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.
Heat shock proteins (HSPs) function as molecular chaperones and are essential for the maintenance and/or restoration of protein homeostasis. The genus Xanthomonas type III effector protein AvrBsT induces hypersensitive cell death in pepper (Capsicum annuum). Here, we report the identification of the pepper CaHSP70a as an AvrBsT-interacting protein. Bimolecular fluorescence complementation and coimmunoprecipitation assays confirm the specific interaction between CaHSP70a and AvrBsT in planta. The CaHSP70a peptide-binding domain is essential for its interaction with AvrBsT. Heat stress (37°C) and Xanthomonas campestris pv vesicatoria (Xcv) infection distinctly induce CaHSP70a in pepper leaves. Cytoplasmic CaHSP70a proteins significantly accumulate in pepper leaves to induce the hypersensitive cell death response by Xcv (avrBsT) infection. Transient CaHSP70a overexpression induces hypersensitive cell death under heat stress, which is accompanied by strong induction of defense- and cell death-related genes. The CaHSP70a peptide-binding domain and ATPase-binding domain are required to trigger cell death under heat stress. Transient coexpression of CaHSP70a and avrBsT leads to cytoplasmic localization of the CaHSP70a-AvrBsT complex and significantly enhances avrBsT-triggered cell death in Nicotiana benthamiana. CaHSP70a silencing in pepper enhances Xcv growth but disrupts the reactive oxygen species burst and cell death response during Xcv infection. Expression of some defense marker genes is significantly reduced in CaHSP70a-silenced leaves, with lower levels of the defense hormones salicylic acid and jasmonic acid. Together, these results suggest that CaHSP70a interacts with the type III effector AvrBsT and is required for cell death and immunity in plants.The heat shock protein HSP70 is a ubiquitous essential protein chaperone and one of the most abundant and diverse heat stress proteins in plants. HSP70s are induced by environmental stresses and are required for plants to cope with heat. HSP70s are involved in protein folding, synthesis, translocation, and macromolecular assemblies such as microtubules (Mayer et al., 2001; Hartl and Hayer-Hartl, 2002). HSP70s protect cells from heat stress by preventing protein aggregation and by facilitating the refolding of denatured proteins. Protein stability can decrease under heat stress conditions and expose hydrophobic patches that cause the aggregation of denatured proteins. HSP70s bind to hydrophobic patches of partially unfolded proteins in an ATP-dependent manner and prevent protein aggregation (Mayer and Bukau, 2005). The modular HSP70 structure consists of a N-terminal ATPase domain and a C-terminal peptide-binding domain that contains a β-sandwich subdomain with a peptide-binding cleft and an α-helical latch-like segment (Zhu et al., 1996; Hartl and Hayer-Hartl, 2002).HSP70s are involved in microbial pathogenesis, cell death responses, and immune responses. Diverse RNA viruses induce HSP70 expression in Arabidopsis (Arabidopsis thaliana; Whitham et al., 2003). Cytoplasmic HSP70s enhance the infection of Nicotiana benthamiana by Tobacco mosaic virus, Potato virus X, Cucumber mosaic virus, and Watermelon mosaic virus (Chen et al., 2008). Recently, the coat protein of Tomato yellow leaf curl virus was suggested to recruit host plant HSP70 during virus infection (Gorovits et al., 2013). HSP70s appear to be involved in regulating viral reproduction, protein folding, and movement, which ultimately promotes viral infection (Boevink and Oparka, 2005; Hafrén et al., 2010). The Pseudomonas syringae effector protein Hopl1 directly binds and manipulates host HSP70, which promotes bacterial virulence (Jelenska et al., 2010). The cytosolic/nuclear heat shock cognate 70 (HSC70) chaperone, which is highly homologous to HSP70 (Tavaria et al., 1996), regulates Arabidopsis immune responses together with SGT1 (for the suppressor of the G2 allele of S-phase kinase-associated protein1 [skp1]; Noël et al., 2007). Cytoplasmic HSP70 is required for the Phytophthora infestans INF1-mediated hypersensitive response (HR) and nonhost resistance to Pseudomonas cichorii in N. benthamiana (Kanzaki et al., 2003). HSP70 is proposed to be involved in both positive and negative regulation of cell death. Selective HSP70 depletion from human cell lines activates a tumor-specific death program that is independent of known caspases and p53 tumor-suppressor protein (Nylandsted et al., 2000), whereas HSP70 promotes tumor necrosis factor-mediated apoptosis by binding IkB kinase γ and impairing nuclear factor-κB signaling in Cos-1 cells (Ran et al., 2004). In N. benthamiana, HSP70 is required for tabtoxinine-β-lactam-induced cell death (Ito et al., 2014). However, HSP70 expression is shown to decrease the cell death triggered by salicylic acid (SA) in Nicotiana tabacum protoplasts (Cronjé et al., 2004). Overexpression of mitochondrial HSP70 suppresses heat- and hydrogen peroxide (H2O2)-induced programmed cell death in rice (Oryza sativa; Qi et al., 2011).The genus Xanthomonas YopJ-like AvrBsT protein activates effector-triggered immunity (ETI) in Arabidopsis Pitztal 0 plants (Cunnac et al., 2007). AvrBsT is a member of the YopJ/AvrRxv family identified in Xanthomonas campestris pv vesicatoria (Xcv; Lewis et al., 2011). AvrBsT alters phospholipid signaling and activates defense responses in Arabidopsis (Kirik and Mudgett, 2009). AvrBsT is an acetyltransferase that acetylates Arabidopsis ACETYLATED INTERACTING PROTEIN1 (ACIP1), a microtubule-associated protein required for plant immunity (Cheong et al., 2014). Xcv strain Bv5-4a secretes the