Fact-checked by Grok 2 weeks ago

Hypersensitive response

The (HR) is a rapid, localized in that occurs at the site of pathogen penetration, serving as a key component of effector-triggered immunity (ETI) to restrict the spread of incompatible microbial invaders. This defense mechanism is typically activated when plant nucleotide-binding (NLR) proteins detect specific pathogen effectors, either directly or indirectly through modifications to host targets, leading to an amplified response that includes a burst of (ROS), ion fluxes, and the production of antimicrobial compounds such as phytoalexins and pathogenesis-related (PR) proteins. The resulting necrotic lesions, often visible within hours of , effectively trap and starve biotrophic and hemibiotrophic by sacrificing infected cells, thereby enhancing localized resistance. At the molecular level, HR exemplifies the gene-for-gene hypothesis, where specific plant resistance (R) genes interact with corresponding pathogen avirulence (Avr) factors to initiate signaling cascades involving salicylic acid accumulation and mitogen-activated protein kinase (MAPK) pathways. Unlike pattern-triggered immunity (PTI), which provides basal defense against a broad range of microbes, HR represents a more robust, tailored reaction that can also induce systemic acquired resistance (SAR), priming distal tissues for enhanced defense against secondary infections. However, HR is not universally protective; it may benefit necrotrophic pathogens that thrive on dead tissue and can impose fitness costs on the host, such as reduced growth in the absence of pathogens. Beyond its role in , shares features with animal and has been studied for its evolutionary conservation across species, influencing breeding strategies for disease-resistant crops. Research continues to elucidate how integrates with other immune outputs, including stomatal closure and callose deposition, to form a multilayered network.

Overview

Definition and characteristics

The hypersensitive response (HR) is a core defense mechanism in plants, characterized by rapid, localized programmed cell death (PCD) at the site of pathogen ingress during incompatible interactions, which effectively restricts pathogen proliferation and spread. This response confines the infection to a small area, preventing systemic dissemination and thereby protecting the plant's overall health. HR is evolutionarily conserved across higher plant species, underscoring its fundamental role in innate immunity. Key observable traits of HR include the collapse of infected cells, leading to tissue and the formation of discrete, visible lesions typically within hours to days after detection. These lesions serve as physical barriers, containing the and limiting its access to nutrients and further tissues. A classic example is the localized necrotic spots observed in tobacco resistance to (TMV), first described in 1929 by Francis O. Holmes in Nicotiana glutinosa, marking the restriction of viral spread. Unlike compatible plant-pathogen interactions, which allow successful and systemic due to the 's ability to evade or suppress defenses, HR occurs exclusively in resistant combinations, resulting in halted pathogen growth and no disease progression beyond the initial site. This distinction highlights HR's specificity in gene-for-gene resistance models, often involving plant resistance genes that recognize pathogen effectors.

Historical context

The hypersensitive response () in was first described over a century ago as a localized form of associated with to fungal . In 1902, H. Marshall Ward observed rapid necrosis in leaves infected with fungi, characterizing it as a defensive limiting pathogen spread. Similar observations followed shortly thereafter, with E.C. Gibson reporting localized in varieties resistant to Puccinia chrysanthemi in 1904, and Dorothea Marryat documenting immune responses in to Puccinia glumarum in 1907. These early accounts established HR as a key mechanism of non-host and gene-based , though the "hypersensitive response" was formalized later by Elvin C. Stakman in 1915, who emphasized its role in cereal interactions. The application of HR to viral pathogens gained prominence in the early 20th century through studies on tobacco mosaic virus (TMV). In 1929, Francis O. Holmes identified local necrotic lesions in Nicotiana glutinosa leaves inoculated with TMV, marking the first detailed description of HR-like localized cell death in a virus-host interaction and enabling the development of the local lesion assay for virus quantification. This phenotype became a model for studying resistance, particularly after the transfer of the N gene from N. glutinosa to cultivated tobacco (Nicotiana tabacum) in the 1930s, which induced similar hypersensitive lesions upon TMV infection. During the mid-20th century, conceptual frameworks linked to genetic specificity. Harold H. Flor's gene-for-gene hypothesis, proposed in 1956 based on flax-rust interactions, posited that plant resistance (R) genes correspond to specific pathogen avirulence (Avr) factors, triggering when matched; this model was extended to viral systems like TMV in during the and . Advancements in the 1980s utilized electron microscopy to reveal ultrastructural features of , such as plasma membrane invagination, organelle degradation, and nuclear fragmentation, drawing parallels to animal and confirming as a programmed cell death () process. Key experimental techniques emerged in the late to study HR in model systems. Infiltration assays, involving syringe or vacuum delivery of pathogens into leaf tissues, were refined in the 1990s using to induce and quantify HR lesions, facilitating genetic screens and effector studies. A major milestone came in 1994 with the cloning of the tobacco N gene, the first R gene against a , which encodes a Toll-interleukin-1 receptor-like protein that recognizes TMV's domain, solidifying the genetic basis of HR.

Genetic Foundations

Key resistance genes

Resistance (R) genes in are dominant alleles that confer specific recognition of avirulence (Avr) effectors, triggering the hypersensitive response () as a localized to restrict spread. These genes encode proteins that detect -derived molecules either directly or indirectly, initiating defense signaling that culminates in . The major classes of R genes include nucleotide-binding (NLR) proteins, which are predominantly cytoplasmic sensors; receptor-like s (RLKs), which are transmembrane proteins with extracellular sensing domains and intracellular activity; and receptor-like proteins (RLPs), which lack domains but associate with RLKs for signaling. NLRs constitute the largest group, comprising about 61% of cloned R genes, and are exemplified by RPM1 and RPS2 in , which mediate against the bacterial pv. tomato through recognition of the AvrRpm1 and AvrRpt2 effectors, respectively. In (), the N gene, an NLR, recognizes the helicase domain of the replicase protein of (TMV), eliciting to confine viral replication. RLPs, such as the Cf genes in tomato (Solanum lycopersicum), detect effectors from the fungal Cladosporium fulvum; for instance, Cf-9 interacts with Avr9 to induce . RLKs, like XA21 in rice (), confer against Xanthomonas oryzae pv. oryzae by perceiving the bacterial sulfated protein Ax21. Genetic mapping and of R genes accelerated in the post-1990s era, building on earlier map-based approaches to isolate functional alleles. Landmark clonings included the Pto gene in 1993, followed by RPS2, , and Cf-9 in 1994, revealing that R proteins localize either cytoplasmically (NLRs) or at the cell surface (RLKs and RLPs). By 2018, over 300 R genes had been cloned, with advanced techniques like RenSeq enabling rapid identification in diverse species. As of 2023, more than 450 R genes have been cloned from 42 plant species. These efforts demonstrated that R products often feature leucine-rich repeats (LRRs) for effector binding and variable domains for specificity. Polymorphism in R genes, particularly in LRR regions, drives diversity in natural plant populations by enabling adaptation to evolving pathogens under balancing selection. For example, Arabidopsis R gene clusters exhibit high nucleotide diversity and allele divergence, correlating with pathogen pressure and facilitating broad resistance spectra. This variability underpins the evolutionary arms race, with allelic series like those in flax R genes showing distinct HR specificities against rust effectors.

