首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
Programmed cell death (PCD) is a crucial process both for plant development and responses to biotic and abiotic stress. There is accumulating evidence that chloroplasts may play a central role during plant PCD as for mitochondria in animal cells, but it is still unclear whether they participate in PCD onset, execution, or both. To tackle this question, we have analyzed the contribution of chloroplast function to the cell death phenotype of the myoinositol phosphate synthase1 (mips1) mutant that forms spontaneous lesions in a light-dependent manner. We show that photosynthetically active chloroplasts are required for PCD to occur in mips1, but this process is independent of the redox state of the chloroplast. Systematic genetic analyses with retrograde signaling mutants reveal that 3′-phosphoadenosine 5′-phosphate, a chloroplast retrograde signal that modulates nuclear gene expression in response to stress, can inhibit cell death and compromises plant innate immunity via inhibition of the RNA-processing 5′-3′ exoribonucleases. Our results provide evidence for the role of chloroplast-derived signal and RNA metabolism in the control of cell death and biotic stress response.Programmed cell death (PCD) is a universal process in multicellular organisms, contributing to the controlled and active degradation of the cell. In plants, PCD is required for processes as diverse as development, self-incompatibility, and stress response. One well-documented example is the induction of PCD upon pathogen attack, allowing the confinement of the infection, and resistance of the plant. The signaling events leading to the onset of PCD have been extensively studied: pathogen recognition triggers activation of mitogen-activated protein kinase cascades, as well as production of reactive oxygen species (ROS) and salicylic acid (SA), which lead to a hypersensitive response (Coll et al., 2011).From a cellular point of view, several classes of plant PCD have been described and compared with the ones found in animal cells (van Doorn, 2011). PCD is thought to have evolved independently in plants and animals, and genes underlying these mechanisms are therefore poorly conserved between the two kingdoms. However, most cellular features are conserved between plant and animal PCD that are both characterized by cell shrinkage, chromatin condensation, DNA laddering, mitochondria permeabilization, and depolarization (Dickman and Fluhr, 2013). In animal cells, mitochondria play a central role in the regulation of apoptosis (Czabotar et al., 2014; Mariño et al., 2014), and this role is likely shared between the two kingdoms (Lord and Gunawardena, 2012). That said, additional mitochondria-independent PCD pathways have clearly evolved in plants.Genetic approaches have greatly contributed to our understanding of cellular pathways governing PCD in plants. For example, the isolation of lesion mimic mutants (LMMs), in which cell death occurs spontaneously, has allowed the identification of several negative regulators of cell death (for review, see Bruggeman et al., 2015b). Interestingly, lesion formation is light dependent in several of these mutants, which include one of the best characterized LMMs—lesions simulating disease1 (lsd1; Dietrich et al., 1994). The LSD1 protein is required for plant acclimation to excess excitation energy (Mateo et al., 2004): when plants are exposed to excessive amounts of light, the redox status of the plastoquinone pool in the chloroplastic electron transfer chain is thought to influence LSD1-dependent signaling to modulate cell death (Mühlenbock et al., 2008). Additionally, we have previously identified the myoinositol phosphate synthase1 (mips1) mutant as a LMM, in which lesion formation is also light dependent (Meng et al., 2009). This mutant is deficient in the myoinositol (MI) phosphate synthase that catalyzes the first committed step of MI biosynthesis and displays pleiotropic defects such as reduced root growth, abnormal vein development, and spontaneous cell death on leaves, together with severe growth reduction after lesions begin to develop (Meng et al., 2009; Donahue et al., 2010). The light-dependent PCD in the mips1 mutant, as observed for lsd1, suggests that chloroplasts may play a role in the MI-dependent cell death regulation. Accumulating evidence suggests that chloroplasts may play a central role in PCD regulation like mitochondria in animal cells (Wang and Bayles, 2013). First, as described in the case of