AvrBsT type III effector protein that induces hypersensitive cell death and strong defense responses in pepper (Capsicum annuum) and N. benthamiana (Orth et al., 2000; Escolar et al., 2001; Kim et al., 2010). AvrBsT-induced HR-like cell death in pepper is likely part of the typical ETI-mediated defense response cascade (Jones and Dangl, 2006; Eitas et al., 2008; Eitas and Dangl, 2010). AvrBsT overexpression in Arabidopsis triggers plant cell death and defense signaling, leading to both disease and defense responses to diverse microbial pathogens (Hwang et al., 2012). Type III effectors such as Hopl1 and AvrBsT are used to identify unknown components of plant defense cascades (Nomura et al., 2006; Block et al., 2008; Jelenska et al., 2010; Kim et al., 2014) that modulate host innate immunity to achieve disease resistance. The pepper SGT1 was identified recently as a host interactor of AvrBsT (Kim et al., 2014). Pepper SGT1 has features of a cochaperone (Shirasu and Schulze-Lefert, 2003), interacts with AvrBsT, and promotes hypersensitive cell death associated with the pepper receptor-like cytoplasmic protein kinase1 (PIK1) phosphorylation cascade.In this study, we used a yeast (Saccharomyces cerevisiae) two-hybrid screen to identify the pepper HSP70a (CaHSP70a) as an interacting partner of the Xanthomonas spp. type III effector AvrBsT. Coimmunoprecipitation and bimolecular fluorescence complementation (BiFC) analyses verify that CaHSP70a interacts with AvrBsT in planta. Transient CaHSP70a overexpression in pepper leaves enhances heat stress sensitivity and leads to a cell death response. Cytoplasmic localization of the AvrBsT-CaHSP70a complex strongly elevates cell death. CaHSP70a expression is rapidly and strongly induced by avrBsT (for avirulent Xcv Dukso1 [Ds1]) infection in pepper. CaHSP70a silencing enhances susceptibility to Xcv infection, attenuates the reactive oxygen species (ROS) burst and cell death response, reduces SA and jasmonic acid (JA) levels, and disrupts expression of the defense response genes C. annuum pathogenesis-related protein1 (CaPR1; Kim and Hwang, 2000), CaPR10 (Choi et al., 2012), and CaDEF1 (for defensin; Do et al., 2004). Taken together, this study demonstrates that CaHSP70a is a target of the Xanthomonas spp. type III effector AvrBsT and acts as a positive regulator of plant cell death and immunity signaling.  相似文献   

2.
3.
4.
5.
6.
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   

7.
8.
9.
The membrane-bound BRASSINOSTEROID INSENSITIVE1-ASSOCIATED RECEPTOR KINASE1 (BAK1) is a common coreceptor in plants and regulates distinct cellular programs ranging from growth and development to defense against pathogens. BAK1 functions through binding to ligand-stimulated transmembrane receptors and activating their kinase domains via transphosphorylation. In the absence of microbes, BAK1 activity may be suppressed by different mechanisms, like interaction with the regulatory BIR (for BAK1-INTERACTING RECEPTOR-LIKE KINASE) proteins. Here, we demonstrated that BAK1 overexpression in Arabidopsis (Arabidopsis thaliana) could cause detrimental effects on plant development, including growth arrest, leaf necrosis, and reduced seed production. Further analysis using an inducible expression system showed that BAK1 accumulation quickly stimulated immune responses, even under axenic conditions, and led to increased resistance to pathogenic Pseudomonas syringae pv tomato DC3000. Intriguingly, our study also revealed that the plasma membrane-associated BAK1 ectodomain was sufficient to induce autoimmunity, indicating a novel mode of action for BAK1 in immunity control. We postulate that an excess of BAK1 or its ectodomain could trigger immune receptor activation in the absence of microbes through unbalancing regulatory interactions, including those with BIRs. Consistently, mutation of SUPPRESSOR OF BIR1-1, which encodes an emerging positive regulator of transmembrane receptors in plants, suppressed the effects of BAK1 overexpression. In conclusion, our findings unravel a new role for the BAK1 ectodomain in the tight regulation of Arabidopsis immune receptors necessary to avoid inappropriate activation of immunity.Plants rely on their innate immune system to detect microbes and mount an active defense against pathogens. The plant immune system is traditionally considered to be composed of two layers (Jones and Dangl, 2006). The first one is based on the activity of pattern-recognition receptors (PRRs) that can detect microbe-associated molecular patterns (MAMPs) and trigger what is termed pattern-triggered immunity (PTI; Boller and Felix, 2009). Many plant pathogens can suppress this basal defense response using virulence factors termed effectors. In a second layer of defense, plants can make use of resistance (R) proteins to recognize the presence of pathogen effectors resulting in effector-triggered immunity (ETI), which resembles an accelerated and amplified PTI response (Jones and Dangl, 2006).Plants utilize plasma membrane-associated receptor-like proteins (RLPs) or receptor-like kinases (RLKs) as PRRs to sense specific signals through their ectodomains (Böhm et al., 2014). RLPs and RLKs require the function of additional RLKs to form active receptor complexes and transfer the external signal to the inside of the cells (Zhang and Thomma, 2013; Cao et al., 2014; Liebrand et al., 2014). The best-known coreceptor is the leucine-rich repeat (LRR)-RLK BRASSINOSTEROID INSENSITIVE1-ASSOCIATED RECEPTOR KINASE1 (BAK1), which was originally identified as a positive regulator and partner for the brassinosteroid (BR) receptor BRASSINOSTEROID INSENSITIVE1 (BRI1; Li et al., 2002; Nam and Li, 