NLR proteins and variants

NLR proteins, also known as -binding receptors, serve as central effectors in the hypersensitive response (HR) by detecting effectors intracellularly and initiating localized to restrict infection. These multidomain proteins typically consist of an N-terminal signaling , a central NB-ARC responsible for binding and oligomerization, and a C-terminal () that facilitates effector sensing. The NB-ARC , part of the STAND ATPase family, toggles between an inactive ADP-bound state and an active ATP-bound conformation, enabling conformational changes that drive immune activation. The varies in repeat number and sequence, allowing specificity in recognizing diverse effectors either directly or indirectly. NLRs exhibit structural variants primarily distinguished by their N-terminal domains, which dictate signaling outputs. TIR-NLRs (TNLs) feature a Toll/interleukin-1 receptor (TIR) domain and include examples like RPP1 in Arabidopsis thaliana, which recognizes effectors from the oomycete pathogen Hyaloperonospora arabidopsidis to trigger HR. In contrast, CC-NLRs (CNLs) possess a coiled-coil (CC) domain, as seen in RPS5, which confers resistance to Pseudomonas syringae by detecting the effector AvrPphB-induced modification of the host kinase PBS1, leading to HR. Within these classes, NLRs function as sensors, which directly perceive effectors, or helpers, which amplify signals from multiple sensors without independent recognition; helper examples include the ADR1 family in Arabidopsis and the NRC family in Solanaceae, which integrate into broader immune networks. Many NLRs operate in pairs, following an integrated model where one NLR incorporates a mimicking a target, enabling the pair to sense effector activity and activate . In this model, the sensor NLR detects effector binding or modification to its integrated , disrupting autoinhibition and activating the paired helper NLR to initiate ; a example is the TIR-NLR pair RRS1/RPS4, where RRS1's C-terminal WRKY acts as a for effectors like AvrRps4, triggering RPS4 oligomerization and immune signaling. Similar occurs in with RGA4/RGA5, where RGA5's RATX1 serves as the for Magnaporthe oryzae effectors. NLR networks extend beyond pairs to monitor guardee proteins, host factors susceptible to effector manipulation, such that perturbation of the guardee activates the NLR and elicits HR. For instance, RPS5 guards the kinase PBS1; cleavage of PBS1 by AvrPphB exposes a cryptic motif that binds RPS5's LRR, promoting its activation. Helper NLRs like ADR1 coordinate these networks by responding to signals from multiple sensor NLRs, enhancing robustness against diverse pathogens. These interactions form intricate webs, where guardee monitoring integrates with pairing to fine-tune effector detection. Recent structural studies using cryo-electron microscopy (cryo-EM) since 2020 have illuminated NLR dynamics in inactive and active states, revealing mechanisms of autoinhibition and . For example, the 2020 cryo-EM structure of the TIR-NLR RPP1 resistosome showed tetramerization upon effector recognition, with the TIR domain forming a channel-like assembly. In 2023-2024, structures of helper CC-NLRs like NRC2 and NRC4 demonstrated dimer-to-hexamer transitions stabilized by phosphates, underscoring exchange and oligomerization as key to HR induction. These findings highlight how LRR-mediated effector binding relieves autoinhibitory interactions, facilitating resistosome assembly essential for downstream signaling.

Activation Mechanisms

Pathogen recognition models

The gene-for-gene model, first proposed by H. H. Flor, posits that specific resistance (R) genes in plants correspond to avirulence (Avr) genes in pathogens, such that the presence of matching R and Avr gene products triggers a defense response including the hypersensitive response (HR). This model explains the specificity of plant-pathogen interactions, where direct recognition of the Avr effector by the R protein is thought to occur in rare cases, such as the interaction between the bacterial effector AvrPto and the tomato R protein Pto, leading to HR initiation. Experimental evidence for direct binding has been demonstrated through yeast two-hybrid assays, where AvrPto and Pto physically interact, and co-immunoprecipitation (co-IP) confirming this association in plant cells. Indirect recognition models predominate in detection, addressing the challenge of evolving effectors that avoid direct binding. The guard hypothesis proposes that nucleotide-binding (NLR) R proteins monitor or "guard" host target proteins (guardees) modified by effectors, activating HR upon detecting these perturbations. A classic example is the RPM1 NLR, which guards the host protein RIN4; bacterial effectors like AvrRpm1 and AvrB phosphorylate RIN4, triggering RPM1 activation and HR, as evidenced by genetic studies showing RIN4 mutations abolish RPM1-dependent resistance. Co-IP experiments have confirmed RIN4's physical association with both effectors and RPM1, supporting the model's mechanistic basis. Complementing the guard model, the decoy hypothesis suggests that certain host proteins act as non-functional mimics (decoys) of true effector targets, luring effectors to bind and thereby activating associated NLRs without the decoy serving a in normal . For instance, RIN4 can function in a decoy-like manner for some effectors, where its modification signals NLR independent of its regulatory functions, as shown in studies differentiating guard and decoy outcomes through pathogen growth assays. This model has been validated by yeast two-hybrid and co-IP data revealing effector-decoy interactions that indirectly engage NLRs, such as in the case of AvrPphB targeting PBS1, a decoy cleaved to activate the NLR RPS5. Receptor-like proteins (RLPs) and receptor-like kinases (RLKs) provide an extracellular layer of recognition, detecting pathogen-associated molecular patterns (PAMPs) or effectors at the cell surface to initiate signaling that can escalate to HR in effector-triggered immunity (ETI).00501-0) The Arabidopsis RLK EFR exemplifies this by binding the bacterial PAMP elongation factor Tu (EF-Tu) or its derived elf18 peptide, triggering immune responses that integrate with intracellular NLR signaling for robust HR in incompatible ETI contexts.00501-0) Ligand-binding assays and genetic complementation have confirmed EFR's specificity, with co-IP showing its complex formation with co-receptors like BAK1 to amplify signaling.00501-0) Pathogen recognition operates in a multi-layered framework, where pattern-triggered immunity (PTI) via RLPs/RLKs provides basal defense that pathogens suppress with effectors, prompting ETI through NLRs and culminating in HR when PTI and ETI signals synergize. This integration amplifies shared downstream pathways, such as MAPK cascades, with ETI restoring PTI-suppressed responses for enhanced HR, as demonstrated by transcriptomic analyses showing overlapping gene induction in PTI-ETI hybrids. Yeast two-hybrid and co-IP studies of NLR-effector-guardee tripartite interactions further illustrate how initial PTI priming sensitizes ETI detection, ensuring rapid HR deployment.

Resistosome assembly

Upon activation by pathogen effectors, nucleotide-binding leucine-rich repeat (NLR) immune receptors in undergo conformational changes that lead to the assembly of oligomeric complexes known as resistosomes. These structures typically consist of 4 to 8 NLR subunits, forming wheel-like or funnel-shaped architectures that facilitate downstream hypersensitive response () signaling. The assembly process is initiated by a nucleotide exchange in the NB-ARC domain, where the inactive ADP-bound state is displaced by ATP , relieving auto-inhibition and promoting oligomerization. This switch-like mechanism exposes oligomerization interfaces in the (LRR) and central nucleotide- (NB) domains, allowing multiple NLR monomers to associate into a stable resistosome. Structural studies have revealed that in the auto-inhibited state, the LRR domain sterically hinders the NB-ARC domain, preventing ATP ; effector recognition disrupts this inhibition, enabling the transition.00141-X) Resistosomes execute HR signaling through distinct mechanisms depending on the NLR subclass. In coiled-coil NLRs (CNLs), such as ZAR1, the resistosome forms a pentameric complex with helper proteins like HOPZ-ETI, adopting a funnel-shaped structure that inserts into the plasma membrane to create cation-permeable pores, triggering ion fluxes including Ca²⁺ influx that amplify defense responses. In toll/interleukin-1 receptor NLRs (TNLs), exemplified by ROQ1, the tetrameric resistosome positions TIR domains to form a holoenzyme with NADase activity, cleaving NAD⁺ into variants like ADPr and vDAP, which serve as signaling molecules to propagate .00141-X) Post-2020 cryo-EM structures have illuminated these processes in greater detail, including auto-inhibitory configurations and activation triggers in diverse NLRs like Sr35 and RPP1, confirming conserved oligomerization motifs while highlighting subclass-specific adaptations for membrane association and enzymatic output.00627-8)

