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.
3.
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   

4.
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   

5.
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   

6.
7.
8.
Fumarylacetoacetate hydrolase (FAH) hydrolyzes fumarylacetoacetate to fumarate and acetoacetate, the final step in the tyrosine (Tyr) degradation pathway that is essential to animals. Deficiency of FAH in animals results in an inborn lethal disorder. However, the role for the Tyr degradation pathway in plants remains to be elucidated. In this study, we isolated an Arabidopsis (Arabidopsis thaliana) short-day sensitive cell death1 (sscd1) mutant that displays a spontaneous cell death phenotype under short-day conditions. The SSCD1 gene was cloned via a map-based cloning approach and found to encode an Arabidopsis putative FAH. The spontaneous cell death phenotype of the sscd1 mutant was completely eliminated by further knockout of the gene encoding the putative homogentisate dioxygenase, which catalyzes homogentisate into maleylacetoacetate (the antepenultimate step) in the Tyr degradation pathway. Furthermore, treatment of Arabidopsis wild-type seedlings with succinylacetone, an abnormal metabolite caused by loss of FAH in the Tyr degradation pathway, mimicked the sscd1 cell death phenotype. These results demonstrate that disruption of FAH leads to cell death in Arabidopsis and suggest that the Tyr degradation pathway is essential for plant survival under short-day conditions.Programmed cell death (PCD) has been defined as a sequence of genetically regulated events that lead to the elimination of specific cells, tissues, or whole organs (Lockshin and Zakeri, 2004). In plants, PCD is essential for developmental processes and defense responses (Dangl et al., 1996; Greenberg, 1996; Durrant et al., 2007). One well-characterized example of plant PCD is the hypersensitive response occurring during incompatible plant-pathogen interactions (Lam, 2004), which results in cell death to form visible lesions at the site of infection by an avirulent pathogen and consequently limits the pathogen spread (Morel and Dangl, 1997).To date, a large number of mutants that display spontaneous cell death lesions have been identified in barley (Hordeum vulgare), maize (Zea mays), rice (Oryza sativa), and Arabidopsis (Arabidopsis thaliana; Marchetti et al., 1983; Wolter et al., 1993; Dietrich et al., 1994; Gray et al., 1997). Because lesions form in the absence of pathogen infection, these mutants have been collectively termed as lesion-mimic mutants. Many genes with regulatory roles in PCD and defense responses, including LESION SIMULATING DISEASE1, ACCELERATED CELL DEATH11, and VASCULAR ASSOCIATED DEATH1, have been cloned and characterized (Dietrich et al., 1997; Brodersen et al., 2002; Lorrain et al., 2004).The appearance of spontaneous cell death lesions in some lesion-mimic mutants is dependent on photoperiod. For example, the Arabidopsis mutant lesion simulating disease1 and myoinositol-1-phosphate synthase1 show lesions under long days (LD; Dietrich et al., 1994; Meng et al., 2009), whereas the lesion simulating disease2, lesion initiation1, enhancing RPW8-mediated HR-like cell death1, and lag one homolog1 display lesions under short days (SD; Dietrich et al., 1994; Ishikawa et al., 2003; Wang et al., 2008; Ternes et al., 2011).Blockage of some metabolic pathways in plants may cause cell death and result in lesion formation. For example, the lesion-mimic phenotypes in the Arabidopsis mutants lesion initiation2 and accelerated cell death2 and the maize mutant lesion mimic22 result from an impairment of porphyrin metabolism (Hu et al., 1998; Ishikawa et al., 2001; Mach et al., 2001). Deficiency in fatty acid, sphingolipid, and myoinositol metabolism also causes cell death in Arabidopsis (Mou et al., 2000; Liang et al., 2003; Wang et al., 2008; Meng et al., 2009; Donahue et al., 2010; Berkey et al., 2012).Tyr degradation is an essential five-step pathway in animals (Lindblad et al., 1977). First, Tyr aminotransferase catalyzes the conversion of Tyr into 4-hydroxyphenylpyruvate, which is further transformed into homogentisate by 4-hydroxyphenylpyruvate dioxygenase. Through the sequential action of homogentisate dioxygenase (HGO), maleylacetoacetate isomerase (MAAI), and fumarylacetoacetate hydrolase (FAH), homogentisate is catalyzed to generate fumarate and acetoacetate (Lindblad et al., 1977). Blockage of this pathway in animals results in metabolic disorder diseases (Lindblad et al., 1977; Ruppert et al., 1992; Grompe et al., 1993). For example, human FAH deficiency causes hereditary tyrosinemia type I (HT1), an inborn lethal disease (St-Louis and Tanguay, 1997). Although the homologous genes putatively encoding these enzymes exist in plants (Dixon et al., 2000; Lopukhina et al., 2001; Dixon and Edwards, 2006), it is unclear whether this pathway is essential for plant growth and development.In this study, we report the isolation and characterization of a recessive short-day sensitive cell death1 (sscd1) mutant in Arabidopsis. Map-based cloning of the corresponding gene revealed that SSCD1 encodes the Arabidopsis putative FAH. Further knockout of the gene encoding the Arabidopsis putative HGO completely eliminated the spontaneous cell death phenotype in the sscd1 mutant. Furthermore, we found that treatment of Arabidopsis wild-type seedlings with succinylacetone, an abnormal metabolite caused by loss of FAH in the Tyr degradation pathway (Lindblad et al., 1977), is able to mimic the sscd1 cell death phenotype. These results demonstrate that disruption of FAH leads to cell death in Arabidopsis and suggest that the Tyr degradation pathway is essential for plant survival under SD.  相似文献   

9.