2002). BRs refer to phytohormones that promote plant growth and development (Fujioka and Yokota, 2003). Thus, loss-of-function mutations in BAK1 negatively impact Arabidopsis (Arabidopsis thaliana) growth due to improper cell elongation. In short, bak1 mutants display compact rosettes with round-shaped leaves and shorter petioles and phenocopy weak bri1 mutations (Li et al., 2002; Nam and Li, 2002). Conversely, certain mutants affected in the BAK1 ectodomain show increased activity in the BR signaling pathway and share phenotypic similarities with BRI1-overexpressing lines (Wang et al., 2001), including elongated hypocotyls, petioles, and leaf blades and an overall increase in height (Jaillais et al., 2011; Chung et al., 2012).Furthermore, BAK1 is involved in the containment of cell death, independently of its function in BR signaling. Arabidopsis bak1 knockout mutants exhibit extensive cell death spreading after microbial infection (Kemmerling et al., 2007). In addition, spontaneous cell death develops in Arabidopsis double mutant plants lacking both BAK1 (also named SOMATIC EMBRYOGENESIS RECEPTOR KINASE3 [SERK3]) and its closest homolog BAK1-LIKE1 (BKK1)/SERK4, causing seedling lethality even in the absence of microbes (He et al., 2007). Similar phenotypes are observed in Arabidopsis, rice (Oryza sativa), and Nicotiana benthamiana by lowering the expression of BAK1 and its homologs (Heese et al., 2007; Jeong et al., 2010; Park et al., 2011). Interestingly, typical defense responses, like the production of reactive oxygen species and constitutive callose deposition, are also detected in those plants, although the basis for this phenomenon remains poorly understood (He et al., 2007; Kemmerling et al., 2007; Park et al., 2011; Gao et al., 2013).On the other hand, BAK1 is widely studied as a key component of immune signaling pathways due to its known association with different PRRs, including RLKs and RLPs (Kim et al., 2013; Böhm et al., 2014). Upon MAMP perception, PRRs induce signaling and physiological defense responses like mitogen-activated protein kinase (MAPK) activation, reactive oxygen species and ethylene production, and modifications in gene expression, all of which contribute to PTI. Among the best-studied examples of BAK1-regulated PRRs are two LRR-receptor kinases, ELONGATION FACTOR Tu RECEPTOR (EFR), which senses the active epitope elf18 of the bacterial elongation factor Tu, and the flagellin receptor FLAGELLIN SENSING2 (FLS2), which senses the active epitope flg22 of bacterial flagellin (Gómez-Gómez and Boller, 2000; Chinchilla et al., 2006; Zipfel et al., 2006). Immediately after flg22 binding to its LRR ectodomain, FLS2 forms a tight complex with BAK1 (Chinchilla et al., 2007; Sun et al., 2013). This heteromerization step may bring the two kinase domains closer and thereby induce, within seconds, the phosphorylation of BAK1 and FLS2 (Schulze et al., 2010; Schwessinger et al., 2011). These steps are sufficient to initiate the immune signaling pathway, even if the ectodomains and kinase domains are switched between FLS2 and BAK1 (Albert et al., 2013).While PRRs, such as FLS2 and EFR, are extremely sensitive to even subnanomolar concentrations of their ligands, a tight control of these receptors is expected, since constitutive activation of defense responses in plants dramatically impairs fitness and growth (Tian et al., 2003; Korves and Bergelson, 2004). However, the mechanisms that underlie the attenuation of PRR activation or prevent these receptors from signaling constitutively remain largely unknown (Macho and Zipfel, 2014). Several independent observations indicate that BAK1 and FLS2 are present in close spatial proximity in preformed complexes at the plasma membrane (Chinchilla et al., 2007; Schulze et al., 2010; Roux et al., 2011). Negative regulation of immune signaling prior to ligand perception could happen within the PRR complex and depend on conformational changes following the association of FLS2 with flg22 (Meindl et al., 2000; Schulze et al., 2010; Mueller et al., 2012). Additionally, other partners might prevent the constitutive interaction of BAK1 with FLS2. Such could be the case for the LRR-RLK BAK1-INTERACTING RECEPTOR-LIKE KINASEs (BIRs): BIR2 was recently discovered as a substrate and negative regulator for BAK1, while the absence of BIR1 leads to the activation of defense induction and strong dwarfism (Gao et al., 2009; Halter et al., 2014b). Furthermore, MAMP signaling may be constrained by phosphatases, as suggested in earlier studies (Felix et al., 1994; Gómez-Gómez et al., 2001) and recently shown for the protein phosphatase 2A, which controls PRR activation likely by modulating the BAK1 phosphostatus (Segonzac et al., 2014). These examples illustrate the variety of mechanisms that may tightly control BAK1 activity.In this work, we show that regulation of BAK1 accumulation is crucial for Arabidopsis fitness, as its overexpression leads to dwarfism and premature death. The phenotype differs from BR mutants and is very reminiscent of or even identical to the autoimmune phenotype of plants showing constitutive activation of R proteins (Oldroyd and Staskawicz, 1998; Bendahmane et al., 2002; Zhang et al., 2003). BAK1 overexpression is associated with constitutive activation of defense pathway(s) involving the general coregulator of RLPs, SUPPRESSOR OF BIR1-1 (SOBIR1; Liebrand et al., 2013, 2014). To our knowledge, this is the first report and comprehensive characterization of such an autoimmunity phenotype for Arabidopsis plants overexpressing BAK1, and it highlights the importance of the regulation of PTI overactivation.  相似文献   

10.
11.
12.
13.
14.
15.
16.
17.