lsd1, excess light energy received by the chloroplast can function as a trigger for PCD. Furthermore, singlet oxygen (1O2), a ROS, can activate the EXECUTER1 (EX1) and EX2 proteins in the chloroplasts to initiate PCD (Lee et al., 2007). Likewise, ROS generated by chloroplasts play a major role for PCD onset during nonhost interaction between tobacco (Nicotiana tabacum) and Xanthomonas campestris (Zurbriggen et al., 2009). Finally, functional chloroplasts have also been shown to be required for PCD in cell suspensions (Gutierrez et al., 2014) and in a number of LMMs (Mateo et al., 2004; Meng et al., 2009; Bruggeman et al., 2015b). Thus, chloroplasts are now recognized as important components of plant defense response against pathogens (Stael et al., 2015) and are proposed to function with mitochondria in the execution of PCD (Van Aken and Van Breusegem, 2015). However, the exact signaling and metabolic contribution of chloroplasts to PCD remain to be elucidated. Furthermore, cross talk between chloroplasts and mitochondria does occur, such as during photorespiration (Sunil et al., 2013), but whether such communication functions sequentially or in parallel in the control of PCD remains to be determined (Van Aken and Van Breusegem, 2015).To further investigate how chloroplasts contribute to the regulation of cell death, we performed both forward and reverse genetics on the mips1 mutant. An extragenic secondary mutation in divinyl protochlorophyllide 8-vinyl reductase involved in chlorophyll biosynthesis leads to chlorophyll deficiency that abolishes the mips1 cell death phenotype, as do changes in CO2 availability. These findings provide evidence for a link between photosynthetic activity and PCD induction in mips1. Additionally, we investigated the contribution of several retrograde signaling pathways (Chan et al., 2015) to the control of PCD in mips1. This process was independent of GENOMES UNCOUPLED (GUN) and EX signaling pathways, but we found that the SAL1-PAP_XRN retrograde signaling pathway inhibits cell death as well as basal defense reactions in Arabidopsis (Arabidopsis thaliana).  相似文献   

3.
4.
5.
A central component of the plant defense response to pathogens is the hypersensitive response (HR), a form of programmed cell death (PCD). Rapid and localized induction of HR PCD ensures that pathogen invasion is prevented. Autophagy has been implicated in the regulation of HR cell death, but the functional relationship between autophagy and HR PCD and the regulation of these processes during the plant immune response remain controversial. Here, we show that a small GTP-binding protein, RabG3b, plays a positive role in autophagy and promotes HR cell death in response to avirulent bacterial pathogens in Arabidopsis (Arabidopsis thaliana). Transgenic plants overexpressing a constitutively active RabG3b (RabG3bCA) displayed accelerated, unrestricted HR PCD within 1 d of infection, in contrast to the autophagy-defective atg5-1 mutant, which gradually developed chlorotic cell death through uninfected sites over several days. Microscopic analyses showed the accumulation of autophagic structures during HR cell death in RabG3bCA cells. Our results suggest that RabG3b contributes to HR cell death via the activation of autophagy, which plays a positive role in plant immunity-triggered HR PCD.In response to the constant attack by microbial pathogens, plants have developed defense mechanisms to protect themselves against harmful diseases caused by various pathogens. Plants primarily rely on two layers of innate immunity to cope with microbial pathogens (Jones and Dangl, 2006). The first layer of plant immunity, which is triggered by pathogen-associated molecular patterns (PAMPs) such as bacterial flagellin, lipopolysaccharides, and fungal chitin, is designated PAMP-triggered immunity (PTI; Boller and He, 2009). Because pathogens have evolved to overcome PTI, plants have developed a second layer of immunity, referred to as effector-triggered immunity (ETI; Dodds and Rathjen, 2010). ETI depends on specific interactions between plant Resistance proteins and pathogen effectors and is often associated with a form of programmed cell death (PCD) termed the hypersensitive response (HR), which inhibits pathogen growth (Coll et al., 2011).Plants use PCD to regulate developmental and defense responses. In addition to pathogen attack, many abiotic stress factors such as heat and ozone exposure elicit PCD in plants (Hayward and Dinesh-Kumar, 2011). PCD also occurs during various developmental