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   

10.
Necrotrophic and biotrophic pathogens are resisted by different plant defenses. While necrotrophic pathogens are sensitive to jasmonic acid (JA)-dependent resistance, biotrophic pathogens are resisted by salicylic acid (SA)- and reactive oxygen species (ROS)-dependent resistance. Although many pathogens switch from biotrophy to necrotrophy during infection, little is known about the signals triggering this transition. This study is based on the observation that the early colonization pattern and symptom development by the ascomycete pathogen Plectosphaerella cucumerina (P. cucumerina) vary between inoculation methods. Using the Arabidopsis (Arabidopsis thaliana) defense response as a proxy for infection strategy, we examined whether P. cucumerina alternates between hemibiotrophic and necrotrophic lifestyles, depending on initial spore density and distribution on the leaf surface. Untargeted metabolome analysis revealed profound differences in metabolic defense signatures upon different inoculation methods. Quantification of JA and SA, marker gene expression, and cell death confirmed that infection from high spore densities activates JA-dependent defenses with excessive cell death, while infection from low spore densities induces SA-dependent defenses with lower levels of cell death. Phenotyping of Arabidopsis mutants in JA, SA, and ROS signaling confirmed that P. cucumerina is differentially resisted by JA- and SA/ROS-dependent defenses, depending on initial spore density and distribution on the leaf. Furthermore, in situ staining for early callose deposition at the infection sites revealed that necrotrophy by P. cucumerina is associated with elevated host defense. We conclude that P. cucumerina adapts to early-acting plant defenses by switching from a hemibiotrophic to a necrotrophic infection program, thereby gaining an advantage of immunity-related cell death in the host.Plant pathogens are often classified as necrotrophic or biotrophic, depending on their infection strategy (Glazebrook, 2005; Nishimura and Dangl, 2010). Necrotrophic pathogens kill living host cells and use the decayed plant tissue as a substrate to colonize the plant, whereas biotrophic pathogens parasitize living plant cells by employing effector molecules that suppress the host immune system (Pel and Pieterse, 2013). Despite this binary classification, the majority of pathogenic microbes employ a hemibiotrophic infection strategy, which is characterized by an initial biotrophic phase followed by a necrotrophic infection strategy at later stages of infection (Perfect and Green, 2001). The pathogenic fungi Magnaporthe grisea, Sclerotinia sclerotiorum, and Mycosphaerella graminicola, the oomycete Phytophthora infestans, and the bacterial pathogen Pseudomonas syringae are examples of hemibiotrophic plant pathogens (Perfect and Green, 2001; Koeck et al., 2011; van Kan et al., 2014; Kabbage et al., 2015).Despite considerable progress in our understanding of plant resistance to necrotrophic and biotrophic pathogens (Glazebrook, 2005; Mengiste, 2012; Lai and Mengiste, 2013), recent debate highlights the dynamic and complex interplay between plant-pathogenic microbes and their hosts, which is raising concerns about the use of infection strategies as a static tool to classify plant pathogens. For instance, the fungal genus Botrytis is often labeled as an archetypal necrotroph, even though there is evidence that it can behave as an endophytic fungus with a biotrophic lifestyle (van Kan et al., 2014). The rice blast fungus Magnaporthe oryzae, which is often classified as a hemibiotrophic leaf pathogen (Perfect and Green, 2001; Koeck et al., 2011), can adopt a purely biotrophic lifestyle when infecting root tissues (Marcel et al., 2010). It remains unclear which signals are responsible for the switch from biotrophy to necrotrophy and whether these signals rely solely on the physiological state of the pathogen, or whether host-derived signals play a role as well (Kabbage et al., 2015).The plant hormones salicylic acid (SA) and jasmonic acid (JA) play a central role in the activation of plant defenses (Glazebrook, 2005; Pieterse et al., 2009, 2012). The first evidence that biotrophic and necrotrophic pathogens are resisted by different immune responses came from Thomma et al. (1998), who demonstrated that Arabidopsis (Arabidopsis thaliana) genotypes impaired in SA signaling show enhanced susceptibility to the biotrophic pathogen Hyaloperonospora arabidopsidis (formerly known as Peronospora parastitica), while JA-insensitive genotypes were more susceptible to the necrotrophic fungus Alternaria brassicicola. In subsequent years, the differential effectiveness of SA- and JA-dependent defense mechanisms has been confirmed in different plant-pathogen interactions, while additional plant hormones, such as ethylene, abscisic acid (ABA), auxins, and cytokinins, have emerged as regulators of SA- and JA-dependent defenses (Bari and Jones, 2009; Cao et al., 2011; Pieterse et al., 2012). Moreover, SA- and JA-dependent defense pathways have been shown to act antagonistically on each other, which allows plants to prioritize an appropriate defense response to attack by biotrophic pathogens, necrotrophic pathogens, or herbivores (Koornneef and Pieterse, 2008; Pieterse et