception of extracellular effectors or ligands by PRRs of the LRR-RLP type is also reported to require helper NLR (RNL) presence [79,113–116]. Another recent finding suggests that signaling by TNLs is contributing to full PTI responses and that PTI induces the rapid transcriptional up-regulation of many TNL genes [117]. Therefore, immunity induced by PRR activation may always depend on an actual NLR activation and thus, is feeding into the classical ETI pathway components [116]. The already very fine line separating PTI and ETI is even further being blurred by two recent and important studies demonstrating a mutual potentiation and interdependency of PTI and ETI pathways [9,10], suggesting convergence of immune signaling initiated by the two receptor-based tiers of plant immunity. The transcriptional activation of important PTI pathway genes and the induced phosphorylation of many essential immune components upon NLR activation is further supporting the idea of a shared or common regulatory or signaling node for cell surface and intracellular immune receptors [97,118,119]. It will be interesting to see, whether such a ‘PTI-pathway’ activation can also be observed in NLR autoimmune mutants or during hybrid incompatibility [68,69,120,121].

The mechanisms leading to the loss of cell integrity during ETI are not well studied. There are a few reports showing that NLR activation or the expression of viral, bacterial and fungal effectors can induce severe morphological cellular phenotypes, like the fusion of the vacuole with the plasma membrane [122], alterations of vacuolar structures (El Kasmi lab unpublished; [63]), the expansion of the nucleus [123], chloroplast clustering around the nucleus [124], cytoplasmic shrinkage, mitochondrial swelling and many more [125]. It is likely that these cellular changes and the loss of plasma membrane integrity is leading to a release or (calcium-) induced production of DAMPs [126,127], molecular signals that are known to be potent immune activators [128]. Production of systemic signals is indeed known to take place during both PTI and ETI to induce priming and a systemic acquired resistance [13,129–131]. However, whether these signals are released prior to cell death and if yes, how they are released from the infected cells is still unclear. One attractive possibility would be that activated CNLs and RNLs could, in addition to being calcium permeable channels or pores, allow the release of such DAMP molecules even before the actual rupture or breakdown of the infected cell. This would be similar to the function of pore forming and cell death inducing proteins, for example gasdermin or MLKL in the animal kingdom, or Het-S/HeLo like proteins in fungal organisms [132–134]. Plants do also posses MLKL like proteins capable of inducing cell death and involved in TNL-mediated resistance signaling [26]. It will be interesting to see whether and how MLKL proteins and NLRs cooperate during plant immunity and if there is any kind of specificity.

NLR activation leads to calcium influxes and probably initiation of other ion fluxes, and this in turn can activate other immune regulatory proteins, including cyclic nucleotide-gated cation channels (CNGCs) [135] (Figure 1), CPKs and transcriptional regulators of the calmodulin-binding protein family that are crucial for expression of salicylic acid biosynthesis genes and genes important for production and release of antimicrobial molecules [95,136].

  • Importance of the field: Intracellular NLR receptors are important components of a plants defense system against host-adapted and non-adapted pathogens. NLR-triggered immunity leads to strong and robust defense responses and thus understanding their mechanism(s) of action is of high importance for agricultural research to generate better and durable resistant crop plants.

  • Current thinking: Recognition of pathogen-derived effectors activates NLR proteins and induces their self-association, which in turn leads to oligomerization of their N-termini and the production of signaling molecules or the formation of a cation channel at cellular membranes, depending on the architecture of the N-terminal domain. NLR activity generally, but not always, leads to a strong immune response in the whole organism and an induced cell death of the infected cell.

  • Future directions: Identifying NLR binding partners that might contribute to their specific functions and regulate their activities, will be crucial in enhancing our understanding of how the different types of NLRs mediate immunity against a wide variety of pathogens. Determining the exact site of NLR action and the mechanism of cell death induction is also one of the fields next important and interesting endeavors.

The author declares that there are no competing interests associated with this manuscript.

Research in the El Kasmi lab is funded by the University of Tübingen, the Deutsche Forschungsgemeinschaft (grant no. DFG-CRC1101-D09 to F.E.K.) and the Reinhard Frank Stiftung (Project ‘helperless plant’ to F.E.K.).

F.E.K. wrote the manuscript.

I would like to thank the anonymous reviewers, my lab members and the Dangl and Nishimura NLR teams for helpful discussions. A huge ‘thank you’ also to S. Saile and S. Sunil for critically reading of the manuscript and S. Saile for helpful comments on the figure.