processes, including endosperm development, tracheary element (TE) differentiation, female gametophyte differentiation, leaf abscission, and senescence (Kuriyama and Fukuda, 2002; Gunawardena, 2008). Recently, plant PCD has been classified into two types, “autolytic” PCD and “nonautolytic” PCD, on the basis of the presence or absence of rapid cytoplasm clearance after tonoplast rupture, respectively (van Doorn et al., 2011). Autolytic PCD, which mainly occurs during plant development, falls under “autophagic” PCD in animals because it is associated with the accumulation of autophagy-related structures in the cytoplasm. Some forms of HR PCD classified as nonautolytic PCD in plants are accompanied by increased vacuolization, indicating the progress of autophagy, and therefore can be placed under autophagic PCD (Hara-Nishimura et al., 2005; Hatsugai et al., 2009).Autophagy is an intracellular process in which double membrane-bound autophagosomes enclose cytoplasmic components and damaged or toxic materials and target them to the vacuole or lysosome for degradation (Chung, 2011). In plants, autophagy plays important roles in the responses to nutrient starvation, senescence, and abiotic and biotic stresses (Liu et al., 2005; Xiong et al., 2005, 2007; Bassham, 2007; Hofius et al., 2009). Accumulating evidence indicates that autophagy regulates immune responses in both animals and plants. Autophagy is essential for the direct elimination of pathogens in mammalian systems (Levine et al., 2011). Invading bacteria and viruses are targeted to autophagosomes and then delivered to the lysosome for degradation in a process called xenophagy (Levine, 2005). In addition to its function in directly killing pathogens, xenophagic degradation can provide microbial antigens for major histocompatibility complex class II presentation to the innate and adaptive immune systems (Levine, 2005; Schmid and Münz, 2007). Furthermore, the human surface receptor CD46 was shown to directly induce autophagy through physical interaction with the autophagic machinery (Joubert et al., 2009). The role of autophagy in plant basal immunity to virulent pathogens has been determined (Patel and Dinesh-Kumar, 2008; Hofius et al., 2009; Lai et al., 2011; Lenz et al., 2011). Arabidopsis (Arabidopsis thaliana) plants defective in AUTOPHAGY-RELATED (ATG) genes exhibited enhanced susceptibility to the necrotrophic fungal pathogens Botrytis cinerea and Alternaria brassicicola, suggesting that the massive breakdown of cytoplasmic materials provides nutrients for the growth of necrotrophic pathogens or that fungal toxin-induced necrotic cell death is enhanced in atg mutants (Lai et al., 2011; Lenz et al., 2011). However, studies on the responses to the biotrophic pathogen Pseudomonas syringae pv tomato DC3000 (Pst DC3000) have yielded contradictory results. Whereas earlier studies reported that bacterial numbers significantly increased in ATG6-antisense (AS) and atg mutant plants (Patel and Dinesh-Kumar, 2008; Hofius et al., 2009), a recent study indicated that atg mutants exhibit increased resistance to Pst DC3000 (Lenz et al., 2011). Although these discrepancies remain to be resolved, salicylic acid (SA) levels and SA-dependent gene expression were both elevated in atg mutants, suggesting that autophagy may negatively regulate SA-associated plant immunity (Yoshimoto et al., 2009; Lenz et al., 2011). These findings indicate that the role of autophagy in plant immunity depends on the lifestyle of the invading pathogens (Lenz et al., 2011).Autophagy plays an important role in the regulation of HR PCD in plant innate immunity (Hayward and Dinesh-Kumar, 2011). Tobacco (Nicotiana tabacum) plants silenced for ATG6/Beclin1 and other ATG genes such as phosphatidylinositol 3-kinase (PI3K)/vacuolar protein sorting34 (VPS34), ATG3, and ATG7 underwent unrestricted HR PCD upon pathogen infection (Liu et al., 2005). ATG6-AS and atg5 mutant Arabidopsis plants also displayed unlimited HR PCD upon infection with the avirulent bacterium Pst DC3000 (AvrRpm1; Patel and Dinesh-Kumar, 2008; Yoshimoto et al., 2009). These studies suggest that autophagy is a “prosurvival” or “antideath” mechanism that negatively regulates HR PCD (Liu and Bassham, 2012). By contrast, a “prodeath” role has been suggested for autophagy in HR PCD regulation (Hofius et al., 2009). Pst DC3000 (AvrRps4)-induced and, to a lesser extent, Pst DC3000 (AvrRpm1)-induced HR PCD was suppressed in atg mutants, suggesting that autophagy plays a positive role and that autophagic cell