al., 2009; Verhage et al., 2010).In addition to plant hormones, reactive oxygen species (ROS) play an important regulatory role in plant defenses (Torres et al., 2006; Lehmann et al., 2015). Within minutes after the perception of pathogen-associated molecular patterns, NADPH oxidases and apoplastic peroxidases generate early ROS bursts (Torres et al., 2002; Daudi et al., 2012; O’Brien et al., 2012), which activate downstream defense signaling cascades (Apel and Hirt, 2004; Torres et al., 2006; Miller et al., 2009; Mittler et al., 2011; Lehmann et al., 2015). ROS play an important regulatory role in the deposition of callose (Luna et al., 2011; Pastor et al., 2013) and can also stimulate SA-dependent defenses (Chaouch et al., 2010; Yun and Chen, 2011; Wang et al., 2014; Mammarella et al., 2015). However, the spread of SA-induced apoptosis during hyperstimulation of the plant immune system is contained by the ROS-generating NADPH oxidase RBOHD (Torres et al., 2005), presumably to allow for the sufficient generation of SA-dependent defense signals from living cells that are adjacent to apoptotic cells. Nitric oxide (NO) plays an additional role in the regulation of SA/ROS-dependent defense (Trapet et al., 2015). This gaseous molecule can stimulate ROS production and cell death in the absence of SA while preventing excessive ROS production at high cellular SA levels via S-nitrosylation of RBOHD (Yun et al., 2011). Recently, it was shown that pathogen-induced accumulation of NO and ROS promotes the production of azelaic acid, a lipid derivative that primes distal plants for SA-dependent defenses (Wang et al., 2014). Hence, NO, ROS, and SA are intertwined in a complex regulatory network to mount local and systemic resistance against biotrophic pathogens. Interestingly, pathogens with a necrotrophic lifestyle can benefit from ROS/SA-dependent defenses and associated cell death (Govrin and Levine, 2000). For instance, Kabbage et al. (2013) demonstrated that S. sclerotiorum utilizes oxalic acid to repress oxidative defense signaling during initial biotrophic colonization, but it stimulates apoptosis at later stages to advance necrotrophic colonization. Moreover, SA-induced repression of JA-dependent resistance not only benefits necrotrophic pathogens but also hemibiotrophic pathogens after having switched from biotrophy to necrotrophy (Glazebrook, 2005; Pieterse et al., 2009, 2012).Plectosphaerella cucumerina ((P. cucumerina, anamorph Plectosporum tabacinum) anamorph Plectosporum tabacinum) is a filamentous ascomycete fungus that can survive saprophytically in soil by decomposing plant material (Palm et al., 1995). The fungus can cause sudden death and blight disease in a variety of crops (Chen et al., 1999; Harrington et al., 2000). Because P. cucumerina can infect Arabidopsis leaves, the P. cucumerina-Arabidopsis interaction has emerged as a popular model system in which to study plant defense reactions to necrotrophic fungi (Berrocal-Lobo et al., 2002; Ton and Mauch-Mani, 2004; Carlucci et al., 2012; Ramos et al., 2013). Various studies have shown that Arabidopsis deploys a wide range of inducible defense strategies against P. cucumerina, including JA-, SA-, ABA-, and auxin-dependent defenses, glucosinolates (Tierens et al., 2001; Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014), callose deposition (García-Andrade et al., 2011; Gamir et al., 2012, 2014; Sánchez-Vallet et al., 2012), and ROS (Tierens et al., 2002; Sánchez-Vallet et al., 2010; Barna et al., 2012; Gamir et al., 2012, 2014; Pastor et al., 2014). Recent metabolomics studies have revealed large-scale metabolic changes in P. cucumerina-infected Arabidopsis, presumably to mobilize chemical defenses (Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014). Furthermore, various chemical agents have been reported to induce resistance against P. cucumerina. These chemicals include β-amino-butyric acid, which primes callose deposition and SA-dependent defenses, benzothiadiazole (BTH or Bion; Görlach et al., 1996; Ton and Mauch-Mani, 2004), which activates SA-related defenses (Lawton et al., 1996; Ton and Mauch-Mani, 2004; Gamir et al., 2014; Luna et al., 2014), JA (Ton and Mauch-Mani, 2004), and ABA, which primes ROS and callose deposition (Ton and Mauch-Mani, 2004; Pastor et al., 2013). However, among all these studies, there is increasing controversy about the exact signaling pathways and defense responses contributing to plant resistance against P. cucumerina. While it is clear that JA and ethylene contribute to basal resistance against the fungus, the exact roles of SA, ABA, and ROS in P. cucumerina resistance vary between studies (Thomma et al., 1998; Ton and Mauch-Mani, 2004; Sánchez-Vallet et al., 2012; Gamir et al., 2014).This study is based on the observation that the disease phenotype during P. cucumerina infection differs according to the inoculation method used. We provide evidence that the fungus follows a hemibiotrophic infection strategy when infecting from relatively low spore densities on the leaf surface. By contrast, when challenged by localized host defense to relatively high spore densities, the fungus switches to a necrotrophic infection program. Our study has uncovered a novel strategy by which plant-pathogenic fungi can take advantage of the early immune response in the host plant.  相似文献   

11.