ADP

adenosine diphosphate

ATP

adenosine triphosphate

CNGCs

cyclic nucleotide-gated cation channels

DAMPs

damage-associated molecular patterns

EDS1

enhanced disease susceptibility 1

ETI

effector-triggered immunity

HR

hypersensitive response

ID

integrated domain

LRR

leucine-rich repeat

MAPK

mitogen-activated protein kinase

MLKL

mixed-linage kinase like

NAM

Nicotinamide

NB

nucleotide-binding

NLRs

nucleotide-binding LRR receptors

NRC

NLR-required for cell death

PAD4

Phytoalexin Deficient 4

PRRs

pattern-recognition receptors

PTI

pattern-triggered immunity

RIN4

RPM1-interacting protein 4

RPP1

Recognition of Peronospora parasitica 1A

SARM1

sterile alpha and TIR motif containing 1

TIR

Toll/Interleukin-1 receptor/Resistance protein

1
Jones
,
J.D.
and
Dangl
,
J.L.
(
2006
)
The plant immune system
.
Nature
444
,
323
329
2
Couto
,
D.
and
Zipfel
,
C.
(
2016
)
Regulation of pattern recognition receptor signalling in plants
.
Nat. Rev. Immunol.
16
,
537
552
3
Wan
,
W.L.
,
Frohlich
,
K.
,
Pruitt
,
R.N.
,
Nurnberger
,
T.
and
Zhang
,
L.
(
2019
)
Plant cell surface immune receptor complex signaling
.
Curr. Opin. Plant Biol.
50
,
18
28
4
Liang
,
X.
and
Zhou
,
J.M.
(
2018
)
Receptor-like cytoplasmic kinases: central players in plant receptor kinase-mediated signaling
.
Annu. Rev. Plant Biol.
69
,
267
299
5
Kasmi F
,
E.
,
Horvath
,
D.
and
Lahaye
,
T.
(
2018
)
Microbial effectors and the role of water and sugar in the infection battle ground
.
Curr. Opin. Plant Biol.
44
,
98
107
6
Varden
,
F.A.
,
De la Concepcion
,
J.C.
,
Maidment
,
J.H.
and
Banfield
,
M.J.
(
2017
)
Taking the stage: effectors in the spotlight
.
Curr. Opin. Plant Biol.
38
,
25
33
7
Jones
,
J.D.
,
Vance
,
R.E.
and
Dangl
,
J.L.
(
2016
)
Intracellular innate immune surveillance devices in plants and animals
.
Science
354
,
aaf6395
8
Duxbury
,
Z.
,
Wu
,
C.H.
and
Ding
,
P.
(
2021
)
A comparative overview of the intracellular guardians of plants and animals: NLRs in innate immunity and beyond
.
Annu. Rev. Plant Biol.
72
,
155
184
9
Ngou
,
B.P.M.
,
Ahn
,
H.K.
,
Ding
,
P.
and
Jones
,
J.D.G.
(
2021
)
Mutual potentiation of plant immunity by cell-surface and intracellular receptors
.
Nature
592
,
110
115
10
Yuan
,
M.
,
Jiang
,
Z.
,
Bi
,
G.
,
Nomura
,
K.
,
Liu
,
M.
,
Wang
,
Y.
et al (
2021
)
Pattern-recognition receptors are required for NLR-mediated plant immunity
.
Nature
592
,
105
109
11
Grant
,
M.
,
Brown
,
I.
,
Adams
,
S.
,
Knight
,
M.
,
Ainslie
,
A.
and
Mansfield
,
J.
(
2000
)
The RPM1 plant disease resistance gene facilitates a rapid and sustained increase in cytosolic calcium that is necessary for the oxidative burst and hypersensitive cell death
.
Plant J.
23
,
441
450
12
Balint-Kurti
,
P.
(
2019
)
The plant hypersensitive response: concepts, control and consequences
.
Mol. Plant Pathol.
20
,
1163
1178
13
Vlot
,
A.C.
,
Sales
,
J.H.
,
Lenk
,
M.
,
Bauer
,
K.
,
Brambilla
,
A.
,
Sommer
,
A.
et al (
2021
)
Systemic propagation of immunity in plants
.
New Phytol.
229
,
1234
1250
14
Macho
,
A.P.
and
Lozano-Duran
,
R.
(
2019
)
Molecular dialogues between viruses and receptor-like kinases in plants
.
Mol. Plant Pathol.
20
,
1191
1195
15
Mei
,
Y.
,
Wang
,
Y.
,
Hu
,
T.
,
He
,
Z.
and
Zhou
,
X.
(
2021
)
The C4 protein encoded by Tomato leaf curl Yunnan virus interferes with mitogen-activated protein kinase cascade-related defense responses through inhibiting the dissociation of the ERECTA/BKI1 complex
.
New Phytol.
231
,
747
762
16
Zhang
,
H.
,
Chen
,
C.
,
Li
,
L.
,
Tan
,
X.
,
Wei
,
Z.
,
Li
,
Y.
et al (
2021
)
A rice LRR receptor-like protein associates with its adaptor kinase OsSOBIR1 to mediate plant immunity against viral infection
.
Plant Biotechnol. J.
17
Urbach
,
J.M.
and
Ausubel
,
F.M.
(
2017
)
The NBS-LRR architectures of plant R-proteins and metazoan NLRs evolved in independent events
.
Proc. Natl Acad. Sci. U.S.A.
114
,
1063
1068
18
Whitham
,
S.
,
Dinesh-Kumar
,
S.P.
,
Choi
,
D.
,
Hehl
,
R.
,
Corr
,
C.
and
Baker
,
B.
(
1994
)
The product of the tobacco mosaic virus resistance gene N: similarity to toll and the interleukin-1 receptor
.
Cell
78
,
1101
1115
19
Bent
,
A.F.
,
Kunkel
,
B.N.
,
Dahlbeck
,
D.
,
Brown
,
K.L.
,
Schmidt