death is involved in RPS4- and RPM1-mediated HR cell death.We previously showed that the small GTP-binding protein RabG3b, isolated from secretome analysis in Arabidopsis (Oh et al., 2005), functions as a component of autophagy and positively regulates TE differentiation via the activation of autophagic cell death (Kwon et al., 2010a, 2010b). Overexpression of a constitutively active RabG3b (RabG3bCA) in plants significantly increased autophagy during PCD associated with TE differentiation, thereby enhancing TE formation and xylem development. Transgenic poplar (Populus alba × Populus tremula var glandulosa) overexpressing Arabidopsis RabG3bCA was further generated, and these exhibited significant stimulation of xylem development together with autophagic activation, suggesting that RabG3b is a positive regulator of autophagy and xylem development in Populus spp. as well as Arabidopsis (Kwon et al., 2011). We also reported that RabG3b is involved in cell death associated with the fungal pathogen A. brassicicola and infection with the fungal toxin fumonisin B1 (FB1) as well as leaf senescence (Kwon et al., 2009). Here, we extend our work to determine the role of RabG3b and autophagy in immunity-associated HR PCD. We found that RabG3bCA transgenic plants accumulated a large number of autophagic structures and displayed accelerated, expanded cell death against a number of PCD inducers, such as FB1 and the bacterial pathogens Pst DC3000 (AvrRpm1) and Pst DC3000 (AvrRpt2). Our results suggest that RabG3b plays a positive role in immunity-associated HR PCD via the activation of autophagic cell death.  相似文献   

6.
7.
Sphingolipids are emerging as second messengers in programmed cell death and plant defense mechanisms. However, their role in plant defense is far from being understood, especially against necrotrophic pathogens. Sphingolipidomics and plant defense responses during pathogenic infection were evaluated in the mutant of long-chain base phosphate (LCB-P) lyase, encoded by the dihydrosphingosine-1-phosphate lyase1 (AtDPL1) gene and regulating long-chain base/LCB-P homeostasis. Atdpl1 mutants exhibit tolerance to the necrotrophic fungus Botrytis cinerea but susceptibility to the hemibiotrophic bacterium Pseudomonas syringae pv tomato (Pst). Here, a direct comparison of sphingolipid profiles in Arabidopsis (Arabidopsis thaliana) during infection with pathogens differing in lifestyles is described. In contrast to long-chain bases (dihydrosphingosine [d18:0] and 4,8-sphingadienine [d18:2]), hydroxyceramide and LCB-P (phytosphingosine-1-phosphate [t18:0-P] and 4-hydroxy-8-sphingenine-1-phosphate [t18:1-P]) levels are higher in Atdpl1-1 than in wild-type plants in response to B. cinerea. Following Pst infection, t18:0-P accumulates more strongly in Atdpl1-1 than in wild-type plants. Moreover, d18:0 and t18:0-P appear as key players in Pst- and B. cinerea-induced cell death and reactive oxygen species accumulation. Salicylic acid levels are similar in both types of plants, independent of the pathogen. In addition, salicylic acid-dependent gene expression is similar in both types of B. cinerea-infected plants but is repressed in Atdpl1-1 after treatment with Pst. Infection with both pathogens triggers higher jasmonic acid, jasmonoyl-isoleucine accumulation, and jasmonic acid-dependent gene expression in Atdpl1-1 mutants. Our results demonstrate that sphingolipids play an important role in plant defense, especially toward necrotrophic pathogens, and highlight a novel connection between the jasmonate signaling pathway, cell death, and sphingolipids.Plants have evolved a complex array of defenses when attacked by microbial pathogens. The success of plant resistance first relies on the capacity of the plant to recognize its invader. Among early events, a transient production of reactive oxygen species (ROS), known as the oxidative burst, is characteristic of successful pathogen recognition (Torres, 2010). Perception of pathogen attack then initiates a large array of immune responses, including modification of cell walls, as well as the production of antimicrobial proteins and metabolites like pathogenesis-related (PR) proteins and phytoalexins, respectively (Schwessinger and Ronald, 2012). The plant hormones salicylic acid (SA), jasmonic acid (JA), and ethylene (ET) are key players in the signaling networks involved in plant resistance (Bari and Jones, 2009; Tsuda and Katagiri, 2010; Robert-Seilaniantz et al., 2011). Interactions between these signal molecules allow the plant to