In Arabidopsis (Arabidopsis thaliana), the Pseudomonas syringae effector proteins AvrB and AvrRpm1 are both detected by the RESISTANCE TO PSEUDOMONAS MACULICOLA1 (RPM1) disease resistance (R) protein. By contrast, soybean (Glycine max) can distinguish between these effectors, with AvrB and AvrRpm1 being detected by the Resistance to Pseudomonas glycinea 1b (Rpg1b) and Rpg1r R proteins, respectively. We have been using these genes to investigate the evolution of R gene specificity and have previously identified RPM1 and Rpg1b. Here, we report the cloning of Rpg1r, which, like RPM1 and Rpg1b, encodes a coiled-coil (CC)-nucleotide-binding (NB)-leucine-rich repeat (LRR) protein. As previously found for Rpg1b, we determined that Rpg1r is not orthologous with RPM1, indicating that the ability to detect both AvrB and AvrRpm1 evolved independently in soybean and Arabidopsis. The tightly linked soybean Rpg1b and Rpg1r genes share a close evolutionary relationship, with Rpg1b containing a recombination event that combined a NB domain closely related to Rpg1r with CC and LRR domains from a more distantly related CC-NB-LRR gene. Using structural modeling, we mapped polymorphisms between Rpg1b and Rpg1r onto the predicted tertiary structure of Rpg1b, which revealed highly polymorphic surfaces within both the CC and LRR domains. Assessment of chimeras between Rpg1b and Rpg1r using a transient expression system revealed that AvrB versus AvrRpm1 specificity is determined by the C-terminal portion of the LRR domain. The P. syringae effector AvrRpt2, which targets RPM1 INTERACTOR4 (RIN4) proteins in both Arabidopsis and soybean, partially blocked recognition of both AvrB and AvrRpm1 in soybean, suggesting that both Rpg1b and Rpg1r may detect these effectors via modification of a RIN4 homolog.Effector triggered immunity in plants involves highly specific recognition events in which plant resistance (R) proteins detect pathogen effector proteins directly or, alternatively, the modifications that they induce on host proteins (Bonardi et al., 2012). The largest group of R proteins belongs to the nucleotide-binding (NB)-leucine-rich repeat (LRR) family (McHale et al., 2006). The NB-LRR family can be further subdivided based on N-terminal domains into the Toll-Interleukin and R protein (TIR) class and non-TIR-NB-LRR class (McHale et al., 2006). The latter most often contain a coiled-coil (CC) domain at the N terminus. The contributions of the TIR, CC, and LRR domains to R protein specificity, and how new specificities evolve, remain important questions.There are relatively few NB-LRR R proteins characterized to date that are thought to detect pathogen effectors directly; these include Pi-ta from rice (Oryza sativa), L and M variants from flax (Linum usitatissimum), and RESISTANCE TO RALSTONIA SOLANACEARUM1 and RESISTANCE TO PERONOSPORA PARASITICA1 (RPP1) from Arabidopsis (Arabidopsis thaliana; Jia et al., 2000; Deslandes et al., 2003; Dodds et al., 2006; Ueda et al., 2006; Catanzariti et al., 2010; Krasileva et al., 2010). In at least some of these examples, the R genes are found in clusters of NB-LRR paralogs in which multiple recognition specificities are represented (Ellis et al., 1995; Botella et al., 1998) or belong to allelic series (Ellis et al., 1995), arrangements that may promote evolution of recognition specificity via recombination between alleles and paralogs. Interestingly, sequence comparisons and domain swaps involving alleles at the L locus implicate both the LRR and TIR regions as determinants of recognition specificity (Ellis et al., 1999; Luck et al., 2000). Subsequently, domain swaps involving paralogs clustered at the barley (Hordeum vulgare) MILDEW A (MLA) and potato (Solanum tuberosum) Resistance to Potato Virus X (Rx)/Globodera pallida (Gpa) loci have provided additional support for the LRR domain playing a key role in conferring recognition specificity (Ellis et al., 1999; Luck et al., 2000; Shen et al., 2003; Rairdan and Moffett, 2006).Several R proteins are known to detect the presence of pathogen effectors indirectly by monitoring the activity of pathogen effectors within the plant cell. For example, the Arabidopsis RESISTANCE TO PSEUDOMONAS MACULICOLA1 (RPM1) and RESISTANCE TO PSEUDOMONAS SYRINGAE2 (RPS2) R proteins detect modification of the effector target RPM1 INTERACTOR4 (RIN4), while the Arabidopsis RPS5 protein detects modification of the effector target AvrPphB SUSCEPTIBLE1 (Mackey et al., 2002, 2003; Axtell and Staskawicz, 2003; Shao et al., 2003). At least for the well-studied examples in Arabidopsis, R proteins that employ indirect recognition mechanisms are encoded by NB-LRR genes that are not members of large clusters, or allelic series, with variants encoding distinct recognition specificities. Correlated with this genomic structure, such loci are typically relatively stable, with RPM1 and RPS5 existing as presence/absence polymorphisms that have been maintained over long evolutionary periods (Stahl et al., 1999; Tian et al., 2002). Both functional and nonfunctional alleles of RPS2 have been isolated, but only a single recognition specificity has been detected at this locus, despite sequence polymorphisms between alleles (Caicedo et al., 1999).Most likely, specificity for this class of R proteins is determined by a combination of the ability to associate with the host protein targeted by the effector and the ability to detect effector-induced modification of this target. Consistent with this hypothesis, it has been shown that the CC domains from at least some R proteins interact with the host proteins they are monitoring, even in the absence of pathogen effectors, in a prerecognition complex (Mackey et al., 2002; Ade et al., 2007). Hence, evolution of recognition specificity in R proteins that employ indirect recognition mechanisms may involve evolution of both the N-terminal CC and LRR domains.To better understand the evolution and function of R proteins that detect pathogen effectors indirectly, we have been studying two soybean (Glycine max) R genes, with known recognition specificities, that are members of a complex NB-LRR cluster. The R genes involved, Resistance to Pseudomonas glycinea 1b (Rpg1b) and Rpg1r, mediate detection of the Pseudomonas syringae effector proteins AvrB and AvrRpm1, respectively (Staskawicz et al., 1984; Ashfield et al., 1995). We have previously cloned Rpg1b, which is a CC-NB-LRR (CNL) gene that maps to a cluster of R genes effective against a diverse range of pathogens (Ashfield et al., 1998, 2004). Rpg1r is present in the same cluster and maps 0.56 centiMorgans from Rpg1b (Ashfield et al., 1995); however, the evolutionary relationship shared by the two R genes is not known. The cluster is associated with numerous NB-LRR genes, of both the CC and TIR subgroups, spread over more than a megabase of soybean chromosome 13 (Peñuela et al., 2002; Hayes et al., 2004; Innes et al., 2008; Ashfield et al., 2012; Wen et al., 2013). The NB-LRR family in this region is evolving rapidly, with duplications/deletions of paralogs, recombination, and positive selection all playing a role (Ashfield et al., 2012).While soybean can distinguish between AvrB and AvrRpm1, both effectors are detected by a single R protein, RPM1, in Arabidopsis (Bisgrove et al., 1994; Grant et al., 1995). It is known that RPM1 recognizes the effector proteins indirectly by detecting effector-dependent phosphorylation of a second Arabidopsis protein, RIN4 (Mackey et al., 2002; Chung et al., 2011; Liu et al., 2011). The available evidence suggests that a related strategy is employed by soybean, at least for the Rpg1b protein, despite the AvrB recognition specificity having evolved independently in these plant species (Ashfield et al., 2004; Selote and Kachroo, 2010; Selote et al., 2013). Soybean contains four RIN4 homologs (Chen et al., 2010), three of which interact physically with Rpg1b, with two required for full resistance conferred by this R gene (Selote and Kachroo, 2010; Selote et al., 2013). It is not known whether RIN4 homologs are required for Rpg1r function.Here, we report the map-based cloning of the soybean Rpg1r gene. Comparison of the Rpg1r protein to Rpg1b, combined with structural modeling, revealed highly polymorphic surfaces in the CC and LRR domains. Transient expression of chimeric Rpg1 proteins demonstrated that specificity for AvrB versus AvrRpm1 is determined by the C-terminal LRR region. Finally, we provide evidence that Rpg1r, like Rpg1b, detects its corresponding pathogen effector indirectly, most likely by monitoring a RIN4 homolog, indicating convergent evolution of recognition mechanisms in separate plant families.  相似文献   

18.