,
R.
,
Giraudat
,
J.
et al (
1994
)
RPS2 of Arabidopsis thaliana: a leucine-rich repeat class of plant disease resistance genes
.
Science
265
,
1856
1860
20
Mindrinos
,
M.
,
Katagiri
,
F.
,
Yu
,
G.L.
and
Ausubel
,
F.M.
(
1994
)
The A. thaliana disease resistance gene RPS2 encodes a protein containing a nucleotide-binding site and leucine-rich repeats
.
Cell
78
,
1089
1099
21
Kourelis
,
J.
and
van der Hoorn
,
R.A.L.
(
2018
)
Defended to the nines: 25 years of resistance gene cloning identifies nine mechanisms for R protein function
.
Plant Cell
30
,
285
299
22
Sukarta
,
O.C.A.
,
Slootweg
,
E.J.
and
Goverse
,
A.
(
2016
)
Structure-informed insights for NLR functioning in plant immunity
.
Semin. Cell Dev. Biol.
56
,
134
149
23
Wang
,
J.
,
Wang
,
J.
,
Hu
,
M.
,
Wu
,
S.
,
Qi
,
J.
,
Wang
,
G.
et al (
2019
)
Ligand-triggered allosteric ADP release primes a plant NLR complex
.
Science
364
,
eaav5868
24
Bentham
,
A.R.
,
Zdrzalek
,
R.
,
De la Concepcion
,
J.C.
and
Banfield
,
M.J.
(
2018
)
Uncoiling CNLs: structure/function approaches to understanding CC domain function in plant NLRs
.
Plant Cell Physiol.
59
,
2398
2408
25
Jubic
,
L.M.
,
Saile
,
S.
,
Furzer
,
O.J.
,
El Kasmi
,
F.
and
Dangl
,
J.L.
(
2019
)
Help wanted: helper NLRs and plant immune responses
.
Curr. Opin. Plant Biol.
50
,
82
94
26
Mahdi
,
L.K.
,
Huang
,
M.
,
Zhang
,
X.
,
Nakano
,
R.T.
,
Kopp
,
L.B.
,
Saur
,
I.M.L.
et al (
2020
)
Discovery of a family of mixed lineage kinase domain-like proteins in plants and their role in innate immune signaling
.
Cell Host Microbe
28
,
813
24 e6
27
Jacob
,
P.
,
Kim
,
N.H.
,
Wu
,
F.
,
El-Kasmi
,
F.
,
Chi
,
Y.
,
Walton
,
W.G.
et al (
2021
)
Plant “helper” immune receptors are Ca(2+)-permeable nonselective cation channels
.
Science
373
,
420
425
28
Casey
,
L.W.
,
Lavrencic
,
P.
,
Bentham
,
A.R.
,
Cesari
,
S.
,
Ericsson
,
D.J.
,
Croll
,
T.
et al (
2016
)
The CC domain structure from the wheat stem rust resistance protein Sr33 challenges paradigms for dimerization in plant NLR proteins
.
Proc. Natl Acad. Sci. U.S.A.
113
,
12856
12861
29
Maekawa
,
T.
,
Cheng
,
W.
,
Spiridon
,
L.N.
,
Toller
,
A.
,
Lukasik
,
E.
,
Saijo
,
Y.
et al (
2011
)
Coiled-coil domain-dependent homodimerization of intracellular barley immune receptors defines a minimal functional module for triggering cell death
.
Cell Host Microbe
9
,
187
199
30
Essuman
,
K.
,
Summers
,
D.W.
,
Sasaki
,
Y.
,
Mao
,
X.
,
Yim
,
A.K.Y.
,
DiAntonio
,
A.
et al (
2018
)
TIR domain proteins are an ancient family of NAD(+)-consuming enzymes
.
Curr. Biol.
28
,
421
30 e4
31
Ve
,
T.
,
Williams
,
S.J.
and
Kobe
,
B.
(
2015
)
Structure and function of toll/interleukin-1 receptor/resistance protein (TIR) domains
.
Apoptosis
20
,
250
261
32
Doron
,
S.
,
Melamed
,
S.
,
Ofir
,
G.
,
Leavitt
,
A.
,
Lopatina
,
A.
,
Keren
,
M.
et al (
2018
)
Systematic discovery of antiphage defense systems in the microbial pangenome
.
Science
359
,
eaar4120
33
Ofir
,
G.
,
Herbst
,
E.
,
Baroz
,
M.
,
Cohen
,
D.
,
Millman
,
A.
,
Doron
,
S.
et al (
2021
)
Antiviral activity of bacterial TIR domains via signaling molecules that trigger cell death
.
bioRxiv
34
Bayless
,
A.M.
and
Nishimura
,
M.T.
(
2020
)
Enzymatic functions for toll/interleukin-1 receptor domain proteins in the plant immune system
.
Front. Genet.
11
,
539
35
Enkhbayar
,
P.
,
Kamiya
,
M.
,
Osaki
,
M.
,
Matsumoto
,
T.
and
Matsushima
,
N.
(
2004
)
Structural principles of leucine-rich repeat (LRR) proteins
.
Proteins
54
,
394
403
36
Burdett
,
H.
,
Bentham
,
A.R.
,
Williams
,
S.J.
,
Dodds
,
P.N.
,
Anderson
,
P.A.
,
Banfield
,
M.J.
et al (
2019
)
The plant “resistosome”: structural insights into immune signaling
.
Cell Host Microbe
26
,
193
201
37
Takken
,
F.L.
,
Albrecht
,
M.
and
Tameling
,
W.I.
(
2006
)
Resistance proteins: molecular switches of plant defence
.
Curr. Opin. Plant Biol.
9
,
383
390
38
Ma
,
S.
,
Lapin
,
D.
,
Liu
,
L.
,
Sun
,
Y.
,
Song
,
W.
,
Zhang
,
X.
et al (
2020
)
Direct pathogen-induced assembly of an NLR immune receptor complex to form a holoenzyme
.
Science
370
,
eabe3069
39
Wang
,
G.F.
,
Ji
,
J.
,
El-Kasmi
,
F.
,
Dangl
,
J.L.
,
Johal
,
G.
and
Balint-Kurti
,
P.J.
(
2015
)