activate and/or modulate an appropriate array of defense responses, depending on the pathogen lifestyle, necrotroph or biotroph (Glazebrook, 2005; Koornneef and Pieterse, 2008). Whereas SA is considered essential for resistance to (hemi)biotrophic pathogens, it is assumed that JA and ET signaling pathways are important for resistance to necrotrophic pathogens in Arabidopsis (Arabidopsis thaliana; Thomma et al., 2001; Glazebrook, 2005). A successful innate immune response often includes the so-called hypersensitive response (HR), a form of rapid programmed cell death (PCD) occurring in a limited area at the site of infection. This suicide of infected cells is thought to limit the spread of biotrophic pathogens, including viruses, bacteria, fungi, and oomycetes (Mur et al., 2008).During the past decade, significant progress has been made in our understanding of the cellular function of plant sphingolipids. Besides being structural components of cell membranes, sphingolipids are bioactive metabolites that regulate important cellular processes such as cell survival and PCD, occurring during either plant development or plant defense (Dunn et al., 2004; Berkey et al., 2012; Markham et al., 2013). The first evidence of the role of sphingolipids in these processes came from the use of the fungal toxins fumonisin B1 (FB1) and AAL, produced by the necrotrophic agent Alternaria alternata f. sp. lycopersici. These toxins are structural sphingosine (d18:1) analogs and function as ceramide synthase inhibitors. They triggered PCD when exogenously applied to plants. Mutant strains in which the production of such toxins is abrogated failed to infect the host plant, implying that toxin accumulation is required for pathogenicity and that the induction of plant PCD could be considered a virulence tool used by necrotrophic pathogens (Berkey et al., 2012). Moreover, several studies revealed that ceramides (Cers) and long-chain bases (LCBs) are also potent inducers of PCD in plants. For example, exogenously applied Cers and LCBs (d18:0, d18:1, or t18:0) induced PCD either in cell suspension cultures (Liang et al., 2003; Lachaud et al., 2010, 2011; Alden et al., 2011) or in whole seedlings (Shi et al., 2007; Takahashi et al., 2009; Saucedo-García et al., 2011). AAL- and FB1-induced PCD seemed to be due to the accumulation of free sphingoid bases (dihydrosphingosine [d18:0] and phytosphingosine [t18:0]; Abbas et al., 1994; Brandwagt et al., 2000; Shi et al., 2007). Spontaneous cell death in lag one homolog1 or l-myoinositol1-phosphate synthase mutant could be due to trihydroxy-LCB and/or Cer accumulation (Donahue et al., 2010; Ternes et al., 2011). Deciphering of Cer participation in the induction of HR and associated PCD also came from studies on accelerated cell death5 (acd5) and enhancing resistance to powdery mildew8 (RPW8)-mediated hypersensitive response (erh1) mutants, which displayed overaccumulation of Cers. These mutants exhibited spontaneous cell death and resistance to biotrophic pathogens, which seemed to be linked with SA and PR protein accumulation (Liang et al., 2003; Wang et al., 2008).Altogether, these data provide evidence of a link between PCD, defense, and sphingolipid metabolism. However, the fatty acid hydroxylase1/2 (atfah1/atfah2) double mutant that accumulates SA and Cers was more tolerant to the obligate biotrophic fungus Golovinomyces cichoracearum but did not display a PCD-like phenotype, suggesting that Cers alone are not involved in the induction of PCD (König et al., 2012). Moreover, Saucedo-García et al. (2011) postulated that dihydroxy-LCBs, but not trihydroxy-LCBs, might be primary mediators for LCB-induced PCD. The sphingoid base hydroxylase sbh1/sbh2 double mutant completely lacking trihydroxy-LCBs showed enhanced expression of PCD marker genes (Chen et al., 2008). On the contrary, increase in t18:0 was specifically sustained in plant interaction with the avirulent Pseudomonas syringae pv tomato (Pst) strain and correlated with a strong PCD induction in leaves (Peer et al., 2010). Thus, the nature of sphingolipids able to induce PCD is still under debate and may evolve depending on plants and their environment. The phosphorylated form of LCBs (LCB-Ps) could abrogate PCD induced by LCBs, Cers, or heat stress in a dose-dependent manner (Shi et al., 2007; Alden et al., 2011). Furthermore, blocking the conversion of LCBs to LCB-Ps by using specific inhibitors induced PCD in cell suspension culture (Alden et al., 2011). Recently, overexpression of rice (Oryza