19.
Holliday junctions (HJs) are physical links between homologous DNA molecules that arise as central intermediary structures during homologous recombination and repair in meiotic and somatic cells. It is necessary for these structures to be resolved to ensure correct chromosome segregation and other functions. In eukaryotes, including plants, homologs of a gene called XPG-like endonuclease1 (GEN1) have been identified that process HJs in a manner analogous to the HJ resolvases of phages, archaea, and bacteria. Here, we report that Arabidopsis (Arabidopsis thaliana), a eukaryotic organism, has two functional GEN1 homologs instead of one. Like all known eukaryotic resolvases, AtGEN1 and Arabidopsis single-strand DNA endonuclease1 both belong to class IV of the Rad2/XPG family of nucleases. Their resolvase activity shares the characteristics of the Escherichia coli radiation and UV sensitive C paradigm for resolvases, which involves resolving HJs by symmetrically oriented incisions in two opposing strands. This leads to ligatable products without the need for further processing. The observation that the sequence context influences the cleavage by the enzymes can be interpreted as a hint for the existence of sequence specificity. The two Arabidopsis paralogs differ in their preferred sequences. The precise cleavage positions observed for the resolution of mobile nicked HJs suggest that these cleavage positions are determined by both the substrate structure and the sequence context at the junction point.To counter the effects of endogenous and exogenous factors that threaten the genome integrity, efficient mechanisms have evolved to ensure the faithful transmission of genetic information (Tuteja et al., 2001). Double-strand breaks, induced by conditions such as ionizing radiation or replication fork (RF) stalling, are among the most deleterious lesions (Jackson and Bartek, 2009). To protect the genome from consequences of these lesions, the cells have ancient double-strand break repair mechanisms, including the homologous recombination (HR) pathway. The HR mechanism is also of great importance in the intentional genetic recombination during sexual reproduction. A key intermediate in HR is the so-called Holliday junction (HJ), a structure that was first suggested in the context of a gene conversion model in fungi (Holliday, 1964) and later shown to arise in somatic and meiotic cells (Szostak et al., 1983; Schwacha and Kleckner, 1995; Cromie et al., 2006; Bzymek et al., 2010).HJs are structures consisting of four DNA strands of two homologous DNA helices (e.g. homologous chromosomes or sister chromatids). They arise through invasion of one single strand from each of two helices into the other double strand. This results in two continuous strands (one per helix) and two strands that cross from one helix into the other. Schematics often depict the HJs with a parallel orientation of the helices, in which the crossing strands cross each other as was originally postulated (Holliday, 1964). However, HJs based on oligonucleotides have been shown to adopt an antiparallel conformation (for review, see Lilley, 2000). In this configuration, the junction resembles the letter H in a lateral view, and the crossing strands actually perform U turns. The crossing strands represent physical links between the two DNA strands involved. If a RF is restored by HR-mediated repair during mitosis, the resulting HJ usually involves the two sister chromatids of one chromosome (Li and Heyer, 2008). In meiosis, the physical links in the shape of HJs arise because of meiotic crossover between homologous chromosomes. In either case, these links must be resolved to ensure unperturbed cell survival.The importance of resolving the HJs for the survival of cells and organisms is highlighted by the phenotypes described for mutants defective for the known pathways of HJ resolution. One of these pathways is the resolution by canonical HJ resolvases, enzymes that cleave the two opposing strands of a HJ in perfectly symmetric positions relative to the junction point, which results in readily ligatable nicked duplex (nD) products (Svendsen and Harper, 2010). This property distinguishes the canonical HJ resolvases from the noncanonical resolvases (see below).The main resolvase of Escherichia coli is radiation and UV sensitive C (RuvC), which is part of the E. coli resolvasome (RuvABC complex; Otsuji et al., 1974; Sharples et al., 1990, 1999). In this complex, a HJ is sandwiched between two RuvA tetramers (Panyutin and Hsieh, 1994). Two RuvB complexes form ATP-dependent motors of branch migration, with two opposing helical arms of the junction threaded through their central openings. For the resolution of the HJ, one RuvA tetramer is replaced by a RuvC homodimer. This homodimer positions two active sites at the center of the junction that are poised to cleave the junction point if a preferred consensus sequence of the form 5′-(A/T)TT(G/C)-3′ is encountered. The requirement for this correct sequence is quite strict; even a single base change can lead to a drastic reduction of the cleavage efficiency (Shah et al., 1994). Isolated EcRuvC is also active in vitro and binds only HJ structures with high specificity. This binding is independent of the sequence context, but the cleavage depends on the specific sequence (Iwasaki et al., 1991; Benson and West, 1994; Dunderdale et al., 1994). The exact cleavage position has been determined to be either one nucleotide 3′ or 5′ from the junction or at the junction point (Bennett and West, 1996; Shida et al., 1996; Osman et al., 2009). The well-characterized EcRuvC is often referred to as a paradigm of canonical HJ resolution.Eukaryotes have evolved a more complex interplay of different HJ resolution pathways (Schwartz and Heyer, 2011; Zakharyevich et al., 2012). A defined complex, consisting of a recombination deficiency Q (RecQ) helicase (AtRECQ4A in Arabidopsis [Arabidopsis thaliana], Bloom syndrome protein in human, and Slow growth suppression1 (Sgs1) in yeast [Saccharomyces cerevisiae]), a type IA topoisomerase (DNA topoisomerase 3-alpha [TOP3A] in Arabidopsis, HsTOPOIIIα in human, and ScTop3 in yeast), and the structural protein RecQ-mediated genome instability1 (AtRMI1 in Arabidopsis, HsRMI1 in human, and ScRmi1 in yeast; RTR complex), mediates the so-called dissolution pathway. The crossing points of a double HJ are brought together by branch migration catalyzed by the helicase followed by decatenation catalyzed by the topoisomerase (Wu and Hickson, 2003; Hartung et al., 2007a, 2008; Mankouri and Hickson, 2007; Yang et al., 2010). In addition to the catalytic activities, a functional RTR complex also requires structural functions based on protein-protein interactions, for which RMI1 plays an essential role (Mullen et al., 2005; Chen and Brill, 2007; Bonnet et al., 2013; Schröpfer et al., 2014). Dissolution leads to noncross-over products and therefore, is a major mechanism in somatic yeast cells (Gangloff et al., 1994; Ira et al., 2003; Matos et al., 2011). In Arabidopsis, the loss of RTR component function leads to elevated rates of HR as well as sensitivity to UV light and methylmethane sulfonate (MMS; Bagherieh-Najjar et al., 2005; Hartung et al., 2007a; Bonnet et al., 2013). Mutants of AtRMI1 and AtTOP3A exhibit severe and unique meiotic phenotypes (Chelysheva et al., 2008; Hartung et al., 2008). This meiosis I arrest is dependent on HR, but the exact nature of the recombination intermediates that are involved remains unclear (Li et al., 