Molecular and functional analyses of a maize autoactive NB-LRR protein identify precise structural requirements for activity
.
PLoS Pathog.
11
,
e1004674
40
Qi
,
D.
,
DeYoung
,
B.J.
and
Innes
,
R.W.
(
2012
)
Structure-function analysis of the coiled-coil and leucine-rich repeat domains of the RPS5 disease resistance protein
.
Plant Physiol.
158
,
1819
1832
41
Michael Weaver
,
L.
,
Swiderski
,
M.R.
,
Li
,
Y.
and
Jones
,
J.D.
(
2006
)
The Arabidopsis thaliana TIR-NB-LRR R-protein, RPP1A; protein localization and constitutive activation of defence by truncated alleles in tobacco and Arabidopsis
.
Plant J.
47
,
829
840
42
Hu
,
Z.
,
Yan
,
C.
,
Liu
,
P.
,
Huang
,
Z.
,
Ma
,
R.
,
Zhang
,
C.
et al (
2013
)
Crystal structure of NLRC4 reveals its autoinhibition mechanism
.
Science
341
,
172
175
43
Martin
,
R.
,
Qi
,
T.
,
Zhang
,
H.
,
Liu
,
F.
,
King
,
M.
,
Toth
,
C.
et al (
2020
)
Structure of the activated ROQ1 resistosome directly recognizing the pathogen effector xopQ
.
Science
370
,
eabd9993
44
Prigozhin
,
D.M.
and
Krasileva
,
K.V.
(
2021
)
Analysis of intraspecies diversity reveals a subset of highly variable plant immune receptors and predicts their binding sites
.
Plant Cell
33
,
998
1015
45
Nishimura
,
M.T.
,
Anderson
,
R.G.
,
Cherkis
,
K.A.
,
Law
,
T.F.
,
Liu
,
Q.L.
,
Machius
,
M.
et al (
2017
)
TIR-only protein RBA1 recognizes a pathogen effector to regulate cell death in Arabidopsis
.
Proc. Natl Acad. Sci. U.S.A.
114
,
E2053
E2E62
46
Jacob
,
F.
,
Kracher
,
B.
,
Mine
,
A.
,
Seyfferth
,
C.
,
Blanvillain-Baufume
,
S.
,
Parker
,
J.E.
et al (
2018
)
A dominant-interfering camta3 mutation compromises primary transcriptional outputs mediated by both cell surface and intracellular immune receptors in Arabidopsis thaliana
.
New Phytol.
217
,
1667
1680
47
Bernoux
,
M.
,
Ve
,
T.
,
Williams
,
S.
,
Warren
,
C.
,
Hatters
,
D.
,
Valkov
,
E.
et al (
2011
)
Structural and functional analysis of a plant resistance protein TIR domain reveals interfaces for self-association, signaling, and autoregulation
.
Cell Host Microbe
9
,
200
211
48
Zhang
,
X.
,
Bernoux
,
M.
,
Bentham
,
A.R.
,
Newman
,
T.E.
,
Ve
,
T.
,
Casey
,
L.W.
et al (
2017
)
Multiple functional self-association interfaces in plant TIR domains
.
Proc. Natl Acad. Sci. U.S.A.
114
,
E2046
E2E52
49
Wang
,
J.
,
Chern
,
M.
and
Chen
,
X.
(
2020
)
Structural dynamics of a plant NLR resistosome: transition from autoinhibition to activation
.
Sci. China Life Sci.
63
,
617
619
50
Dangl
,
J.L.
and
Jones
,
J.D.
(
2001
)
Plant pathogens and integrated defence responses to infection
.
Nature
411
,
826
833
51
Le Roux
,
C.
,
Huet
,
G.
,
Jauneau
,
A.
,
Camborde
,
L.
,
Tremousaygue
,
D.
,
Kraut
,
A.
et al (
2015
)
A receptor pair with an integrated decoy converts pathogen disabling of transcription factors to immunity
.
Cell
161
,
1074
1088
52
Sarris
,
P.F.
,
Duxbury
,
Z.
,
Huh
,
S.U.
,
Ma
,
Y.
,
Segonzac
,
C.
,
Sklenar
,
J.
et al (
2015
)
A plant immune receptor detects pathogen effectors that target WRKY transcription factors
.
Cell
161
,
1089
1100
53
Maqbool
,
A.
,
Saitoh
,
H.
,
Franceschetti
,
M.
,
Stevenson
,
C.E.
,
Uemura
,
A.
,
Kanzaki
,
H.
et al (
2015
)
Structural basis of pathogen recognition by an integrated HMA domain in a plant NLR immune receptor
.
eLife
4
,
e08709
54
De la Concepcion
,
J.C.
,
Franceschetti
,
M.
,
MacLean
,
D.
,
Terauchi
,
R.
,
Kamoun
,
S.
and
Banfield
,
M.J.
(
2019
)
Protein engineering expands the effector recognition profile of a rice NLR immune receptor
.
eLife
8
,
e47713
55
Mackey
,
D.
,
Belkhadir
,
Y.
,
Alonso
,
J.M.
,
Ecker
,
J.R.
and
Dangl
,
J.L.
(
2003
)
Arabidopsis RIN4 is a target of the type III virulence effector AvrRpt2 and modulates RPS2-mediated resistance
.
Cell
112
,
379
389
56
Mackey
,
D.
,
Holt
, III,
B.F.
,
Wiig
,
A.
and
Dangl
,
J.L.
(
2002
)
RIN4 interacts with pseudomonas syringae type III effector molecules and is required for RPM1-mediated resistance in Arabidopsis
.
Cell
108
,
743
754
57
Axtell
,
M.J.
and
Staskawicz
,
B.J.
(
2003
)
Initiation of RPS2-specified disease resistance in Arabidopsis is coupled to the AvrRpt2-directed elimination of RIN4
.
Cell
112
,
369
377
58
Chung
,
E.H.
,
da Cunha
,
L.
,
Wu
,
A.J.
,
Gao
,
Z.
,
Cherkis
,
K.
,
Afzal
,
A.J.
et al (
2011
)