12.
13.
14.
15.
Lipid peroxide-derived toxic carbonyl compounds (oxylipin carbonyls), produced downstream of reactive oxygen species (ROS), were recently revealed to mediate abiotic stress-induced damage of plants. Here, we investigated how oxylipin carbonyls cause cell death. When tobacco (Nicotiana tabacum) Bright Yellow-2 (BY-2) cells were exposed to hydrogen peroxide, several species of short-chain oxylipin carbonyls [i.e. 4-hydroxy-(E)-2-nonenal and acrolein] accumulated and the cells underwent programmed cell death (PCD), as judged based on DNA fragmentation, an increase in terminal deoxynucleotidyl transferase dUTP nick end labeling-positive nuclei, and cytoplasm retraction. These oxylipin carbonyls caused PCD in BY-2 cells and roots of tobacco and Arabidopsis (Arabidopsis thaliana). To test the possibility that oxylipin carbonyls mediate an oxidative signal to cause PCD, we performed pharmacological and genetic experiments. Carnosine and hydralazine, having distinct chemistry for scavenging carbonyls, significantly suppressed the increase in oxylipin carbonyls and blocked PCD in BY-2 cells and Arabidopsis roots, but they did not affect the levels of ROS and lipid peroxides. A transgenic tobacco line that overproduces 2-alkenal reductase, an Arabidopsis enzyme to detoxify α,β-unsaturated carbonyls, suffered less PCD in root epidermis after hydrogen peroxide or salt treatment than did the wild type, whereas the ROS level increases due to the stress treatments were not different between the lines. From these results, we conclude that oxylipin carbonyls are involved in the PCD process in oxidatively stressed cells. Our comparison of the ability of distinct carbonyls to induce PCD in BY-2 cells revealed that acrolein and 4-hydroxy-(E)-2-nonenal are the most potent carbonyls. The physiological relevance and possible mechanisms of the carbonyl-induced PCD are discussed.In plants, environmental stressors such as extreme temperatures, drought, intense UV-B radiation, and soil salinity can cause tissue damage, growth inhibition, and even death. These detrimental effects are often ascribed to the action of reactive oxygen species (ROS) produced in the stressed plants for the following reasons: (1) various environmental stressors commonly cause the oxidation of biomolecules in plants; and (2) transgenic plants with enhanced antioxidant capacities show improved tolerance to environmental stressors (Suzuki et al., 2014). The production of ROS such as superoxide anion radical and hydrogen peroxide (H2O2) is intrinsically associated with photosynthesis and respiration (Foyer and Noctor, 2003; Asada, 2006).Plant cells are equipped with abundant antioxidant molecules such as α-tocopherol, β-carotene, and ascorbic acid and an array of ROS-scavenging enzymes such as superoxide dismutase and ascorbate peroxidase to maintain low intracellular ROS levels. When plants are exposed to severe and prolonged environmental stress, the balance between the production and scavenging of ROS is disrupted and the cellular metabolism reaches a new state of higher ROS production and lower antioxidant capacity. Then, the oxidation of vital biomolecules such as proteins and DNA proceeds, and as a consequence, cells undergo oxidative injury (Mano, 2002). The cause-effect relationship between ROS and tissue injury in plants is thus widely accepted, but the biochemical processes between the generation of ROS and cell death are poorly understood.Increasing evidence shows that oxylipin carbonyls mediate the oxidative injury of plants (Yamauchi et al., 2012; for review, see Mano, 2012; Farmer and Mueller, 2013). Oxylipin carbonyls are a group of carbonyl compounds derived from oxygenated lipids and fatty acids. The production of oxylipin carbonyls in living cells is explained as follows. Lipids in the membranes are constitutively oxidized by ROS to form lipid peroxides (LOOHs; Mène-Saffrané et al., 2007) because they are the most immediate and abundant targets near the ROS production sites. There are two types of LOOH formation