2004; Hartung et al., 2007b; Knoll et al., 2014).Dissolution acts in parallel with a second pathway mediated by the structure-specific endonuclease MMS and UV-sensitive protein81 (MUS81) as shown by the fact that the additional mutation of ScSgs1/AtRECQ4A leads to synthetic lethality (Mullen et al., 2001; Hartung et al., 2006; Mannuss et al., 2010). Single mutants of MUS81 in yeast, human, Drosophila melanogaster, and Arabidopsis are sensitive to DNA-damaging agents that perturb RFs and show reduced HR after induction of double-strand breaks (Boddy et al., 2001; Hanada et al., 2006; Hartung et al., 2006). The MUS81 homologs form heterodimers with the noncatalytic subunit essential meiotic endonuclease1 (EME1; ScMms4 in S. cerevisiae). SpMus81-Eme1 was, to our knowledge, the first nuclear endonuclease reported to be capable of resolving HJs (Boddy et al., 2001). The Arabidopsis complexes can be formed with the two different subunits: AtEME1A or AtEME1B (Geuting et al., 2009). AtMUS81-EME1A/B, like the fission yeast ortholog, preferentially cleaves nicked Holliday junctions (nHJs) and 3′-flaps but also shows weaker activity on intact HJs in vitro (Boddy et al., 2001; Osman et al., 2003; Geuting et al., 2009; Schwartz and Heyer, 2011). MUS81 homologs are key players in meiotic cross-over generation (Osman et al., 2003; Berchowitz et al., 2007; Higgins et al., 2008). Although cross-over formation is solely dependent on SpMus81 in fission yeast, this function was shown to be shared with ScYen1 in budding yeast (Osman et al., 2003; Blanco et al., 2010; Ho et al., 2010; Tay and Wu, 2010). Tightly regulated by cell division cycle5-dependent hyperphosphorylation at the end of prophase I, the main activity of ScMus81-Mms4 is timed to coordinate with the formation of chiasmata and HJs that link the homologous chromosomes. This role in meiosis I is shown by the failure of chromosome segregation at the end of meiosis I in ScMus81 mutants (Matos et al., 2011). Interestingly, the chromosomes could be segregated at the end of meiosis II because of the presence of ScYen1. In contrast to canonical HJ resolvases, the hallmark of the MUS81-EME1 cleavage mechanism is the asymmetry of the second incision relative to either a first incision or a preexisting nick. This difference classifies MUS81-EME1 as a noncanonical resolvase. Its products need additional processing by gap-filling or flap-cleaving enzymes to allow religation (Boddy et al., 2001; Geuting et al., 2009).In very recent studies, HsMUS81-EME1 was found to constitute an essential canonical HJ resolvase with HsSLX1-SLX4 (SLX for synthetic lethal of unknown function), in which a first incision is made by HsSLX1-SLX4 followed by the enhanced action of the HsMUS81-EME1 subunits on the resulting nHJ (Garner et al., 2013; Wyatt et al., 2013). HsSLX1-SLX4 had previously been described as a canonical resolvase, albeit producing only a low level of symmetrically cut ligatable products (Fekairi et al., 2009).In addition to the mechanisms described above, an activity resembling that of EcRuvC had long been known to be present in mammalian cell-free extracts. In 2008, the group of Steven C. West succeeded in identifying, to their knowledge, the first nuclear proteins analogous to the EcRuvC paradigm: ScYen1 and Homo sapiens XPG-like endonuclease1 (HsGEN1; Ip et al., 2008). These proteins are members of the large and well-characterized Rad2/XPG family of nucleases. The Rad2/XPG family consists of the Xeroderma pigmentosum group G-complementing protein (XPG) endonucleases of the nucleotide excision repair (class I), the flap endonuclease1 (FEN1) replication-associated flap endonucleases (class II), the exodeoxyribonuclease1 (EXO1) exonucleases of recombination and repair (class III), and class IV (containing the [putative] eukaryotic HJ resolvases). This last class was introduced after the identification of the rice (Oryza sativa) single-strand DNA endonuclease1 (OsSEND-1) based on sequence homology. The class IV members show a domain composition homologous to FEN1 and EXO1, with no spacer region between their N-terminal XPG (XPG-N) and internal XPG (XPG-I) domains, whereas the primary structure of these domains is more similar to the sequence of the nuclease domain of XPG (Furukawa et al., 2003).Although all Rad2/XPG homologs share a common cleavage mechanism as observed for the typical 5′-flap substrate (Tsutakawa et al., 2011; Tsutakawa and Tainer, 2012), the striking evolutionary difference between classes I, II, and III on the one hand and the HJ resolvases (class IV) on the other hand is the ability of class IV members to form homodimers in vitro at their preferred substrate, the HJs (Rass et al., 2010). The homodimer configuration ensures the presence of two active sites positioned on the opposing strands of the HJ, which is necessary for resolution. The mode of eukaryotic HJ resolution is largely similar to the bacterial paradigm: (1) cleavage occurs one nucleotide in the 3′ direction of a static junction point (equivalent to the main cleavage site on 5′-flaps), (2) the incisions occur with almost perfect point symmetry, (3) the incisions result in readily ligatable nDs, and (4) certain sites within a migratable HJ core are preferred, providing evidence for a (yet to be determined) sequence specificity (Ip et al., 2008; Bailly et al., 2010; Rass et al., 2010; Yang et al., 2012).In the absence of MUS81-EME1/Mms4, the proteins HsGEN1, ScYen1, and CeGEN-1 have been shown to play a role in response to replication-associated perturbations, such as MMS- and UV-induced DNA damage (Bailly et al., 2010; Blanco et al., 2010; Tay and Wu, 2010; Gao et al., 2012; Muñoz-Galván et al., 2012). It is also likely that these proteins provide a backup mechanism in mitosis and meiosis, ensuring proper chromosome segregation after a failure of other mechanisms, including MUS81-EME1/Mms4 (Blanco et al., 2010; Matos et al., 2011).Although canonical HJ resolvases in animals and fungi are a current topic of great interest, very little is known about these proteins in plants. In rice, two members of the Rad2/XPG class IV have been described: OsSEND-1 (the founding member) and OsGEN-like (OsGEN-L). OsSEND-1 was shown to digest single-stranded circular DNA, and its expression is induced on MMS-induced genotoxic stress, whereas OsGEN-L is implicated in late spore development (Furukawa et al., 2003; Moritoh et al., 2005). Both studies (Furukawa et al., 2003; Moritoh et al., 2005) proposed putative homologs in other plants, and the gene locus At1g01880 of Arabidopsis, coding for the protein AtGEN1, is considered the ortholog of HsGEN1 and ScYen1 (Ip et al., 2008). However, currently, only OsGEN-L has been further investigated and described to possess in vitro properties similar to both Rad2/XPG nucleases and EcRuvC. This protein shows a well-defined 5′-flap activity as well as a poorly characterized ability, similar to that of EcRuvC, to resolve mobile HJs (Yang et al., 2012).Thus, of two members of Rad2/XPG class IV of plants, only one member has so far been analyzed with respect to a possible HJ resolvase activity. However, Arabidopsis expression data show that both proteins are expressed in plants and do not reveal marked differences (Laubinger et al., 2008). In this study, the goal was, therefore, to characterize the in vitro activities of not only AtGEN1 but also, AtSEND1, focusing on the idea that Arabidopsis and (seed) plants in general might encode not one but actually two HJ resolvases with functional homology to EcRuvC.  相似文献   

20.