Specific threonine phosphorylation of a host target by two unrelated type III effectors activates a host innate immune receptor in plants
.
Cell Host Microbe
9
,
125
136
59
Liu
,
J.
,
Elmore
,
J.M.
,
Lin
,
Z.J.
and
Coaker
,
G.
(
2011
)
A receptor-like cytoplasmic kinase phosphorylates the host target RIN4, leading to the activation of a plant innate immune receptor
.
Cell Host Microbe
9
,
137
146
60
Redditt
,
T.J.
,
Chung
,
E.H.
,
Karimi
,
H.Z.
,
Rodibaugh
,
N.
,
Zhang
,
Y.
,
Trinidad
,
J.C.
et al (
2019
)
Avrrpm1 functions as an ADP-ribosyl transferase to modify NOI domain-containing proteins, including Arabidopsis and soybean RPM1-interacting Protein4
.
Plant Cell
31
,
2664
2681
61
Wang
,
J.
,
Hu
,
M.
,
Wang
,
J.
,
Qi
,
J.
,
Han
,
Z.
,
Wang
,
G.
et al (
2019
)
Reconstitution and structure of a plant NLR resistosome conferring immunity
.
Science
364
,
eaav5870
62
Martel
,
A.
,
Laflamme
,
B.
,
Seto
,
D.
,
Bastedo
,
D.P.
,
Dillon
,
M.M.
,
Almeida
,
R.N.D.
et al (
2020
)
Immunodiversity of the Arabidopsis ZAR1 NLR is conveyed by receptor-like cytoplasmic kinase sensors
.
Front Plant Sci.
11
,
1290
63
Bi
,
G.
,
Su
,
M.
,
Li
,
N.
,
Liang
,
Y.
,
Dang
,
S.
,
Xu
,
J.
et al (
2021
)
The ZAR1 resistosome is a calcium-permeable channel triggering plant immune signaling
.
Cell
184
,
3528
3541.e12
64
Saile
,
S.C.
,
Ackermann
,
F.M.
,
Sunil
,
S.
,
Bayless
,
A.
,
Stöbbe
,
E.
,
Bonardi
,
V.
et al (
2021
)
Coiled-coil and RPW8-type immune receptors function at the plasma membrane in a phospholipid dependent manner
.
bioRxiv
65
Kasmi F
,
E.
,
Chung
,
E.H.
,
Anderson
,
R.G.
,
Li
,
J.
,
Wan
,
L.
,
Eitas
,
T.K.
et al (
2017
)
Signaling from the plasma-membrane localized plant immune receptor RPM1 requires self-association of the full-length protein
.
Proc. Natl Acad. Sci. U.S.A.
114
,
E7385
E7E94
66
Williams
,
S.J.
,
Sohn
,
K.H.
,
Wan
,
L.
,
Bernoux
,
M.
,
Sarris
,
P.F.
,
Segonzac
,
C.
et al (
2014
)
Structural basis for assembly and function of a heterodimeric plant immune receptor
.
Science
344
,
299
303
67
Mestre
,
P.
and
Baulcombe
,
D.C.
(
2006
)
Elicitor-mediated oligomerization of the tobacco N disease resistance protein
.
Plant Cell
18
,
491
501
68
Tran
,
D.T.N.
,
Chung
,
E.H.
,
Habring-Muller
,
A.
,
Demar
,
M.
,
Schwab
,
R.
,
Dangl
,
J.L.
et al (
2017
)
Activation of a plant NLR complex through heteromeric association with an autoimmune risk variant of another NLR
.
Curr. Biol.
27
,
1148
1160
69
Li
,
L.
,
Habring
,
A.
,
Wang
,
K.
and
Weigel
,
D.
(
2020
)
Atypical resistance protein RPW8/HR triggers oligomerization of the NLR immune receptor RPP7 and autoimmunity
.
Cell Host Microbe
27
,
405
417.e6
70
Ade
,
J.
,
DeYoung
,
B.J.
,
Golstein
,
C.
and
Innes
,
R.W.
(
2007
)
Indirect activation of a plant nucleotide binding site-leucine-rich repeat protein by a bacterial protease
.
Proc. Natl Acad. Sci. U.S.A.
104
,
2531
2536
71
Saile
,
S.C.
,
Jacob
,
P.
,
Castel
,
B.
,
Jubic
,
L.M.
,
Salas-Gonzalez
,
I.
,
Backer
,
M.
et al (
2020
)
Two unequally redundant “helper” immune receptor families mediate Arabidopsis thaliana intracellular “sensor” immune receptor functions
.
PLoS Biol.
18
,
e3000783
72
Peart
,
J.R.
,
Mestre
,
P.
,
Lu
,
R.
,
Malcuit
,
I.
and
Baulcombe
,
D.C.
(
2005
)
NRG1, a CC-NB-LRR protein, together with N, a TIR-NB-LRR protein, mediates resistance against tobacco mosaic virus
.
Curr Biol.
15
,
968
973
73
Bonardi
,
V.
,
Tang
,
S.
,
Stallmann
,
A.
,
Roberts
,
M.
,
Cherkis
,
K.
and
Dangl
,
J.L.
(
2011
)
Expanded functions for a family of plant intracellular immune receptors beyond specific recognition of pathogen effectors
.
Proc. Natl Acad. Sci. U.S.A.
108
,
16463
16468
74
Dong
,
O.X.
,
Tong
,
M.
,
Bonardi
,
V.
,
El Kasmi
,
F.
,
Woloshen
,
V.
,
Wunsch
,
L.K.
et al (
2016
)
TNL-mediated immunity in Arabidopsis requires complex regulation of the redundant ADR1 gene family
.
New Phytol.
210
,
960
973
75
Castel
,
B.
,
Ngou
,
P.M.
,
Cevik
,
V.
,
Redkar
,
A.
,
Kim
,
D.S.
,
Yang
,
Y.
et al (
2019
)
Diverse NLR immune receptors activate defence via the RPW8-NLR NRG1
.
New Phytol.
222
,
966
980
76
Wu
,
Z.
,
Li
,
M.
,
Dong
,
O.X.
,
Xia
,
S.
,
Liang
,
W.
,
Bao
,
Y.
et al (
2019
)
Differential regulation of TNL-mediated immune signaling by redundant helper CNLs