NLR networks and pairing

NLR immune receptors in frequently operate as sensor-helper pairs, where the NLR detects effectors and recruits a helper NLR to amplify signaling and trigger the hypersensitive response (). In , the sensor NLR Pik-1, containing an integrated heavy metal-associated (HMA) domain, recognizes effectors such as AVR-Pik from the blast fungus Magnaporthe oryzae, leading to the recruitment of the helper NLR Pik-2, which facilitates downstream immune activation including . This paired architecture ensures specificity in effector recognition while enhancing signal robustness, as Pik-1 alone induces weak responses, but co-expression with Pik-2 robustly elicits . Similarly, in , the Pias-1/Pias-2 pair detects the Magnaporthe oryzae effector AVR-Pias, demonstrating conserved sensor-helper functionality across monocots. Beyond simple pairs, NLRs form intricate networks through indirect interactions mediated by integrated domains (IDs) or chaperone complexes, enabling coordinated responses to diverse pathogens. Sensor NLRs with IDs, such as the HMA in Pik-1, facilitate effector binding and subsequent networking with helpers via shared structural motifs that promote oligomerization. Chaperone complexes, particularly the HSP90-SGT1-RAR1 system, stabilize NLR proteins and support their assembly into functional networks by preventing auto-activation and aiding maturation. In the family, the ROQ1 sensor NLR recognizes effector XopQ and effector HopQ1, integrating into the NRC (NLR required for ) network of helper NLRs (NRC1-4) that amplify HR signaling across multiple sensors. In , ADR1-family helpers (ADR1, ADR1-L1, ADR1-L2) enhance responses from diverse sensors like RPS2 and RPP4, forming a that boosts HR intensity without direct effector recognition. These networked and paired configurations have evolved through gene duplications, generating specialized architectures that fine-tune HR specificity and robustness against evolving pathogens. Ancestral NLR pairs underwent tandem duplications and diversification, as seen in the expansion of the NRC superclade in asterids from a single sensor-helper progenitor into a dispersed network comprising up to half of the NLR repertoire in some Solanaceae species. This evolutionary process allows subfunctionalization, where duplicated NLRs partition roles in pathogen detection and signaling, reducing fitness costs while broadening immune coverage. In Arabidopsis, duplications within the ADR1 clade similarly created redundant helpers that collectively guard against effector targeting.00305-X) Experimental evidence for NLR pairing and networking relies heavily on transient expression assays in model plants like , which recapitulate in a controlled manner. Co-expression of Pik-1 and Pik-2 with AVR-Pik elicits effector-dependent , whereas individual NLRs fail to do so, confirming the necessity of pairing for activation. Similarly, ROQ1-triggered in requires NRC helpers, as silencing NRC genes abolishes the response in these assays. These approaches highlight pair- and network-dependent mechanisms, linking genetic interactions to physiological outcomes without stable transformation.

Regulatory Processes

Activation thresholds

The activation threshold of the hypersensitive response () in is determined by several key factors that modulate the sensitivity and timing of effector-triggered immunity (ETI), balancing effective restriction with minimal costs to the host. Effector concentration plays a critical role, as higher levels of pathogen-delivered effectors can overcome recognition barriers, facilitating NLR activation and subsequent HR initiation when delivered via type III secretion systems. Similarly, the binding affinity between resistance (R) proteins and avirulence (Avr) effectors sets a quantitative for response elicitation; for instance, a 10-fold reduction in affinity between AVR-PikE and the Pikp-HMA domain diminishes HR triggering, indicating that sufficient molecular interactions are required to surpass the activation barrier. Cellular , particularly ATP levels, further influences this threshold, as NLR proteins undergo exchange from to ATP upon effector perception, enabling resistosome oligomerization essential for HR signaling; disruptions in cytosolic ATP correlate with altered intracellular morphology and delayed hypersensitive in cells. HR activation is developmentally regulated, with full responses typically restricted to adult tissues to prevent deleterious during early growth phases. In seedlings, HR is often suppressed through microRNA-mediated silencing of NLR genes, such as miR482/2118, which targets multiple resistance loci to maintain developmental and avoid spontaneous ; this suppression diminishes as plants mature, allowing robust ETI in vegetative stages. NLR pairing can briefly influence these thresholds by enhancing signal amplification in mature tissues, though detailed mechanisms are addressed elsewhere. Environmental cues like and fine-tune HR thresholds by modulating (ROS) production and signaling fidelity. Higher intensity enhances chloroplast-derived ROS, lowering the activation threshold for HR by amplifying oxidative bursts necessary for execution, as observed in where light acclimation accelerates hypersensitive lesion formation. Temperature exerts a suppressive effect at elevated levels (above 20–30°C), inhibiting NB-LRR protein stability and R gene-mediated responses, thereby raising the threshold for HR induction and promoting susceptibility. In quantitative resistance conferred by polygenic traits, partial HR manifests as attenuated cell death responses that lower the overall activation threshold compared to monogenic qualitative resistance, enabling finer control over defense intensity across diverse pathogen pressures. This polygenic architecture contributes to durable resistance by integrating multiple weak ETI signals, reducing lesion spread without complete host sacrifice. Threshold variations are commonly assessed through lesion size measurements, where smaller or irregular HR lesions correlate with higher activation sensitivity in resistant genotypes, as quantified in maize association studies linking genomic loci to phenotypic metrics of cell death extent.

Feedback and suppression

The hypersensitive response (HR) in plants is fine-tuned by positive feedback mechanisms that amplify initial signaling to ensure robust defense activation. Mitogen-activated protein kinase (MAPK) cascades play a central role in this amplification, where upstream kinases activate downstream MAPKs to propagate signals leading to HR execution. For instance, in , the MAPK cascade involving MEKK1, MKK4/MKK5, and MPK3/MPK6 positively regulates defense gene expression and HR upon pathogen recognition, creating a feed-forward loop that sustains production and signaling. This positive regulation contrasts with other MAPKs like MPK4, which typically dampen responses but can integrate into broader cascades for signal intensification in specific contexts. Negative regulatory loops counteract excessive HR to prevent runaway cell death and autoimmunity. Protein phosphatases, such as protein phosphatase 2A (PP2A), deactivate key immune components by dephosphorylating upstream regulators like BAK1, thereby terminating signaling. Additionally, ubiquitination pathways promote NLR degradation via the 26S proteasome, providing a post-translational brake on HR. The SCF^{CPR1} complex, for example, polyubiquitinates activated NLRs like SNC1, facilitating their turnover and limiting sustained immune responses. Reversible ubiquitination further allows dynamic control in plant immunity. Hormonal signals integrate into these feedback loops to modulate HR intensity and duration. Salicylic acid (SA) promotes HR by stabilizing NLR signaling and enhancing execution, as elevated SA levels in infected tissues sustain MAPK activation and defense gene expression. In contrast, (JA) antagonizes HR duration by suppressing SA-dependent pathways, often through crosstalk at transcription factors like NPR1, thereby attenuating prolonged in favor of growth recovery. This antagonism is evident in dual-pathogen challenges, where JA signaling shortens HR to prevent . Cell-autonomous controls, such as , degrade activated immune complexes post-HR to restore . selectively targets NLR oligomers and resistosomes via ATG8-mediated engulfment, preventing their accumulation and excessive signaling after pathogen clearance. In and , autophagy-deficient mutants exhibit amplified HR and heightened susceptibility due to undegraded defense components, underscoring its role in terminating localized . Mutations in regulatory genes highlight the precision of these feedbacks, with loss-of-function leading to uncontrolled HR. The Arabidopsis ssi2 mutant, defective in plastidial glycerol-3-phosphate acyltransferase, accumulates SA and exhibits spontaneous, spreading lesions resembling runaway HR, coupled with dwarfism and constitutive defense activation. This phenotype demonstrates how disrupted lipid signaling disrupts negative feedbacks, causing unchecked cell death propagation.