sativa) LCB kinase in transgenic tobacco (Nicotiana tabacum) plants reduced PCD after treatment with FB1 (Zhang et al., 2013). Genetic mutation on LCB-P lyase encoded by the AtDPL1 gene, modifying the LCB-LCB-P ratio, could impact PCD levels after treatment with FB1 (Tsegaye et al., 2007). Altogether, these data point to the existence of a rheostat between LCBs and their phosphorylated forms that controls plant cell fate toward cell death or survival.Data on plant sphingolipid functions are still fragmentary. Only a few reports have described interconnections between sphingolipids, cell death, and plant defense responses, almost exclusively in response to (hemi)biotrophic pathogens. Knowledge about such relations in response to necrotrophic pathogens is still in its infancy (Rivas-San Vicente et al., 2013; Bi et al., 2014). In this report, the link between sphingolipids, cell death, and plant defense has been explored in response to Botrytis cinerea infection and in comparison with Pst infection. For this purpose, Atdpl1 mutant plants, disturbed in LCB/LCB-P accumulation without displaying any phenotype under standard growth conditions (Tsegaye et al., 2007), have been analyzed after pathogen infection. Our results revealed that modification of sphingolipid contents not only impacted plant tolerance to hemibiotrophs but also greatly affected resistance to necrotrophs. Whereas the SA signaling pathway is globally repressed in Atdpl1-1 compared with wild-type plants, the JA signaling pathway is significantly enhanced. Cell death and ROS accumulation are markedly modified in Atdpl1-1 mutant plants. We further demonstrated that phytosphingosine-1-phosphate (t18:0-P) and d18:0 are key players in pathogen-induced cell death and ROS generation. Here, we thus established a link between JA signaling, PCD, and sphingolipid metabolism.  相似文献   

8.
In Arabidopsis thaliana, responses to pathogen-associated molecular patterns (PAMPs) are mediated by cell surface pattern recognition receptors (PRRs) and include the accumulation of reactive oxygen species, callose deposition in the cell wall, and the generation of the signal molecule salicylic acid (SA). SA acts in a positive feedback loop with ACCELERATED CELL DEATH6 (ACD6), a membrane protein that contributes to immunity. This work shows that PRRs associate with and are part of the ACD6/SA feedback loop. ACD6 positively regulates the abundance of several PRRs and affects the responsiveness of plants to two PAMPs. SA accumulation also causes increased levels of PRRs and potentiates the responsiveness of plants to PAMPs. Finally, SA induces PRR- and ACD6-dependent signaling to induce callose deposition independent of the presence of PAMPs. This PAMP-independent effect of SA causes a transient reduction of PRRs and ACD6-dependent reduced responsiveness to PAMPs. Thus, SA has a dynamic effect on the regulation and function of PRRs. Within a few hours, SA signaling promotes defenses and downregulates PRRs, whereas later (within 24 to 48 h) SA signaling upregulates PRRs, and plants are rendered more responsive to PAMPs. These results implicate multiple modes of signaling for PRRs in response to PAMPs and SA.  相似文献   

9.
10.
11.
12.
13.
Salicylic acid (SA) signaling acts in defense and plant development. The only gene demonstrated to be required for the response to SA is Arabidopsis thaliana NON-EXPRESSER OF PATHOGENESIS-RELATED GENE 1 (NPR1), and npr1 mutants are insensitive to SA. By focusing on the effect of analogs of SA on plant development, we identified mutants in additional genes acting in the SA response. In this work, we describe a gene necessary for the SA Non-Recognition-of-BTH4 (NRB4). Three nrb4 alleles recovered from the screen cause phenotypes similar to the wild type in the tested conditions, except for SA-related phenotypes. Plants with NRB4 null alleles express profound insensitivity to SA, even more than npr1. NRB4 null mutants are also sterile and their growth is compromised. Plants carrying weaker nrb4 alleles are also insensitive to SA, with some quantitative differences in some phenotypes, like systemic acquired resistance or pathogen growth restriction. When weak alleles are used, NPR1 and NRB4 mutations produce an additive phenotype, but we did not find evidence of a genetic interaction in F1 nor biochemical interaction in yeast or in planta. NRB4 is predicted to be a subunit of Mediator, the ortholog of MED15 in Arabidopsis. Mechanistically, NRB4 functions downstream of NPR1 to regulate the SA response.  相似文献   

14.