.
New Phytol.
222
,
938
953
77
Lapin
,
D.
,
Kovacova
,
V.
,
Sun
,
X.
,
Dongus
,
J.A.
,
Bhandari
,
D.
,
von Born
,
P.
et al (
2019
)
A coevolved EDS1-SAG101-NRG1 module mediates cell death signaling by TIR-domain immune receptors
.
Plant Cell
31
,
2430
2455
78
Wu
,
C.H.
,
Abd-El-Haliem
,
A.
,
Bozkurt
,
T.O.
,
Belhaj
,
K.
,
Terauchi
,
R.
,
Vossen
,
J.H.
et al (
2017
)
NLR network mediates immunity to diverse plant pathogens
.
Proc. Natl Acad. Sci. U.S.A.
114
,
8113
8118
79
Wu
,
C.H.
,
Derevnina
,
L.
and
Kamoun
,
S.
(
2018
)
Receptor networks underpin plant immunity
.
Science
360
,
1300
1301
80
Adachi
,
H.
,
Derevnina
,
L.
and
Kamoun
,
S.
(
2019
)
NLR singletons, pairs, and networks: evolution, assembly, and regulation of the intracellular immunoreceptor circuitry of plants
.
Curr. Opin. Plant Biol.
50
,
121
131
81
Wu
,
Z.
,
Tian
,
L.
,
Liu
,
X.
,
Zhang
,
Y.
and
Li
,
X.
(
2021
)
TIR signaling promotes the interactions between EDS1/PAD4 and ADR1-L1 and oligomerization of ADR1-L1
.
bioRxiv
82
Adachi
,
H.
,
Contreras
,
M.P.
,
Harant
,
A.
,
Wu
,
C.H.
,
Derevnina
,
L.
,
Sakai
,
T.
et al (
2019
)
An N-terminal motif in NLR immune receptors is functionally conserved across distantly related plant species
.
ELife
8
,
e49956
83
Duggan
,
C.
,
Moratto
,
E.
,
Savage
,
Z.
,
Hamilton
,
E.
,
Adachi
,
H.
,
Wu
,
C.-H.
et al (
2021
)
Dynamic accumulation of a helper NLR at the plant-pathogen interface underpins pathogen recognition
.
PNAS
118
, e2104997118
84
Lee
,
H.Y.
,
Mang
,
H.
,
Choi
,
E.
,
Seo
,
Y.E.
,
Kim
,
M.S.
,
Oh
,
S.
et al (
2021
)
Genome-wide functional analysis of hot pepper immune receptors reveals an autonomous NLR clade in seed plants
.
New Phytol.
229
,
532
547
85
Coll
,
N.S.
,
Vercammen
,
D.
,
Smidler
,
A.
,
Clover
,
C.
,
Van Breusegem
,
F.
,
Dangl
,
J.L.
et al (
2010
)
Arabidopsis type I metacaspases control cell death
.
Science
330
,
1393
1397
86
Century
,
K.S.
,
Holub
,
E.B.
and
Staskawicz
,
B.J.
(
1995
)
NDR1, a locus of Arabidopsis thaliana that is required for disease resistance to both a bacterial and a fungal pathogen
.
Proc. Natl Acad. Sci. U.S.A.
92
,
6597
6601
87
Sun
,
X.
,
Lapin
,
D.
,
Feehan
,
J.M.
,
Stolze
,
S.C.
,
Kramer
,
K.
,
Dongus
,
J.A.
et al (
2021
)
Pathogen effector recognition-dependent association of NRG1 with EDS1 and SAG101 in TNL receptor immunity
.
Nat. Commun.
12
,
3335
88
Lapin
,
D.
,
Bhandari
,
D.D.
and
Parker
,
J.E.
(
2020
)
Origins and immunity networking functions of EDS1 family proteins
.
Annu. Rev. Phytopathol.
58
,
253
276
89
Feehan
,
J.M.
,
Castel
,
B.
,
Bentham
,
A.R.
and
Jones
,
J.D.
(
2020
)
Plant NLRs get by with a little help from their friends
.
Curr. Opin. Plant Biol.
56
,
99
108
90
Wan
,
L.
,
Essuman
,
K.
,
Anderson
,
R.G.
,
Sasaki
,
Y.
,
Monteiro
,
F.
,
Chung
,
E.H.
et al (
2019
)
TIR domains of plant immune receptors are NAD(+)-cleaving enzymes that promote cell death
.
Science
365
,
799
803
91
Horsefield
,
S.
,
Burdett
,
H.
,
Zhang
,
X.
,
Manik
,
M.K.
,
Shi
,
Y.
,
Chen
,
J.
et al (
2019
)
NAD(+) cleavage activity by animal and plant TIR domains in cell death pathways
.
Science
365
,
793
799
92
Essuman
,
K.
,
Summers
,
D.W.
,
Sasaki
,
Y.
,
Mao
,
X.
,
DiAntonio
,
A.
and
Milbrandt
,
J.
(
2017
)
The SARM1 toll/interleukin-1 receptor domain possesses intrinsic NAD(+) cleavage activity that promotes pathological axonal degeneration
.
Neuron
93
,
1334
1343 e5
93
Gerdts
,
J.
,
Brace
,
E.J.
,
Sasaki
,
Y.
,
DiAntonio
,
A.
and
Milbrandt
,
J.
(
2015
)
SARM1 activation triggers axon degeneration locally via NAD(+) destruction
.
Science
348
,
453
457
94
Duxbury
,
Z.
,
Wang
,
S.
,
MacKenzie
,
C.I.
,
Tenthorey
,
J.L.
,
Zhang
,
X.
,
Huh
,
S.U.
et al (
2020
)
Induced proximity of a TIR signaling domain on a plant-mammalian NLR chimera activates defense in plants
.
Proc. Natl Acad. Sci. U.S.A.
117
,
18832
18839
95
Gao
,
X.
,
Chen
,
X.
,
Lin
,
W.
,
Chen
,
S.
,
Lu
,
D.
,
Niu
,
Y.
et al (
2013
)
Bifurcation of Arabidopsis NLR immune signaling via Ca(2)(+)-dependent protein kinases
.
PLoS Pathog.
9
,
e1003127
96
Qi
,
J.
,
Wang
,
J.
,
Gong
,
Z.
and
Zhou
,
J.M.
(
2017
)
Apoplastic ROS signaling in plant immunity
.
Curr. Opin. Plant Biol.
38
,
92
100
97
Lu
,
Y.