reaction from ROS (Halliwell and Gutteridge, 2007). One is the radical-dependent reaction. Highly oxidizing radicals, such as hydroxyl radical (standard reduction potential of the HO/H2O pair, +2.31 V) and the protonated form of superoxide radical (HO2/H2O2, +1.06 V), can abstract a hydrogen atom from a lipid molecule, especially at the central carbon of a pentadiene structure in a polyunsaturated fatty acid, to form a radical. This organic radical rapidly reacts with molecular oxygen, forming a lipid hydroperoxyl radical, which then abstracts a hydrogen atom from a neighboring molecule and becomes a LOOH. The other reaction is the addition of singlet oxygen to a double bond of an unsaturated fatty acid to form an endoperoxide or a hydroperoxide (both are LOOHs). A variety of LOOH species are formed, depending on the source fatty acid and also by the oxygenation mechanism (Montillet et al., 2004). LOOH molecules are unstable, and in the presence of redox catalysts such as transition metal ions or free radicals, they decompose to form various aldehydes and ketones (i.e. oxylipin carbonyls; Farmer and Mueller, 2013). The chemical species of oxylipin carbonyl formed in the cells differ according to the fatty acids and the type of ROS involved (Grosch, 1987; Mano et al., 2014a).More than a dozen species of oxylipin carbonyls are formed in plants (for review, see Mano et al., 2009). Oxylipin carbonyls are constitutively formed in plants under normal physiological conditions, and the levels of certain types of oxylipin carbonyls rise severalfold under stress conditions, detected as increases in the free carbonyl content (Mano et al., 2010; Yin et al., 2010; Kai et al., 2012) and by the extent of the carbonyl modification of target proteins (Winger et al., 2007; Mano et al., 2014b). Among the oxylipin carbonyls, the α,β-unsaturated carbonyls, such as acrolein and 4-hydroxy-(E)-2-nonenal (HNE), have high reactivity and cytotoxicity (Esterbauer et al., 1991; Alméras et al., 2003). They strongly inactivate lipoate enzymes in mitochondria (Taylor et al., 2002) and thiol-regulated enzymes in chloroplasts (Mano et al., 2009) in vitro and cause tissue injury in leaves when they are fumigated (Matsui et al., 2012).The physiological relevance of oxylipin carbonyls has been shown by the observation that the overexpression of different carbonyl-scavenging enzymes commonly confers stress tolerance to transgenic plants (for review, see Mano, 2012). For example, 2-alkenal reductase (AER)-overproducing tobacco (Nicotiana tabacum) showed tolerance to aluminum (Yin et al., 2010), aldehyde dehydrogenase-overproducing Arabidopsis (Arabidopsis thaliana) showed tolerance to osmotic and oxidative stress (Sunkar et al., 2003), and aldehyde reductase-overproducing tobacco showed tolerance to chemical and drought stress (Oberschall et al., 2000). In addition, the genetic suppression of a carbonyl-scavenging enzyme made plants susceptible to stressors (Kotchoni et al., 2006; Shin et al., 2009; Yamauchi et al., 2012; Tang et al., 2014). Under stress conditions, there are positive correlations between the levels of certain carbonyls and the extent of tissue injury (Mano et al., 2010; Yin et al., 2010; Yamauchi et al., 2012). Thus, it is evident that oxylipin carbonyls, downstream products of ROS, are causes of oxidative damage in plant cells.To investigate how oxylipin carbonyls damage cells in oxidatively stressed plants, we here examined the mode of cell death that is induced by oxylipin carbonyls and identified the carbonyl species responsible for the cell death. We observed that oxylipin carbonyls cause programmed cell death (PCD), and our results demonstrated that the oxylipin carbonyls mediate the oxidative stress-induced PCD in tobacco Bright Yellow-2 (BY-2) cultured cells and in roots of tobacco and Arabidopsis plants. We then estimated the relative strengths of distinct carbonyl species to initiate the PCD program. Our findings demonstrate a critical role of the lipid metabolites in ROS signaling.  相似文献   

16.