Key Mediators

Reactive oxygen species

The hypersensitive response (HR) in plants is characterized by a rapid oxidative burst, during which (ROS) such as (H₂O₂) and (O₂⁻) are produced in large quantities. This burst is primarily driven by plasma membrane-bound NADPH oxidases, specifically respiratory burst oxidase homologs D and F (RBOHD and RBOHF), which catalyze the transfer of electrons from NADPH to oxygen, generating apoplastic that is subsequently dismutated to H₂O₂. The production is biphasic: an initial occurs within minutes of , followed by a sustained that amplifies the response. In addition to NADPH oxidases, other cellular compartments contribute to the ROS pool during HR. Chloroplasts generate ROS through over-reduction of the photosynthetic , particularly under light conditions, producing , , and H₂O₂. Peroxisomes also serve as significant sources, releasing ROS via and β-oxidation processes, which add to the overall oxidative at sites. These multiple origins ensure a robust and multifaceted ROS accumulation that supports localized defense. ROS fulfill critical functions in HR as both direct effectors and signaling molecules. They exert antimicrobial effects by directly damaging pathogen cells through oxidation of proteins, lipids, and DNA, thereby restricting pathogen proliferation at the invasion site. In terms of cell death induction, elevated ROS levels trigger hypersensitive cell death via lipid peroxidation of plasma and organelle membranes, leading to membrane rupture and irreversible cellular damage. Furthermore, ROS act as redox signals, interacting with sensors such as thioredoxins and peroxiredoxins to activate downstream defense pathways, including gene expression for pathogenesis-related proteins. A key aspect of ROS involvement in HR is their propagation as waves that amplify the response beyond the initial infection focus. These ROS waves, primarily H₂O₂-based, originate in chloroplasts or the and spread to adjacent cells via activation in a relay-like manner. This propagation enhances HR by priming neighboring tissues for activation without causing widespread damage. To prevent uncontrolled ROS accumulation and unintended tissue damage, employ antioxidants that maintain redox balance during HR. (SOD) enzymes, localized in various compartments including chloroplasts and peroxisomes, rapidly convert to H₂O₂, facilitating controlled signaling while mitigating . Other antioxidants, such as catalases and ascorbate peroxidases, further scavenge excess ROS, ensuring the HR remains confined and effective.

Ion fluxes and signaling molecules

During the hypersensitive response () in , rapid influx of calcium ions (Ca²⁺) into the cytosol occurs through plasma membrane channels, generating characteristic spikes in cytosolic Ca²⁺ concentration ([Ca²⁺]cyt) that serve as early signaling events. Cyclic nucleotide-gated channels (CNGCs), such as AtCNGC2 and AtCNGC4 in , mediate this Ca²⁺ influx in response to recognition, contributing to downstream defense activation including production and . These channels are activated by cyclic nucleotides like cGMP or , often elevated during attack, and their dysfunction in mutants (e.g., dnd1 for AtCNGC2) impairs HR while enhancing resistance to certain pathogens like . The Ca²⁺ signals are decoded by Ca²⁺-binding proteins, including calmodulins (CaMs), which interact with CNGCs to modulate channel activity and propagate signals to transcription factors and defense enzymes. Potassium (K⁺) efflux and anion efflux from cells are prominent ion movements during , leading to membrane that amplifies local and systemic signaling. elicitors like harpin from Erwinia amylovora rapidly enhance K⁺ outward-rectifying currents within minutes, promoting K⁺ loss and contributing to leakage associated with in and suspension cells. Concurrently, anion efflux, particularly of (Cl⁻), through channels such as ALMT12, drives membrane , which in turn activates voltage-gated K⁺ channels for further efflux. This cascade restricts spread and induces stomatal closure in , limiting apoplastic bacterial entry during defense responses. Nitric oxide (NO) emerges as a key redox-active signaling molecule during HR, functioning to potentiate and defense in response to avirulent pathogens. In cells challenged with pv. glycinea, exogenous NO donors like enhance hypersensitive , while NO scavengers suppress it, indicating NO's role in amplifying HR signals through S-nitrosylation of target proteins. NO synthesis, primarily via or nitric oxide synthase-like activity, integrates with Ca²⁺ signaling to regulate mitogen-activated protein kinases and hypersensitive response outcomes. Additionally, (SA) biosynthesis is upregulated during HR, with levels rising dramatically in infected tissues to reinforce local . In tobacco responding to , SA derives from trans-cinnamic acid via the pathway through , with tracer studies confirming increased flux and accumulation up to 100-fold within hours of . This SA surge activates pathogenesis-related genes and contributes to HR containment of viral spread. Symplastic diffusion through plasmodesmata enables the intercellular spread of HR signals from dying cells to neighboring tissues, facilitating coordinated defense. These cytoplasmic channels, with a typical size exclusion limit of 1-10 kDa, allow mobile molecules like precursors or small peptides to propagate signals, triggering callose deposition that subsequently closes plasmodesmata to isolate infected zones. In , proteins such as AZI1 mediate symplastic transport of derivatives from HR sites, priming distal cells for resistance without pathogen dissemination. During HR to avirulent s, this regulated diffusion ensures rapid local reinforcement while preventing uncontrolled signal overload. Pharmacological assays demonstrate the centrality of ion fluxes in HR, as channel blockers like lanthanum (La³⁺) abolish key responses. In tobacco and Arabidopsis, pretreatment with 10 mM La³⁺ inhibits Ca²⁺ influx and membrane depolarization induced by elicitors or cold shock analogs of defense, preventing HR cell death while preserving pathogen resistance in some cases. This blockade highlights Ca²⁺ channels' essential role, though La³⁺'s nonspecific effects on K⁺ and anion channels underscore the interconnectedness of ion signaling in HR execution.

Pathogen Counterstrategies

Evasion tactics

employ passive evasion tactics to circumvent the hypersensitive response () by avoiding detection by (R) genes, particularly nucleotide-binding (NLR) receptors, without actively interfering with signaling pathways. These strategies include structural and behavioral adaptations that blend pathogen components with host features or exploit temporal and spatial mismatches in immune surveillance. Such tactics enable pathogens to establish below the threshold required for HR activation, which involves rapid localized to restrict pathogen spread. One key evasion mechanism is host , where pathogens secrete effectors structurally resembling host proteins to evade surveillance. For instance, the effector RaxX from Xanthomonas oryzae pv. oryzae mimics the plant sulfated PSY; however, the non-sulfated variant of RaxX fails to activate the NLR receptor XA21, thereby avoiding HR induction and allowing bacterial proliferation. This reduces the likelihood of as a foreign , permitting stealthy host colonization. Temporal avoidance represents another passive strategy, wherein pathogens initiate low-level infections during host developmental stages or conditions when R gene expression and immune priming are suboptimal, preventing the accumulation of signals needed to trigger HR. Compartmentalization in biotrophic fungi further aids evasion by sequestering effectors away from cytoplasmic NLRs. Haustoria, specialized feeding structures, form an enclosed interface with host cells, secreting effectors into the extrahaustorial space bounded by a host-derived membrane; this spatial separation delays effector translocation into the host cytoplasm, minimizing premature NLR activation and HR. In oomycetes like Phytophthora infestans, distinct secretion pathways direct apoplastic effectors to this compartment, enhancing stealth during early infection stages. Pathogen populations also evolve genetic diversity through avirulence (Avr) gene mutations that abolish recognition motifs, leading to loss of HR triggering. In Pseudomonas syringae pv. tomato, mutations such as W200* and R228* in the avrRpt2 gene eliminate interaction with the Arabidopsis RPS2 NLR, allowing virulent growth without eliciting HR while preserving effector functionality on susceptible hosts. Similarly, aphid effectors exemplify evasion in insect vectors; salivary proteins from Myzus persicae mask pattern-triggered immunity (PTI) responses, such as the oxidative burst, thereby preventing escalation to effector-triggered immunity and associated HR. These mutations and masking tactics underscore the evolutionary arms race, where pathogen variability outpaces host surveillance.