and
Tsuda
,
K.
(
2021
)
Intimate association of PRR- and NLR-mediated signaling in plant immunity
.
Mol. Plant Microbe Interact.
34
,
3
14
98
Thomma
,
B.P.
,
Nurnberger
,
T.
and
Joosten
,
M.H.
(
2011
)
Of PAMPs and effectors: the blurred PTI-ETI dichotomy
.
Plant Cell
23
,
4
15
99
van der Burgh
,
A.M.
and
Joosten
,
M.
(
2019
)
Plant immunity: thinking outside and inside the box
.
Trends Plant Sci.
24
,
587
601
100
Albert
,
I.
,
Bohm
,
H.
,
Albert
,
M.
,
Feiler
,
C.E.
,
Imkampe
,
J.
,
Wallmeroth
,
N.
et al (
2015
)
An RLP23-SOBIR1-BAK1 complex mediates NLP-triggered immunity
.
Nat. Plants
1
,
15140
101
van der Burgh
,
A.M.
,
Postma
,
J.
,
Robatzek
,
S.
and
Joosten
,
M.
(
2019
)
Kinase activity of SOBIR1 and BAK1 is required for immune signalling
.
Mol. Plant Pathol.
20
,
410
422
102
Ron
,
M.
and
Avni
,
A.
(
2004
)
The receptor for the fungal elicitor ethylene-inducing xylanase is a member of a resistance-like gene family in tomato
.
Plant Cell
16
,
1604
1615
103
Naito
,
K.
,
Taguchi
,
F.
,
Suzuki
,
T.
,
Inagaki
,
Y.
,
Toyoda
,
K.
,
Shiraishi
,
T.
et al (
2008
)
Amino acid sequence of bacterial microbe-associated molecular pattern flg22 is required for virulence
.
Mol. Plant Microbe Interact.
21
,
1165
1174
104
Kemmerling
,
B.
,
Schwedt
,
A.
,
Rodriguez
,
P.
,
Mazzotta
,
S.
,
Frank
,
M.
,
Qamar
,
S.A.
et al (
2007
)
The BRI1-associated kinase 1, BAK1, has a brassinolide-independent role in plant cell-death control
.
Curr. Biol.
17
,
1116
1122
105
He
,
K.
,
Gou
,
X.
,
Yuan
,
T.
,
Lin
,
H.
,
Asami
,
T.
,
Yoshida
,
S.
et al (
2007
)
BAK1 and BKK1 regulate brassinosteroid-dependent growth and brassinosteroid-independent cell-death pathways
.
Curr. Biol.
17
,
1109
1115
106
Imkampe
,
J.
,
Halter
,
T.
,
Huang
,
S.
,
Schulze
,
S.
,
Mazzotta
,
S.
,
Schmidt
,
N.
et al (
2017
)
The Arabidopsis leucine-rich repeat receptor kinase BIR3 negatively regulates BAK1 receptor complex formation and stabilizes BAK1
.
Plant Cell
29
,
2285
2303
107
Halter
,
T.
,
Imkampe
,
J.
,
Mazzotta
,
S.
,
Wierzba
,
M.
,
Postel
,
S.
,
Bucherl
,
C.
et al (
2014
)
The leucine-rich repeat receptor kinase BIR2 is a negative regulator of BAK1 in plant immunity
.
Curr. Biol.
24
,
134
143
108
Gao
,
M.
,
Wang
,
X.
,
Wang
,
D.
,
Xu
,
F.
,
Ding
,
X.
,
Zhang
,
Z.
et al (
2009
)
Regulation of cell death and innate immunity by two receptor-like kinases in Arabidopsis
.
Cell Host Microbe
6
,
34
44
109
Gao
,
X.
,
Ruan
,
X.
,
Sun
,
Y.
,
Wang
,
X.
and
Feng
,
B.
(
2018
)
BAKing up to survive a battle: functional dynamics of BAK1 in plant programmed cell death
.
Front. Plant Sci.
9
,
1913
110
Huang
,
W.R.H.
,
Schol
,
C.
,
Villanueva
,
S.L.
,
Heidstra
,
R.
and
Joosten
,
M.
(
2021
)
Knocking out SOBIR1 in Nicotiana benthamiana abolishes functionality of transgenic receptor-like protein Cf-4
.
Plant Physiol.
185
,
290
294
111
Wu
,
Y.
,
Gao
,
Y.
,
Zhan
,
Y.
,
Kui
,
H.
,
Liu
,
H.
,
Yan
,
L.
et al (
2020
)
Loss of the common immune coreceptor BAK1 leads to NLR-dependent cell death
.
Proc. Natl Acad. Sci. U.S.A.
117
,
27044
27053
112
Schulze
,
S.
,
Yu
,
L.
,
Ehinger
,
A.
,
Kolb
,
D.
,
Saile
,
S.C.
,
Stahl
,
M.
et al (
2021
)
The TIR-NBS-LRR protein CSA1 is required for autoimmune cell death in Arabidopsis pattern recognition co-receptor bak1 and bir3mutants
.
bioRxiv
113
Catanzariti
,
A.M.
,
Do
,
H.T.
,
Bru
,
P.
,
de Sain
,
M.
,
Thatcher
,
L.F.
,
Rep
,
M.
et al (
2017
)
The tomato I gene for Fusarium wilt resistance encodes an atypical leucine-rich repeat receptor-like protein whose function is nevertheless dependent on SOBIR1 and SERK3/BAK1
.
Plant J.
89
,
1195
1209
114
Catanzariti
,
A.M.
,
Lim
,
G.T.T.
and
Jones
,
D.A.
(
2015
)
The tomato I-3 gene: a novel gene for resistance to fusarium wilt disease
.
New Phytol.
207
,
106
118
115
Leibman-Markus
,
M.
,
Pizarro
,
L.
,
Bar
,
M.
,
Coaker
,
G.
and
Avni
,
A.
(
2018
)
NRC proteins: a critical node for pattern and effector mediated signaling
.
Plant Signal. Behav.
13
,
e1507404
116
Pruitt
,
R.N.
,
Locci
,
F.
,
Wanke
,
F.
,
Zhang
,
L.
,
Saile
,
S.C.
,
Joe
,
A.
et al (
2021
)
The EDS1-PAD4-ADR1 node mediates Arabidopsis thaliana pattern-triggered immunity
.
Nature
in press
117
Tian
,
H.
,
Chen
,
S.
,
Wu
,
Z.
,
Ao
,
K.
,