We have established an efficient transient expression system with several vacuolar reporters to study the roles of endosomal sorting complex required for transport (ESCRT)-III subunits in regulating the formation of intraluminal vesicles of prevacuolar compartments (PVCs)/multivesicular bodies (MVBs) in plant cells. By measuring the distributions of reporters on/within the membrane of PVC/MVB or tonoplast, we have identified dominant negative mutants of ESCRT-III subunits that affect membrane protein degradation from both secretory and endocytic pathways. In addition, induced expression of these mutants resulted in reduction in luminal vesicles of PVC/MVB, along with increased detection of membrane-attaching vesicles inside the PVC/MVB. Transgenic Arabidopsis (Arabidopsis thaliana) plants with induced expression of ESCRT-III dominant negative mutants also displayed severe cotyledon developmental defects with reduced cell size, loss of the central vacuole, and abnormal chloroplast development in mesophyll cells, pointing out an essential role of the ESCRT-III complex in postembryonic development in plants. Finally, membrane dissociation of ESCRT-III components is important for their biological functions and is regulated by direct interaction among Vacuolar Protein Sorting-Associated Protein20-1 (VPS20.1), Sucrose Nonfermenting7-1, VPS2.1, and the adenosine triphosphatase VPS4/SUPPRESSOR OF K+ TRANSPORT GROWTH DEFECT1.Endomembrane trafficking in plant cells is complicated such that secretory, endocytic, and recycling pathways are usually integrated with each other at the post-Golgi compartments, among which, the trans-Golgi network (TGN) and prevacuolar compartment (PVC)/multivesicular body (MVB) are best studied (Tse et al., 2004; Lam et al., 2007a, 2007b; Müller et al., 2007; Foresti and Denecke, 2008; Hwang, 2008; Otegui and Spitzer, 2008; Robinson et al., 2008; Richter et al., 2009; Ding et al., 2012; Gao et al., 2014). Following the endocytic trafficking of a lipophilic dye, FM4-64, the TGN and PVC/MVB are sequentially labeled and thus are defined as the early and late endosome, respectively, in plant cells (Lam et al., 2007a; Chow et al., 2008). While the TGN is a tubular vesicular-like structure that may include several different microdomains and fit its biological function as a sorting station (Chow et al., 2008; Kang et al., 2011), the PVC/MVB is 200 to 500 nm in size with multiple luminal vesicles of approximately 40 nm (Tse et al., 2004). Membrane cargoes destined for degradation are sequestered into these tiny luminal vesicles and delivered to the lumen of the lytic vacuole (LV) via direct fusion between the PVC/MVB and the LV (Spitzer et al., 2009; Viotti et al., 2010; Cai et al., 2012). Therefore, the PVC/MVB functions between the TGN and LV as an intermediate organelle and decides the fate of membrane cargoes in the LV.In yeast (Saccharomyces cerevisiae), carboxypeptidase S (CPS) is synthesized as a type II integral membrane protein and sorted from the Golgi to the lumen of the vacuole (Spormann et al., 1992). Genetic analyses on the trafficking of CPS have led to the identification of approximately 17 class E genes (Piper et al., 1995; Babst et al., 1997, 2002a, 2002b; Odorizzi et al., 1998; Katzmann et al., 2001) that constitute the core endosomal sorting complex required for transport (ESCRT) machinery. The evolutionarily conserved ESCRT complex consists of several functionally different subcomplexes, ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III and the ESCRT-III-associated/Vacuolar Protein Sorting4 (VPS4) complex. Together, they form a complex protein-protein interaction network that coordinates sorting of cargoes and inward budding of the membrane on the MVB (Hurley and Hanson, 2010; Henne et al., 2011). Cargo proteins carrying ubiquitin signals are thought to be passed from one ESCRT subcomplex to the next, starting with their recognition by ESCRT-0 (Bilodeau et al., 2002, 2003; Hislop and von Zastrow, 2011; Le Bras et al., 2011; Shields and Piper, 2011; Urbé, 2011). ESCRT-0 recruits the ESCRT-I complex, a heterotetramer of VPS23, VPS28, VPS37, and MVB12, from the cytosol to the endosomal membrane (Katzmann et al., 2001, 2003). The C terminus of VPS28 interacts with the N terminus of VPS36, a member of the ESCRT-II complex (Kostelansky et al., 2006; Teo et al., 2006). Then, cargoes passed from ESCRT-I and ESCRT-II are concentrated in certain membrane domains of the endosome by ESCRT-III, which includes four coiled-coil proteins and is sufficient to induce the membrane invagination (Babst et al., 2002b; Saksena et al., 2009; Wollert et al., 2009). Finally, the ESCRT components are disassociated from the membrane by the adenosine triphosphatase (ATPase) associated with diverse cellular activities (AAA) VPS4/SUPPRESSOR OF K+ TRANSPORT GROWTH DEFECT1 (SKD1) before releasing the internal vesicles (Babst et al., 1997, 1998).Putative homologs of ESCRT-I–ESCRT-III and ESCRT-III-associated components have been identified in plants, except for ESCRT-0, which is only present in Opisthokonta (Winter and Hauser, 2006; Leung et al., 2008; Schellmann and Pimpl, 2009). To date, only a few plant ESCRT components have been studied in detail. The Arabidopsis (Arabidopsis thaliana) AAA ATPase SKD1 localized to the PVC/MVB and showed ATPase activity that was regulated by Lysosomal Trafficking Regulator-Interacting Protein5, a plant homolog of Vps Twenty Associated1 Protein (Haas et al., 2007). Expression of the dominant negative form of SKD1 caused an increase in the size of the MVB and a reduction in the number of internal vesicles (Haas et al., 2007). This protein also contributes to the maintenance of the central vacuole and might be associated with cell cycle regulation, as leaf trichomes expressing its dominant negative mutant form lost the central vacuole and frequently contained multiple nuclei (Shahriari et al., 2010). Double null mutants of CHARGED MULTIVESICULAR BODY PROTEIN, chmp1achmp1b, displayed severe growth defects and were seedling lethal. This may be due to the mislocalization of plasma membrane (PM) proteins, including those involved in auxin transport such as PINFORMED1, PINFORMED2, and AUXIN-RESISTANT1, from the vacuolar degradation pathway to the tonoplast of the LV (Spitzer et al., 2009).Plant ESCRT components usually contain several homologs, with the possibility of functional redundancy. Single mutants of individual ESCRT components may not result in an obvious phenotype, whereas knockout of all homologs of an ESCRT component by generating double or triple mutants may be lethal to the plant. As a first step to carry out systematic analysis on each ESCRT complex in plant cells, here, we established an efficient analysis system to monitor the localization changes of four vacuolar reporters that accumulate either in the lumen (LRR84A-GFP, EMP12-GFP, and aleurain-GFP) or on the tonoplast (GFP-VIT1) of the LV and identified several ESCRT-III dominant negative mutants. We reported that ESCRT-III subunits were involved in the release of PVC/MVB’s internal vesicles from the limiting membrane and were required for membrane protein degradation from secretory and endocytic pathways. In addition, transgenic Arabidopsis plants with induced expression of ESCRT-III dominant negative mutants showed severe cotyledon developmental defects. We also showed that membrane dissociation of ESCRT-III subunits was regulated by direct interaction with SKD1.  相似文献   

17.