Effector-mediated suppression

Pathogen effectors actively suppress the hypersensitive response (HR) by targeting key components of plant immune signaling, thereby promoting virulence and inhibiting localized cell death. These effectors, secreted via type III secretion systems, interfere with recognition, activation, or downstream signaling of resistance (R) proteins, particularly nucleotide-binding leucine-rich repeat (NLR) receptors that trigger HR upon pathogen detection. By directly modifying host proteins, effectors disrupt the rapid programmed cell death associated with HR, allowing pathogen proliferation in susceptible hosts. One major target of effectors is the production of reactive oxygen species (ROS), a hallmark of HR initiation mediated by respiratory burst oxidase homologs (RBOHs). For instance, the Pseudomonas syringae effector HopAO1, a tyrosine phosphatase, dephosphorylates the pattern recognition receptor EFR at a critical tyrosine residue (Y836), thereby suppressing the ROS burst generated by RBOHD and preventing HR activation in Arabidopsis thaliana. This inhibition reduces apoplastic ROS accumulation, a key early signal for HR, and enhances bacterial virulence. Effectors also interfere with NLR function by promoting their degradation or inactivation through chaperone disruption. The P. syringae effector HopBF1 mimics a host client protein to bind and phosphorylate at Ser100, inhibiting its activity and chaperone function essential for NLR stability and activation. This leads to reduced NLR accumulation, suppression of effector-triggered immunity (ETI), and diminished in plants like Nicotiana benthamiana and A. thaliana. In addition, effectors block downstream signaling pathways critical for HR execution, such as MAPK cascades and pathogenesis-related () gene expression. The P. syringae effector AvrPto directly binds and inhibits receptor-like kinases, including FLS2 and BAK1, preventing their complex formation and upon ligand binding, which in turn suppresses MAPK activation and PR gene induction associated with HR. Similarly, HopF2 ADP-ribosylates the guardee protein RIN4, a negative regulator of basal defense that is monitored by NLRs like RPS2 and RPM1; this modification alters RIN4's state, preventing NLR activation and HR in A. thaliana. HopF2 also targets MAP kinase kinases (MKKs), such as MKK4/5, via to inhibit their activity and block defense signaling. This effector-mediated suppression exemplifies an , where plants evolve new R genes to recognize and counter these suppressors, while pathogens adapt effectors to evade detection. For example, variants of RIN4 have arisen in A. thaliana accessions that resist HopF2 modification, restoring NLR-mediated HR against P. syringae. Such co-evolution drives diversity in both effector repertoires and plant immune receptors, balancing virulence and resistance across pathosystems.

Physiological Consequences

Local cell death outcomes

The hypersensitive response (HR) in culminates in (PCD) at the site of recognition, executing a localized sacrifice of infected cells to restrict microbial spread. This process involves distinct morphological and biochemical changes, including DNA laddering, where genomic DNA fragments into nucleosomal units detectable by , akin to apoptotic fragmentation in animals. Vacuolar collapse follows, releasing hydrolytic enzymes that dismantle cellular contents and facilitate rapid autolysis, often observed within hours of HR initiation. Concurrently, surviving cells at lesion borders undergo cell wall fortification through lignification and suberization, reinforcing physical barriers against egress. These events are bolstered by reinforcements that enhance . Callose deposition, a β-1,3-glucan , rapidly accumulates in plasmodesmata and s surrounding lesions, sealing intercellular connections to prevent movement. , such as lignins and , also accumulate in these borders, exerting direct effects and cross-linking components for added rigidity, thereby trapping and starving pathogens within the dead . HR-mediated PCD imposes fitness trade-offs on the host plant. The energy-intensive processes of execution and defense compound synthesis divert resources from growth and reproduction, leading to reduced and seed yield in resistant genotypes under pressure. In autoimmune mutants that constitutively activate HR-like responses, such as lesion mimic mutants, uncontrolled risks escalate, causing spontaneous and severe even without . Lesion containment during HR varies in scale, from pinpoint spots under low pathogen loads to expanding necrotic areas with higher inoculum densities, allowing adaptive restriction based on infection intensity. Microscopy evidence, including electron micrographs, reveals chromatin condensation and margination in HR-affected nuclei, mirroring apoptotic nuclear dismantling and confirming the regulated nature of this death program. These local outcomes are mediated in part by and ion fluxes that amplify signaling. Recent studies as of 2025 have proposed a 'concentric circle' model for transcellular regulation of effector-triggered immunity (ETI)-induced cell death during HR, highlighting how neighboring cells coordinate to contain lesions and prevent excessive spread.

Systemic immunity induction

The hypersensitive response (HR) at a local infection site triggers systemic acquired resistance (SAR), a salicylic acid (SA)-dependent mechanism that establishes long-distance signaling to prime uninfected distal tissues for heightened defense against secondary pathogen attacks. This priming enhances the speed and amplitude of immune responses in systemic leaves, stems, and roots without causing widespread cell death, thereby conferring broad-spectrum resistance lasting from weeks to several months. SAR is marked by the sustained upregulation of pathogenesis-related (PR) genes, such as PR1 and BGL2, which encode antimicrobial proteins and contribute to the primed state, although their direct antimicrobial role is secondary to the overall sensitization of defenses. Central to SAR induction are mobile chemical signals generated during HR that travel via the to distal tissues. (AzA), a nine-carbon , accumulates systemically following local HR and primes tissues for rapid SA biosynthesis upon reinfection, with levels increasing within 6–48 hours post-inoculation in models like Arabidopsis thaliana. (Pip), an lysine-derived amino acid, serves as a key amplifier, promoting and free radical production (e.g., and ) to enhance SAR; its systemic levels rise up to sevenfold in response to avirulent pathogens. The N-hydroxy derivative of Pip (N-OH-Pip) acts as a direct, potent mobile inducer, biosynthesized by enzymes like ALD1 and FMO1, and elicits SAR even in SA-deficient mutants by activating immune-related transcription factors. These signals often act additively, with AzA and Pip requiring lipid transfer proteins like DIR1 for transport. Local SA accumulation during HR, while not highly mobile itself, coordinates the initial synthesis of these signals. Grafting experiments provide direct evidence of SAR signal mobility, demonstrating that HR induction in one part of a chimeric plant confers resistance to distal portions. For example, inoculation of rootstocks with avirulent Pseudomonas syringae in tobacco or cucumber grafts triggers PR gene expression and reduced lesion sizes in uninfected scions, with the signal crossing graft unions in as little as 48 hours and persisting for weeks. Similar results in Arabidopsistobacco heterografts highlight the non-species-specific nature of the signal, supporting phloem-based transport of AzA, Pip, and related metabolites. Mobile RNA species further contribute to systemic immunity by facilitating long-distance of factors. Small interfering RNAs (siRNAs), including pathogen-inducible nat-siRNAs like nat-siRNAATGB2, move cell-to-cell and systemically via plasmodesmata and , targeting effector genes such as bacterial Avr genes (e.g., avrRpt2) to suppress proliferation in distal tissues. This RNA mobility integrates with by amplifying -triggered silencing, enhancing resistance to avirulent without requiring transcription in recipient tissues. Recent research as of 2024 has shown that RNA silencing is activated by N gene-mediated , further linking it to establishment. The integration of HR and SAR creates synergistic defenses particularly potent against biotrophic pathogens, such as Hyaloperonospora arabidopsidis or Pseudomonas syringae pv. tomato, by combining local confinement with systemic priming of SA-responsive pathways. This synergy restricts biotroph nutrient acquisition through coordinated upregulation of PR genes and oxidative bursts, significantly reducing disease severity in challenged systemic tissues compared to non-primed plants.

Evolutionary and Comparative Aspects

Role in plant speciation

The hypersensitive response (HR) in is mediated by () genes, particularly those encoding nucleotide-binding (NLR) proteins, which often cluster in the and serve as key loci for . These clusters experience balancing selection that maintains high allelic diversity to counter evolving pathogens, but this polymorphism can lead to hybrid incompatibilities when divergent alleles interact in interspecific crosses, contributing to . Such Dobzhansky-Muller incompatibilities arise from mismatched products triggering autoimmune-like HR in hybrids, promoting divergence. In (Zea mays), divergence from its wild progenitor teosinte (Zea mays ssp. parviglumis) involves NLR variants that have undergone rapid evolution, with teosinte-derived alleles introgressed into domesticated lines enhancing resistance but also contributing to genetic barriers in hybrids. Similarly, in the genus (including and ), mismatched R cause hybrid necrosis, as seen with a NLR of recent origin that elicits autoimmune in interspecific Solanum hybrids, exemplifying how R divergence drives postzygotic isolation. Geographic patterns of alleles reflect local to co-evolved , where balancing selection favors regionally specific variants that provide resistance advantages in native environments but reduce hybrid fitness when crossed with . For instance, NLR in wild populations shows clinal variation correlated with pressures, facilitating ecotypic . Hybrid dysgenesis, characterized by incompatible activation in F1 hybrids, leads to lethality or severe weakness, as observed in crosses between cultivated rice () and wild relatives like . This involves interactions between NLR-like loci such as HWA1 and HWA2, which trigger constitutive -like cell death and growth inhibition, acting as a barrier to during . Recent genomic studies from the 2020s, using scans for selective sweeps and blocks, reveal that variants from wild relatives have been selectively introgressed into crops during to bolster immunity, yet these often carry linked incompatibilities that trace back to events. For example, in and , NLR introgressions from wild progenitors show signatures of positive selection but also inviability risks, highlighting ' dual role in and . For instance, a 2025 study identified in driven by the Ne1 (alpha/beta ) and Ne2 interaction, triggering autoimmune responses akin to , further illustrating immune roles in postzygotic .