Yaghmaiean
,
H.
,
Sun
,
T.
et al (
2021
)
Activation of TIR signaling is required for pattern-triggered immunity
.
bioRxiv
118
Kadota
,
Y.
,
Liebrand
,
T.W.H.
,
Goto
,
Y.
,
Sklenar
,
J.
,
Derbyshire
,
P.
,
Menke
,
F.L.H.
et al (
2019
)
Quantitative phosphoproteomic analysis reveals common regulatory mechanisms between effector- and PAMP-triggered immunity in plants
.
New Phytol.
221
,
2160
2175
119
Navarro
,
L.
,
Zipfel
,
C.
,
Rowland
,
O.
,
Keller
,
I.
,
Robatzek
,
S.
,
Boller
,
T.
et al (
2004
)
The transcriptional innate immune response to flg22. interplay and overlap with Avr gene-dependent defense responses and bacterial pathogenesis
.
Plant Physiol.
135
,
1113
1128
120
Gao
,
Z.
,
Chung
,
E.H.
,
Eitas
,
T.K.
and
Dangl
,
J.L.
(
2011
)
Plant intracellular innate immune receptor resistance to pseudomonas syringae pv. maculicola 1 (RPM1) is activated at, and functions on, the plasma membrane
.
Proc. Natl Acad. Sci. U.S.A.
108
,
7619
7624
121
Roberts
,
M.
,
Tang
,
S.
,
Stallmann
,
A.
,
Dangl
,
J.L.
and
Bonardi
,
V.
(
2013
)
Genetic requirements for signaling from an autoactive plant NB-LRR intracellular innate immune receptor
.
PLoS Genet.
9
,
e1003465
122
Hatsugai
,
N.
,
Iwasaki
,
S.
,
Tamura
,
K.
,
Kondo
,
M.
,
Fuji
,
K.
,
Ogasawara
,
K.
et al (
2009
)
A novel membrane fusion-mediated plant immunity against bacterial pathogens
.
Genes Dev.
23
,
2496
2506
123
Tsushima
,
A.
,
Narusaka
,
M.
,
Gan
,
P.
,
Kumakura
,
N.
,
Hiroyama
,
R.
,
Kato
,
N.
et al (
2021
)
The conserved Colletotrichum spp. effector CEC3 induces nuclear expansion and cell death in plants
.
bioRxiv
124
Ding
,
X.
,
Jimenez-Gongora
,
T.
,
Krenz
,
B.
and
Lozano-Duran
,
R.
(
2019
)
Chloroplast clustering around the nucleus is a general response to pathogen perception in Nicotiana benthamiana
.
Mol. Plant Pathol.
20
,
1298
1306
125
Mur
,
L.A.
,
Kenton
,
P.
,
Lloyd
,
A.J.
,
Ougham
,
H.
and
Prats
,
E.
(
2008
)
The hypersensitive response; the centenary is upon us but how much do we know?
J. Exp. Bot.
59
,
501
520
126
Shen
,
W.
,
Liu
,
J.
and
Li
,
J.F.
(
2019
)
Type-II metacaspases mediate the processing of plant elicitor peptides in Arabidopsis
.
Mol. Plant
12
,
1524
1533
127
Hander
,
T.
,
Fernandez-Fernandez
,
A.D.
,
Kumpf
,
R.P.
,
Willems
,
P.
,
Schatowitz
,
H.
,
Rombaut
,
D.
et al (
2019
)
Damage on plants activates Ca2+-dependent metacaspases for release of immunomodulatory peptides
.
Science
363
,
eaar7486
128
Choi
,
H.W.
and
Klessig
,
D.F.
(
2016
)
DAMPs, MAMPs, and NAMPs in plant innate immunity
.
BMC Plant Biol.
16
,
232
129
Hartmann
,
M.
,
Zeier
,
T.
,
Bernsdorff
,
F.
,
Reichel-Deland
,
V.
,
Kim
,
D.
,
Hohmann
,
M.
et al (
2018
)
Flavin monooxygenase-generated N-hydroxypipecolic acid is a critical element of plant systemic immunity
.
Cell
173
,
456
469 e16
130
Hartmann
,
M.
and
Zeier
,
J.
(
2019
)
N-hydroxypipecolic acid and salicylic acid: a metabolic duo for systemic acquired resistance
.
Curr. Opin. Plant Biol.
50
,
44
57
131
Chen
,
Y.-C.
,
Holmes
,
E.C.
,
Rajniak
,
J.
,
Kim
,
J.-G.
,
Tang
,
S.
,
Fischer
,
C.R.
et al (
2018
)
N-hydroxy-pipecolic acid is a mobile metabolite that induces systemic disease resistance in Arabidopsis
.
Proc. Natl Acad. Sci. U.S.A.
115
,
E4920
E49E9
132
Liu
,
X.
,
Zhang
,
Z.
,
Ruan
,
J.
,
Pan
,
Y.
,
Magupalli
,
V.G.
,
Wu
,
H.
et al (
2016
)
Inflammasome-activated gasdermin D causes pyroptosis by forming membrane pores
.
Nature
535
,
153
158
133
Daskalov
,
A.
,
Habenstein
,
B.
,
Sabate
,
R.
,
Berbon
,
M.
,
Martinez
,
D.
,
Chaignepain
,
S.
et al (
2016
)
Identification of a novel cell death-inducing domain reveals that fungal amyloid-controlled programmed cell death is related to necroptosis
.
Proc. Natl Acad. Sci. U.S.A.
113
,
2720
2725
134
Flores-Romero
,
H.
,
Ros
,
U.
and
Garcia-Saez
,
A.J.
(
2020
)
Pore formation in regulated cell death
.
EMBO J.
39
,
e105753
135
Tian
,
W.
,
Hou
,
C.
,
Ren
,
Z.
,
Wang
,
C.
,
Zhao
,
F.
,
Dahlbeck
,
D.
et al (
2019
)
A calmodulin-gated calcium channel links pathogen patterns to plant immunity
.
Nature
572
,
131
135
136
Zhou
,
J.M.
and
Zhang
,
Y.
(
2020
)
Plant immunity: danger perception and signaling
.
Cell
181
,
978
989
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).