In plants, the copy number of the mitochondrial DNA (mtDNA) can be much lower than the number of mitochondria. The biological significance and regulatory mechanisms of this phenomenon remain poorly understood. Here, using the pollen vegetative cell, we examined the role of the Arabidopsis (Arabidopsis thaliana) mtDNA-binding protein WHIRLY2 (AtWHY2). AtWHY2 decreases during pollen development, in parallel with the rapid degradation of mtDNA; to examine the importance of this decrease, we used the pollen vegetative cell-specific promoter Lat52 to express AtWHY2. The transgenic plants (LWHY2) had very high mtDNA levels in pollen, more than 10 times more than in the wild type (ecotype Columbia-0). LWHY2 plants were fertile, morphologically normal, and set seeds; however, reciprocal crosses with heterozygous plants showed reduced transmission of LWHY2-1 through the male and slower growth of LWHY2-1 pollen tubes. We found that LWHY2-1 pollen had significantly more reactive oxygen species and less ATP compared with the wild type, indicating an effect on mitochondrial respiration. These findings reveal that AtWHY2 affects mtDNA copy number in pollen and suggest that low mtDNA copy numbers might be the normal means by which plant cells maintain mitochondrial genetic information.Reflecting their endosymbiotic origin, mitochondria contain DNA genomes (mtDNA) encoding several key proteins for oxidative phosphorylation. As most genes identified in the mitochondrial genome are indispensable for mitochondrial function, it is generally believed that each mitochondrion must possess at least one full copy of the genome. Indeed, this seems to be the case in animals. For example, although the number of mitochondria per cell varies in human, mouse, rabbit, and rat cell lines, the mtDNA copy number per mitochondrion remains constant at 2.6 ± 0.3 (Robin and Wong, 1988). Also, in mouse egg cells, each mitochondrion contains an estimated one to two copies of the mtDNA (Pikó and Matsumoto, 1976).Plant cells, however, have very few copies of the mtDNA compared with the number of mitochondria. For example, in the Cucurbitaceae, cells containing 110 to 140 copies of the mtDNA have 360 to 1,100 mitochondria (Bendich and Gauriloff, 1984). In Arabidopsis (Arabidopsis thaliana), leaf cells each contain approximately 670 mitochondria (Sheahan et al., 2005) and approximately 50 copies of the mtDNA (Draper and Hays, 2000). Thus, in plant cells, each mitochondrion does not possess one complete copy of the mtDNA, a phenomenon that occurs commonly in somatic cells of plants (Preuten et al., 2010). In addition, work in Arabidopsis, barley (Hordeum vulgare), and tobacco (Nicotiana tabacum) showed that cells in leaves, stems, and roots contain few copies of the mtDNA (40–160), whereas cells in root tips contain more copies (300–450; Preuten et al., 2010). This is consistent with the mitochondrial nucleoid diminishment previously observed in developing root and shoot tips (Fujie et al., 1993, 1994), which suggests that the low copy numbers in plant cells result from a decrease in the mtDNA copy number in nondividing cells during development.One question raised by these findings is whether some mitochondria have complete mtDNAs while others have no mtDNA or whether mitochondria have partial mtDNAs. Using techniques for the direct visualization of small amounts of DNA, our group revealed that up to two-thirds of mitochondria in Arabidopsis mesophyll cells totally lack mtDNA and the remaining one-third of mitochondria possess mtDNA of about 100 kb on average (Wang et al., 2010). This agrees well with a previously reported value for mtDNA copy number (about 50 copies per cell; Draper and Hays, 2000) and is consistent with the idea that plant mitochondrial genomes exist as submolecules smaller than the total genomic sizes (Satoh et al., 1993; Kubo and Newton, 2008). Among plant cells possessing low mtDNA copy numbers, the vegetative cell in the pollen grains is an extreme case; a mature pollen grain of Antirrhinum majus, containing many more mitochondria than a somatic cell, possesses only 16 copies of the mtDNA (Wang et al., 2010). Similar to the changes observed in somatic cells, this extremely low level of mtDNA in pollen vegetative cells results from a rapid decrease in mtDNA copy number during pollen development (Sodmergen et al., 1991; Nagata et al., 1999). In A. majus, the vegetative cell in its initial developmental stage has 482.7 copies