Parallels to animal immunity

The hypersensitive response () in shares fundamental parallels with animal innate immunity, particularly in the recognition of pathogens and the deployment of () to restrict infection. Both systems rely on receptors (PRRs) and nucleotide-binding (NLR) proteins to detect pathogen-associated molecular patterns (PAMPs) or effectors, triggering rapid defense signaling that culminates in localized . In , HR is activated by effector-triggered immunity (ETI) via NLRs, mirroring animal NLRs that form to sense intracellular threats. This conserved architecture suggests an ancient evolutionary origin, with STAND ATPases and TIR domains serving as scaffolds for immune signaling across kingdoms. A key similarity lies in the execution of , where HR exhibits hybrid features resembling animal and necroptosis more closely than . HR involves rapid plasma membrane rupture, cytoplasm shrinkage, and release of damage-associated molecular patterns (DAMPs) like , which alert neighboring cells and induce systemic resistance—analogous to , where gasdermin pores release inflammatory cytokines such as IL-1β, or necroptosis, mediated by MLKL oligomerization to form lytic pores. Unlike the non-lytic apoptotic bodies in animals, HR lacks such containment due to rigid cell walls but achieves similar pathogen containment through necrosis-like leakage. (ROS) bursts and calcium influxes regulate both HR and these animal PCD forms, amplifying signaling via MAPK cascades. Signaling pathways further underscore these parallels, with plant resistance (R) proteins functioning like animal Toll-like receptors (TLRs) in detecting avirulence factors and initiating defense . For instance, plant NLR resistosomes, such as ZAR1, oligomerize to form calcium-permeable channels, akin to animal that activate for pyroptotic execution. Co-chaperones like SGT1 in plants parallel animal complexes in stabilizing immune receptors, ensuring precise activation. Proteolytic components also converge: plant metacaspases and vacuolar processing enzymes (VPEs) drive HR, comparable to animal in and . These mechanisms highlight independent evolution from shared prokaryotic precursors, enabling effective innate immunity without adaptive components in plants.