15.
16.
17.
18.
Plant activators are compounds, such as analogs of the defense hormone salicylic acid (SA), that protect plants from pathogens by activating the plant immune system. Although some plant activators have been widely used in agriculture, the molecular mechanisms of immune induction are largely unknown. Using a newly established high-throughput screening procedure that screens for compounds that specifically potentiate pathogen-activated cell death in Arabidopsis thaliana cultured suspension cells, we identified five compounds that prime the immune response. These compounds enhanced disease resistance against pathogenic Pseudomonas bacteria in Arabidopsis plants. Pretreatments increased the accumulation of endogenous SA, but reduced its metabolite, SA-O-β-d-glucoside. Inducing compounds inhibited two SA glucosyltransferases (SAGTs) in vitro. Double knockout plants that lack both SAGTs consistently exhibited enhanced disease resistance. Our results demonstrate that manipulation of the active free SA pool via SA-inactivating enzymes can be a useful strategy for fortifying plant disease resistance and may identify useful crop protectants.  相似文献   

19.
20.
Nitric oxide (NO) is a small redox molecule that acts as a signal in different physiological and stress-related processes in plants. Recent evidence suggests that the biological activity of NO is also mediated by S-nitrosylation, a well-known redox-based posttranslational protein modification. Here, we show that during programmed cell death (PCD), induced by both heat shock (HS) or hydrogen peroxide (H2O2) in tobacco (Nicotiana tabacum) Bright Yellow-2 cells, an increase in S-nitrosylating agents occurred. NO increased in both experimentally induced PCDs, although with different intensities. In H2O2-treated cells, the increase in NO was lower than in cells exposed to HS. However, a simultaneous increase in S-nitrosoglutathione (GSNO), another NO source for S-nitrosylation, occurred in H2O2-treated cells, while a decrease in this metabolite was evident after HS. Consistently, different levels of activity and expression of GSNO reductase, the enzyme responsible for GSNO removal, were found in cells subjected to the two different PCD-inducing stimuli: low in H2O2-treated cells and high in the heat-shocked ones. Irrespective of the type of S-nitrosylating agent, S-nitrosylated proteins formed upon exposure to both of the PCD-inducing stimuli. Interestingly, cytosolic ascorbate peroxidase (cAPX), a key enzyme controlling H2O2 levels in plants, was found to be S-nitrosylated at the onset of both PCDs. In vivo and in vitro experiments showed that S-nitrosylation of cAPX was responsible for the rapid decrease in its activity. The possibility that S-nitrosylation induces cAPX ubiquitination and degradation and acts as part of the signaling pathway leading to PCD is discussed.Nitric oxide (NO) is a gaseous and diffusible redox molecule that acts as a signaling compound in both animal and plant systems (Pacher et al., 2007; Besson-Bard et al., 2008). In plants, NO has been found to play a key role in several physiological processes, such as germination, lateral root development, flowering, senescence, stomatal closure, and growth of pollen tubes (Beligni and Lamattina, 2000; Neill et al., 2002; Correa-Aragunde et al., 2004; He et al., 2004; Prado et al., 2004; Carimi et al., 2005). In addition, NO has been reported to be involved in plant responses to both biotic and abiotic stresses (Leitner et al., 2009; Siddiqui et al., 2011) and in the signaling pathways leading to programmed cell death (PCD; Delledonne et al., 1998; de Pinto et al., 2006; De Michele et al., 2009; Lin et al., 2012; Serrano et al., 2012).The cellular environment may greatly influence the chemical reactivity of NO, giving rise to different biologically active NO-derived compounds, collectively named reactive nitrogen species, which amplify and differentiate its ability to activate physiological and stress-related processes. Many of the biological properties of NO are due to its high affinity with transition metals of metalloproteins as well as its reactivity with reactive oxygen species (ROS; Hill et al., 2010). However, recent evidence suggests that protein S-nitrosylation, due to the addition of NO to reactive Cys thiols, may