of the mtDNA per cell, indicating a 30-fold decrease (482.7/16) during development (Wang et al., 2010). These results from both somatic and reproductive cells led to the intriguing idea that the mtDNA copy number in plants decreases in parallel with cell differentiation, to a very low value, and thus that several mitochondria must share the genetic information carried on a single copy of the mtDNA. Plant cell mitochondria undergo frequent and coupled fusions and divisions, which may explain how mitochondria share this information (Arimura et al., 2004). However, the biological significance of why plant cells lose their mtDNA, and how this benefits these cells, remains unknown. Given that pollen germination, pollen tube elongation, and sperm cell delivery all require energy conversion, the extremely low mtDNA copy numbers, such as in pollen vegetative cells, must not compromise mitochondrial function.The mtDNA copy numbers remain constant in various tissues, however, indicating that cellular mechanisms accurately regulate the levels of mtDNA in relation to cell type (Robin and Wong, 1988; Preuten et al., 2010). In yeast and animals, this regulation involves the core enzymes of mtDNA replication, such as DNA polymerase-γ (Sharief et al., 1999), RNA polymerase (Wanrooij et al., 2008), and mitochondrial helicase (Liu et al., 2009), as well as a group of DNA-binding proteins such as ARS-binding factor2 protein in yeast (Saccharomyces cerevisiae; Newman et al., 1996), MITOCHONDRIAL TRANSCRIPTION FACTOR A (TFAM) in human (Alam et al., 2003), and mitochondrial single-stranded DNA binding protein in Drosophila spp. (Maier et al., 2001). Overexpression of TFAM causes an increase in the mtDNA copy number, and RNA interference of TFAM decreases the mtDNA copy number (Ekstrand et al., 2004; Kanki et al., 2004). Also, the homozygous knockout of TFAM in mouse results in embryos that lack mtDNA and thus fail to survive (Larsson et al., 1998). Clearly, protein factors within mitochondrial nucleoids play a crucial role in the regulation of mtDNA copy number.Recent investigation in Arabidopsis revealed that, similar to the case in yeast and animal cells, DNA polymerase, the core enzyme of mtDNA replication, functions to maintain mtDNA levels in plants. Mutation of Arabidopsis PolIA or PolIB (homologs of bacterial DNA polymerase I) causes a reduction in mtDNA copy number, and double mutation of these proteins is lethal (Parent et al., 2011). Also, an Mg2+-dependent exonuclease, DEFECTIVE IN POLLEN ORGANELLE DNA DEGRADATION1 (DPD1), degrades organelle DNA, helping to produce the proper amounts of mtDNA in pollen cells (Matsushima et al., 2011; Tang et al., 2012). These results provide insights into the molecular control of mtDNA levels in plants, via both mtDNA replication and mtDNA degradation. Except for these enzymes, however, other protein factors (such as TFAM in animals) have not been identified in plants. The DNA-binding proteins, such as MutS Homolog1 (MSH1), Organellar Single-Strand DNA Binding Protein1 (OSB1), Recombinase A1 (RecA1), RecA3, and WHIRLY2 (WHY2), identified so far in plant mitochondria likely participate in genomic maintenance by affecting substoichiometric shifting (Abdelnoor et al., 2003), stoichiometric transmission (Zaegel et al., 2006), genomic stability (Shedge et al., 2007; Odahara et al., 2009), and DNA repair (Cappadocia et al., 2010). None of these plant nucleoid factors (DNA-binding proteins) has been implicated in the control of mtDNA copy number; thus, the mechanisms by which nonenzyme protein factors regulate mtDNA copy number in plants remain obscure.To test whether nucleoid DNA-binding proteins can affect mtDNA copy number, we examined the effect of producing Arabidopsis WHY2, a single-stranded DNA-binding protein (Cappadocia et al., 2010), in the pollen vegetative cell, which generally does not express WHY2 (Honys and Twell, 2004). We found that expression of WHY2 resulted in a 10-fold increase in mtDNA copy number in the pollen vegetative cell. This increase affected mitochondrial respiration, mitochondrial size, and pollen tube growth. Thus, our results uncover a novel function for WHY2, a member of the plant Whirly protein family, in regulating mtDNA amounts and indicate that, in plants, low mtDNA copy number does not compromise mitochondrial function but rather promotes proper mitochondrial function.  相似文献   

18.
19.
20.
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