References

  1. [1]
    The plant immune system - Nature
    Nov 16, 2006 · Plants respond to infection using a two-branched innate immune system. The first branch recognizes and responds to molecules common to many classes of microbes.
  2. [2]
    The plant hypersensitive response: concepts, control and ...
    Jul 15, 2019 · Simply put, the plant hypersensitive response (HR) is a rapid localized cell death that occurs at the point of pathogen penetration and is ...
  3. [3]
    Hypersensitive Response - an overview | ScienceDirect Topics
    The hypersensitive response, often referred to as HR, is a localized induced cell defense in the host plant at the site of infection by a pathogen (Fig. 6-10A).
  4. [4]
    Programmed cell death in the plant immune system - Nature
    Apr 8, 2011 · The hypersensitive response (HR) cell death in plants displays morphological features ... Generalized induction of defense responses in ...Immune Surveillance Systems... · Cell Death At The Center Of... · Regulators Of Plant Cell...
  5. [5]
    hypersensitive response; the centenary is upon us but how much do ...
    The resistance phenomenon known as the hypersensitive response (HR) was first described by pioneering plant pathologists around 100 years ago.
  6. [6]
    Local Lesions in Tobacco Mosaic
    HOLMES-TOBACCO MOSAIC. 43 first indications of breakdown appear. These rapidly developing necrotic lesions appear first as tiny glistening dark spots. The ...
  7. [7]
    [PDF] Controlling Tobacco Mosaic Virus in Tobacco through Resistance
    These five species all gave a hypersensitive response (HR) on the leaves that were inoculated. A hypersensitive reaction is characteristic of resistance to ...
  8. [8]
    Hypersensitive cell death, autofluorescence, and insoluble silicon ...
    Hypersensitive cell death (HR) of adaxial leaf epidermal cells of barley ... electron microscopy. Canadian Journal of Botany, 57 (1979), pp. 898-913.<|control11|><|separator|>
  9. [9]
    Bacterial Leaf Infiltration Assay for Fine Characterization of Plant ...
    Oct 1, 2015 · The following protocol describes an optimized syringe infiltration method to deliver virulent Psm ES4326 into leaves of adult soil-grown Arabidopsis plants.
  10. [10]
    Defended to the Nines: 25 Years of Resistance Gene Cloning ...
    The first R gene to be cloned, maize (Zea mays) Hm1, was published over 25 years ago, and since then, many different R genes have been identified and isolated.
  11. [11]
    The Arabidopsis thaliana RPM1 disease resistance gene product is ...
    This set of responses typically includes localized cell death at the site of pathogen infection, termed the hypersensitive response (HR). In the absence of ...
  12. [12]
    A Genome-Wide Survey of R Gene Polymorphisms in Arabidopsis
    R genes located in cluster I tend to have high levels of polymorphism and more diverged alleles.
  13. [13]
  14. [14]
    NLR immune receptors: structure and function in plant disease ...
    Aug 21, 2023 · Effectors manipulate the host by either suppressing its immune responses or by promoting its nutrient supply to increase the pathogen's fitness.
  15. [15]
    Direct pathogen-induced assembly of an NLR immune receptor ...
    Dec 4, 2020 · Nucleotide-binding/leucine-rich repeat (NLR) immune receptors detect pathogen effectors and trigger a plant's immune response.
  16. [16]
    A novel conserved mechanism for plant NLR protein pairs
    In this model, a plant protein targeted by an effector has been duplicated and fused to one member of the NLR pair, where it acts as a bait to trigger defense ...Missing: RKS1- | Show results with:RKS1-
  17. [17]
    Oligomerization-mediated autoinhibition and cofactor binding of a ...
    Jun 12, 2024 · Nucleotide-binding leucine-rich repeat (NLR) proteins play a pivotal role in plant immunity by recognizing pathogen effectors.
  18. [18]
    Article RIN4 Interacts with Pseudomonas syringae Type III Effector ...
    We further suggest that RPM1 “guards” the plant by perceiving the Avr-dependent perturbation of RIN4 and inducing disease resistance. Results. RIN4 Interacts ...Missing: paper | Show results with:paper
  19. [19]
    From Guard to Decoy: A New Model for Perception of Plant ... - NIH
    The Guard model monitors pathogen targets. The Decoy model uses a target mimic for effector perception, with a decoy having no function in the absence of its ...
  20. [20]
    PTI‐ETI synergistic signal mechanisms in plant immunity - Yu - 2024
    Mar 12, 2024 · PTI and ETI collaborate synergistically to bolster disease resistance and collectively trigger a cascade of downstream defence responses.
  21. [21]
    Rice Resistance Protein Pair RGA4/RGA5 Recognizes the ...
    Yeast two-hybrid, coimmunoprecipitation, and fluorescence resonance energy transfer–fluorescence lifetime imaging experiments revealed direct binding of AVR-Pia ...
  22. [22]
    Reconstitution and structure of a plant NLR resistosome conferring ...
    Apr 5, 2019 · We reconstituted an active complex containing the Arabidopsis coiled-coil NLR ZAR1, the pseudokinase RKS1, uridylated protein kinase PBL2 ...
  23. [23]
    Structure of the activated ROQ1 resistosome directly recognizing the ...
    Dec 4, 2020 · Here we describe the 3.8-angstrom-resolution cryo–electron microscopy structure of the activated ROQ1 (recognition of XopQ 1), an NLR native to ...
  24. [24]
    A wheat resistosome defines common principles of immune receptor ...
    Sep 26, 2022 · Here we report the cryo-electron microscopy structure of the wheat CNL Sr35 5 in complex with the effector AvrSr35 6 of the wheat stem rust pathogen.
  25. [25]
    Ligand-triggered allosteric ADP release primes a plant NLR complex
    Apr 5, 2019 · NLRs are believed to function as a nucleotide [adenosine diphosphate (ADP) or adenosine triphosphate (ATP)]–operated molecular switch, with ADP- ...
  26. [26]
    Structural basis of NLR activation and innate immune signalling in ...
    The central NB-ARC domain, consisting of NBD, HD1 and WHD, functions as a molecular switch that regulates NLR activity by binding adenosine nucleotides, ADP or ...
  27. [27]
    Plasma membrane association and resistosome formation of plant ...
    Plants express diverse intracellular immune receptors that activate defense against pathogen infections. These include “sensor” NLRs (Nucleotide-binding ...
  28. [28]
    The rice NLR pair Pikp-1/Pikp-2 initiates cell death through receptor ...
    Plant NLR immune receptors are multidomain proteins that can function as specialized sensor/helper pairs. Paired NLR immune receptors are generally thought ...
  29. [29]
    NLR receptors in plant immunity: making sense of the alphabet soup
    Aug 21, 2023 · The genetically linked Arabidopsis TIR‐NLR pair RRS1/RPS4 similarly works via negative regulation. RPS4 is constitutively active in Arabidopsis, ...
  30. [30]
    A genetically linked pair of NLR immune receptors shows ... - PNAS
    Here, we identified the Pias gene from rice, which encodes the NLR pair Pias-1 “helper” and Pias-2 “sensor.” These proteins function together to detect the ...
  31. [31]
    NLRs in plant immunity: Structural insights and molecular mechanisms
    Sensor NLRs can typically recognize specific molecules or proteins produced by pathogens and are responsible for initiating the immune response, often contain ...
  32. [32]
    The HSP90-SGT1 Chaperone Complex for NLR Immune Sensors
    Nov 17, 2008 · NLR proteins recognize, directly or indirectly, pathogen-derived molecules and trigger immune responses. To function as a sensor, NLR proteins ...
  33. [33]
    Arabidopsis ADR1 helper NLR immune receptors localize ... - PubMed
    Our results show that PM localization and stability of some RNLs and one CC-type NLR (CNL) depend on the direct interaction with PM phospholipids.
  34. [34]
    NLR receptor networks in plants | Essays in Biochemistry
    In this article, we highlight key aspects of immune receptor networks in plant NLR biology and discuss NLR network architecture, the advantages of this ...
  35. [35]
    NLR network mediates immunity to diverse plant pathogens - PNAS
    Jul 11, 2017 · We discovered that a large NLR immune signaling network with a complex genetic architecture confers immunity to oomycetes, bacteria, viruses, nematodes, and ...Nlr Network Mediates... · Results And Discussion · Nrc Clade And Its Sister...
  36. [36]
    Activation of a Plant NLR Complex through Heteromeric Association ...
    Apr 24, 2017 · A convenient platform for testing NLR activities is transient expression in Nicotiana benthamiana. Co-expression of full-length DM1 and DM2d ...
  37. [37]
    Optimization of immune receptor-related hypersensitive cell death ...
    May 2, 2022 · Optimized transient HR cell death assay conditions for NLR studies using tobacco plants are proposed. When temperature, humidity, and leaf ...
  38. [38]
    Effector‐dependent activation and oligomerization of plant NRC ...
    Many Solanaceae NLRs require NRC (NLR‐Required for Cell death) class of helper NLRs. We show here that Rpi‐amr3, a sensor NLR from Solanum americanum, detects ...
  39. [39]
    MEKK1, MKK1/MKK2 and MPK4 function together in a mitogen ...
    Nov 4, 2008 · MKK1 and MKK2 function together with MPK4 and MEKK1 in a MAP kinase cascade to negatively regulate innate immune responses in plants.
  40. [40]
    Protein Phosphatase 2A in the Regulatory Network Underlying ...
    Jun 10, 2016 · Altogether, PP2A activity is an important contributor to negative regulation of a variety of plant defense responses, notably cell death.Missing: NLRs RPM1
  41. [41]
    Reversible ubiquitination conferred by domain shuffling controls ...
    Feb 26, 2025 · Plant intracellular NLR immune receptors can function individually or in pairs to detect pathogen effectors and activate immune responses.
  42. [42]
    Conventional and unconventional ubiquitination in plant immunity
    Dec 7, 2016 · Ubiquitination is one of the most abundant types of protein post-translational modification (PTM) in plant cells.
  43. [43]
    Salicylic acid and jasmonic acid in plant immunity - PubMed Central
    Plants rely on SA to ward off biotrophic and hemibiotrophic pathogens, whereas JA-induced responses primarily contribute to defense against necrotrophic ...
  44. [44]
    Salicylic acid and jasmonic acid crosstalk in plant immunity
    The phytohormones salicylic acid (SA) and jasmonic acid (JA) are major phytohormones to mediate plant immunity against pathogens [4,5]. SA and JA signaling play ...Sa-Ja Crosstalk · Suppression Of Ja By Sa · Conclusions And PerspectivesMissing: hypersensitive | Show results with:hypersensitive
  45. [45]
    Salicylic Acid Suppresses Jasmonic Acid Signaling Downstream of ...
    Interactions between the plant hormones salicylic acid and jasmonic acid play an important role in the regulation of plant defense responses against pathog.Missing: hypersensitive | Show results with:hypersensitive
  46. [46]
    The Complex Roles of Autophagy in Plant Immunity - PMC
    Autophagy is a conserved cellular recycling process and has well established roles in nutrient starvation responses and cellular homeostasis.