act as a key mechanism of NO signaling in plants (Wang et al., 2006; Astier et al., 2011). NO is also able to react with reduced glutathione (GSH), the most abundant cellular thiol, thus producing S-nitrosoglutathione (GSNO), which also acts as an endogenous trans-nitrosylating agent. GSNO is also considered as a NO store and donor and, as it is more stable than NO, acts as a long-distance NO transporter through the floematic flux (Malik et al., 2011). S-Nitrosoglutathione reductase (GSNOR), which is an enzyme conserved from bacteria to humans, has been suggested to play a role in regulating S-nitrosothiols (SNO) and the turnover of S-nitrosylated proteins in plants (Liu et al., 2001; Rusterucci et al., 2007).A number of proteins involved in metabolism, stress responses, and redox homeostasis have been identified as potential targets for S-nitrosylation in Arabidopsis (Arabidopsis thaliana; Lindermayr et al., 2005). During the hypersensitive response (HR), 16 proteins were identified to be S-nitrosylated in the seedlings of the same species (Romero-Puertas et al., 2008); in Citrus species, S-nitrosylation of about 50 proteins occurred in the NO-mediated resistance to high salinity (Tanou et al., 2009).However, while the number of candidate proteins for S-nitrosylation is increasing, the functional significance of protein S-nitrosylation has been explained only in a few cases, such as for nonsymbiotic hemoglobin (Perazzolli et al., 2004), glyceraldehyde 3-phosphate dehydrogenase (Lindermayr et al., 2005; Wawer et al., 2010), Met adenosyltransferase (Lindermayr et al., 2006), and metacaspase9 (Belenghi et al., 2007). Of particular interest are the cases in which S-nitrosylation involves enzymes controlling ROS homeostasis. For instance, it has been reported that S-nitrosylation of peroxiredoxin IIE regulates the antioxidant function of this enzyme and might contribute to the HR (Romero-Puertas et al., 2007). It has also been shown that in the immunity response, S-nitrosylation of NADPH oxidase inactivates the enzyme, thus reducing ROS production and controlling HR development (Yun et al., 2011).Recently, S-nitrosylation has also been shown to be involved in PCD of nitric oxide excess1 (noe1) rice (Oryza sativa) plants, which are mutated in the OsCATC gene coding for catalase (Lin et al., 2012). In these plants, which show PCD-like phenotypes under high-light conditions, glyceraldehyde 3-phosphate dehydrogenase and thioredoxin are S-nitrosylated. This suggests that the NO-dependent regulation of these proteins is involved in plant PCD, similar to what occurs in animal apoptosis (Sumbayev, 2003; Hara et al., 2005; Lin et al., 2012). The increase in hydrogen peroxide (H2O2) after exposure to high light in noe1 plants is responsible for the production of NO required for leaf cell death induction (Lin et al., 2012). There is a strict relationship between H2O2 and NO in PCD activation (Delledonne et al., 2001; de Pinto et al., 2002); however, the mechanism of this interplay is largely still unknown (for review, see Zaninotto et al., 2006; Zhao, 2007; Yoshioka et al., 2011). NO can induce ROS production and vice versa, and their reciprocal modulation in terms of intensity and timing seems to be crucial in determining PCD activation and in controlling HR development (Delledonne et al., 2001; Zhao, 2007; Yun et al., 2011).In previous papers, we demonstrated that heat shock (HS) at 55°C and treatment with 50 mm H2O2 promote PCD in tobacco (Nicotiana tabacum) Bright Yellow-2 (BY-2) cells (Vacca et al., 2004; de Pinto et al., 2006; Locato et al., 2008). In both experimental conditions, NO production and decrease in cytosolic ascorbate peroxidase (cAPX) were observed as early events in the PCD pathway, and cAPX decrease has been suggested to contribute to determining the redox environment required for PCD (de Pinto et al., 2006; Locato et al., 2008).In this study, the production of nitrosylating agents (NO and GSNO) in the first hours of PCD induction by HS or H2O2 treatment in tobacco BY-2 cells and their role in PCD were studied. The possibility that S-nitrosylation could be a first step in regulating cAPX activity and turnover as part of the signaling pathway leading to PCD was also investigated.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号