  47. [47]
    Role of autophagy in disease resistance and hypersensitive ...
    A few research reports have appeared recently to shed light on the roles of autophagy in plant–pathogen interactions and in disease-associated host cell death.Missing: NLR | Show results with:NLR
  48. [48]
    Plastidial Fatty Acid Signaling Modulates Salicylic Acid
    18:1 levels in ssi2 plants were increased by performing epistatic analyses between ssi2 and several mutants in FA pathways that cause an increase in the levels ...
  49. [49]
    To die or not to die? Lessons from lesion mimic mutants - Frontiers
    Extensive progress in our understanding of plant PCD in response to stress came from forward genetic approaches and the identification of many mutants ...
  50. [50]
    ROS signaling in the hypersensitive response - PubMed Central - NIH
    The hypersensitive response (HR), first described by Stakmann in 1915, is the landmark of successful pathogen recognition during non-host and incompatible host ...Missing: definition characteristics
  51. [51]
    Full article: ROS signaling in the hypersensitive response
    4 A biphasic oxidative burst leading to the generation of reactive oxygen species (ROS) commonly precedes cell death, and the signaling role played by these ...
  52. [52]
  53. [53]
    An insight on superoxide dismutase (SOD) from plants for ...
    Superoxide dismutase (SOD) is an antioxidant enzyme functional for physiological defense strategies in animals and plants against free radicals and reactive ...
  54. [54]
  55. [55]
    Harpin, a hypersensitive response elicitor from Erwinia amylovora ...
    May 25, 2001 · In suspension cell cultures from tobacco or Arabidopsis, harpin induces early responses such as potassium efflux and the rapid inhibition of ATP ...
  56. [56]
    Nitric oxide signalling functions in plant-pathogen interactions
    In this review we will examine the synthesis of NO, its effects, functions and signalling giving rise to the hypersensitive response and systemic acquired ...
  57. [57]
    Nitric oxide functions as a signal in plant disease resistance - PubMed
    Nitric oxide, which acts as a signal in the immune, nervous and vascular systems, potentiates the induction of hypersensitive cell death in soybean cells.Missing: NO | Show results with:NO
  58. [58]
    Pathway of Salicylic Acid Biosynthesis in Healthy and Virus ...
    During the hypersensitive response of Nicotiana tabacum L. cv Xanthi-nc to tobacco mosaic virus (TMV), SA levels rise dramatically.
  59. [59]
    Nonselective Block by La3+ of Arabidopsis Ion Channels Involved in ...
    Lanthanide ions such as La3+ are frequently used as blockers to test the involvement of calcium channels in plant and animal signal transduction pathways.Missing: hypersensitive response
  60. [60]
    Molecular mimicry modulates plant host responses to pathogens
    Nov 22, 2017 · Together, these studies indicate that molecular mimics can suppress host immune responses, facilitate infection and/or enhance plant health, ...Missing: hypersensitive tactics
  61. [61]
    Avoidance and suppression of plant defenses by herbivores and ...
    While prolonged pathogen infection can give rise to local cell death, i.e. the hypersensitive ... Plant pathogens and integrated defence responses to infection.Signal Transduction In Plant... · Avoidance Of Plant Defenses · Suppression Of Plant...
  62. [62]
  63. [63]
    A Functional Genomics Approach Identifies Candidate Effectors from ...
    Nov 18, 2010 · To determine whether aphid candidate effectors can suppress PTI, we assessed whether any of our 48 candidates suppressed the oxidative burst ...Missing: masking HR
  64. [64]
    How pathogen effectors suppress NLR-mediated immunity
    Effectors disrupt NLR-mediated immunity by directly binding or indirectly affecting recognition, signalling, and/or NLR protein homeostasis.
  65. [65]
    Suppression of NLR-mediated plant immune detection by bacterial ...
    In this review we focus on bacterially driven suppression of the latter, known as effector-triggered immunity (ETI) and dependent on diverse NOD-like receptors ...
  66. [66]
  67. [67]
    Convergent Evolution of Pathogen Effectors toward Reactive ...
    Sep 29, 2017 · Pathogen effectors may have evolved to converge on a common host protein network to suppress the common plant immune system, including the ROS burst and cell ...
  68. [68]
  69. [69]
    A bacterial effector mimics a host HSP90 client to undermine immunity
    A bacterial effector operates through a “betrayal-like” mechanism by masquerading as an HSP90 client as a means to achieve specificity for its target.
  70. [70]
    Induction and Signaling of an Apoptotic Response Typified by DNA ...
    In many plant−microbe interactions, recognition of an in- compatible pathogen triggers the hypersensitive response. (HR), resulting in rapid cell death ...
  71. [71]
    Hypersensitive response-related death - ResearchGate
    Aug 7, 2025 · The final, preeminent step of TE PCD is a rapid collapse of the vacuole occurring after completion of secondary cell wall synthesis. Vacuole ...
  72. [72]
    The Hypersensitive Response : A Case Of Cell Death Induction In ...
    One common feature of disease resistance is the rapid development of cell death at and immediately surrounding infection sites, called the Hypersensitive ...
  73. [73]
    The Plant Cell Wall: A Dynamic Barrier Against Pathogen Invasion
    This review summarizes recent advances in our understanding of cell wall-associated defenses induced by pathogen perception.
  74. [74]
    Mechanisms to Mitigate the Trade-Off between Growth and Defense
    The edr1-1 mutant constitutively expressed the priming phenotype and similarly incurred slight fitness costs, but these costs were substantially lower than the ...
  75. [75]
    (PDF) Mighty Dwarfs: Arabidopsis Autoimmune Mutants and Their ...
    the hypersensitive response in Arabidopsis dnd1 mutant. ... Autoimmune mutants have been instrumental in understanding the fine tuning of plant defense responses.
  76. [76]
    Differences in Intensity and Specificity of Hypersensitive Response ...
    To determine the significance of our observations, statistical analysis was performed on necrotic lesion size data. ... hypersensitive response and non-host ...
  77. [77]
    Programmed cell death (PCD) control in plants: New insights from ...
    Notable hallmarks of apoptosis in HR response include membrane blebbing, chromatin condensation, genomic DNA fragmentation, cell-wall alterations, ion fluxes ...
  78. [78]
    Systemic Acquired Resistance - PMC - PubMed Central - NIH
    Early grafting experiments have shown that a primary infected leaf of a plant can produce a systemic signal that is graft transmissible from rootstock to scion.
  79. [79]
    Systemic acquired resistance: turning local infection into ... - PubMed
    Consequently, the rest of the plant is protected from secondary infection for a period of weeks to months. SAR can even be passed on to progeny through ...
  80. [80]
    Long-distance communication and signal amplification in systemic ...
    Feb 21, 2013 · This review summarizes the involvement and interaction between long-distance SAR signals and details the recently discovered role of Pip in defense ...
  81. [81]
    Pipecolic Acid, an Endogenous Mediator of Defense Amplification ...
    Pipecolic acid (Pip) is a critical signal for plant immunity, acting as a regulator of defense amplification and priming, and is necessary for SAR.
  82. [82]
  83. [83]
    A pathogen-inducible endogenous siRNA in plant immunity - PNAS
    siRNA-mediated gene silencing plays an essential role in antiviral defense in both plant and animal systems (35, 36). However, these siRNAs generated from viral ...Missing: mobile piRNAs
  84. [84]
    Systemic Acquired Resistance and Salicylic Acid: Past, Present, and ...
    Jul 10, 2018 · Here, we present a historical overview of the progress that has been made to date in elucidating the role of SA in signaling plant immune responses.
  85. [85]
    Balancing Selection at the Tomato RCR3 Guardee Gene Family ...
    Multiple studies have reported high genetic diversity at R genes maintained by balancing selection.
  86. [86]
    Hybrid Incompatibility of the Plant Immune System - PubMed Central
    Hybridization is a core element in modern rice breeding as beneficial combinations of two parental genomes often result in the expression of heterosis.
  87. [87]
    The NLRomes of Zea mays NAM founder lines and Zea luxurians ...
    Mar 16, 2023 · The NLRomes of the maize NAM founder lines and Zea luxurians possess high sequence diversity, presence–absence variation, transchromosomal mobility.Missing: variants | Show results with:variants
  88. [88]
    A singleton NLR of recent origin causes hybrid necrosis in ... - bioRxiv
    May 20, 2020 · Hybrid necrosis in plants arises from conflict between divergent alleles of immunity genes contributed by different parents, resulting in ...<|control11|><|separator|>
  89. [89]
    Local adaptation of both plant and pathogen: an arms‐race ...
    Jun 22, 2025 · The balance between host and parasite coevolution is driven by differences in each species' population and quantitative genetic characteristics.
  90. [90]
    Genomic variability as a driver of plant–pathogen coevolution? - PMC
    Feb 1, 2014 · Allelic variation in many R-genes is likely to be ecologically and functionally relevant. Population genetic analyses in several plant species ...
  91. [91]
    Hypersensitive Response-Like Reaction Is Associated with Hybrid ...
    ... R genes is necessary to mediate hybrid incompatibility [2]. Hypersensitive reaction plays significant roles in plant pathogenic resistance, and a lot of the ...
  92. [92]
    A two-locus interaction causes interspecific hybrid weakness in rice
    Feb 21, 2014 · Hybrid weakness, the poorer development of hybrids compared with their parents, hinders gene exchange between different species at the ...
  93. [93]
    HWA1- and HWA2-Mediated Hybrid Weakness in Rice Involves Cell ...
    In rice, hybrid weakness has been reported to result from interactions between the HWI1 locus, which encodes the LRR-RLK gene (R gene), and the HWI2 locus, ...
  94. [94]
    Genomic introgression through interspecific hybridization ...
    Jan 30, 2019 · We envision that interspecific introgression serves as an important mechanism for counteracting the reduction of genetic diversity in domesticated crops.Introgression-Mediated Gene... · Introgression And Asymmetric... · Pairwise Ibd DetectionMissing: 2020 | Show results with:2020
  95. [95]
    Domestication Reduces Plant Immune Receptor Gene Repertoires ...
    This study compares the IRG repertoires of diverse crop plants and their respective wild relatives within a comparative genomics framework. The results show ...
  96. [96]
    The relevance of gene flow with wild relatives in ... - Journals
    Apr 15, 2020 · The widespread use of genomic tools has allowed for a deeper understanding of the genetics and the evolutionary dynamics of domestication.
  97. [97]
  98. [98]
  99. [99]
  100. [100]