Organelle movement and positioning play important roles in fundamental cellular activities and adaptive responses to environmental stress in plants. To optimize photosynthetic light utilization, chloroplasts move toward weak blue light (the accumulation response) and escape from strong blue light (the avoidance response). Nuclei also move in response to strong blue light by utilizing the light-induced movement of attached plastids in leaf cells. Blue light receptor phototropins and several factors for chloroplast photorelocation movement have been identified through molecular genetic analysis of Arabidopsis (Arabidopsis thaliana). PLASTID MOVEMENT IMPAIRED1 (PMI1) is a plant-specific C2-domain protein that is required for efficient chloroplast photorelocation movement. There are two PLASTID MOVEMENT IMPAIRED1-RELATED (PMIR) genes, PMIR1 and PMIR2, in the Arabidopsis genome. However, the mechanism in which PMI1 regulates chloroplast and nuclear photorelocation movements and the involvement of PMIR1 and PMIR2 in these organelle movements remained unknown. Here, we analyzed chloroplast and nuclear photorelocation movements in mutant lines of PMI1, PMIR1, and PMIR2. In mesophyll cells, the pmi1 single mutant showed severe defects in both chloroplast and nuclear photorelocation movements resulting from the impaired regulation of chloroplast-actin filaments. In pavement cells, pmi1 mutant plants were partially defective in both plastid and nuclear photorelocation movements, but pmi1pmir1 and pmi1pmir1pmir2 mutant lines lacked the blue light-induced movement responses of plastids and nuclei completely. These results indicated that PMI1 is essential for chloroplast and nuclear photorelocation movements in mesophyll cells and that both PMI1 and PMIR1 are indispensable for photorelocation movements of plastids and thus, nuclei in pavement cells.In plants, organelles move within the cell and become appropriately positioned to accomplish their functions and adapt to the environment (for review, see Wada and Suetsugu, 2004). Light-induced chloroplast movement (chloroplast photorelocation movement) is one of the best characterized organelle movements in plants (Suetsugu and Wada, 2012). Under weak light conditions, chloroplasts move toward light to capture light efficiently (the accumulation response; Zurzycki, 1955). Under strong light conditions, chloroplasts escape from light to avoid photodamage (the avoidance response; Kasahara et al., 2002; Sztatelman et al., 2010; Davis and Hangarter, 2012; Cazzaniga et al., 2013). In most green plant species, these responses are induced primarily by the blue light receptor phototropin (phot) in response to a range of wavelengths from UVA to blue light (approximately 320–500 nm; for review, see Suetsugu and Wada, 2012; Wada and Suetsugu, 2013; Kong and Wada, 2014). Phot-mediated chloroplast movement has been shown in land plants, such as Arabidopsis (Arabidopsis thaliana; Jarillo et al., 2001; Kagawa et al., 2001; Sakai et al., 2001), the fern Adiantum capillus-veneris (Kagawa et al., 2004), the moss Physcomitrella patens (Kasahara et al., 2004), and the liverwort Marchantia polymorpha (Komatsu et al., 2014). Two phots in Arabidopsis, phot1 and phot2, redundantly mediate the accumulation response (Sakai et al., 2001), whereas phot2 primarily regulates the avoidance response (Jarillo et al., 2001; Kagawa et al., 2001; Luesse et al., 2010). M. polymorpha has only one phot that mediates both the accumulation and avoidance responses (Komatsu et al., 2014), although two or more phots mediate chloroplast photorelocation movement in A. capillus-veneris (Kagawa et al., 2004) and P. patens (Kasahara et al., 2004). Thus, duplication and functional diversification of PHOT genes have occurred during land plant evolution, and plants have gained a sophisticated light sensing system for chloroplast photorelocation movement.In general, movements of plant organelles, including chloroplasts, are dependent on actin filaments (for review, see Wada and Suetsugu, 2004). Most organelles common in eukaryotes, such as mitochondria, peroxisomes, and Golgi bodies, use the myosin motor for their movements, but there is no clear evidence that chloroplast movement is myosin dependent (for review, see Suetsugu et al., 2010a). Land plants have innovated a novel actin-based motility system that is specialized for chloroplast movement as well as a photoreceptor system (for review, see Suetsugu et al., 2010a; Wada and Suetsugu, 2013; Kong and Wada, 2014). Chloroplast-actin (cp-actin) filaments, which were first found in Arabidopsis, are short actin filaments specifically localized around the chloroplast periphery at the interface between the chloroplast and the plasma membrane (Kadota et al., 2009). Strong blue light induces the rapid disappearance of cp-actin filaments and then, their subsequent reappearance preferentially at the front region of the moving chloroplasts. This asymmetric distribution of cp-actin filaments is essential for directional chloroplast movement (Kadota et al., 2009; Kong et al., 2013a). The greater the difference in the amount of cp-actin filaments between the front and rear regions of chloroplasts becomes, the faster the chloroplasts move, in which the magnitude of the difference is determined by fluence rate (Kagawa and Wada, 2004; Kadota et al., 2009; Kong et al., 2013a). Strong blue light-induced disappearance of cp-actin filaments is regulated in a phot2-dependent manner before the intensive polymerization of cp-actin filaments at the front region occurs (Kadota et al., 2009; Ichikawa et al., 2011; Kong et al., 2013a). This phot2-dependent response contributes to the greater difference in the amount of cp-actin filaments between the front and rear regions of chloroplasts. Similar behavior of cp-actin filaments has also been observed in A. capillus-veneris (Tsuboi and Wada, 2012) and P. patens (Yamashita et al., 2011).Like chloroplasts, nuclei also show light-mediated movement and positioning (nuclear photorelocation movement) in land plants (for review, see Higa et al., 2014b). In gametophytic cells of A. capillus-veneris, weak light induced the accumulation responses of both chloroplasts and nuclei, whereas strong light induced avoidance responses (Kagawa and Wada, 1993, 1995; Tsuboi et al., 2007). However, in mesophyll cells of Arabidopsis, strong blue light induced both chloroplast and nuclear avoidance responses, but weak blue light induced only the chloroplast accumulation response (Iwabuchi et al., 2007, 2010; Higa et al., 2014a). In Arabidopsis pavement cells, small numbers of tiny plastids were found and showed autofluorescence under the confocal laser-scanning microscopy (Iwabuchi et al., 2010; Higa et al., 2014a). Hereafter, the plastid in the pavement cells is called the pavement cell plastid. Strong blue light-induced avoidance responses of pavement cell plastids and nuclei were induced in a phot2-dependent manner, but the accumulation response was not detected for either organelle (Iwabuchi et al., 2007, 2010; Higa et al., 2014a). In both Arabidopsis and A. capillus-veneris, phots mediate nuclear photorelocation movement, and phot2 mediates the nuclear avoidance response (Iwabuchi et al., 2007, 2010; Tsuboi et al., 2007). The nuclear avoidance response is dependent on actin filaments in both mesophyll and pavement cells of Arabidopsis (Iwabuchi et al., 2010). Recently, it was shown that the nuclear avoidance response relies on cp-actin-dependent movement of pavement cell plastids, where nuclei are associated with pavement cell plastids of Arabidopsis (Higa et al., 2014a). In mesophyll cells, nuclear avoidance response is likely dependent on cp-actin filament-mediated chloroplast movement, because the mutants deficient in chloroplast movement were also defective in nuclear avoidance response (Higa et al., 2014a). Thus, phots mediate both chloroplast (and pavement cell plastid) and nuclear photorelocation movement by regulating cp-actin filaments.Molecular genetic analyses of Arabidopsis mutants deficient in chloroplast photorelocation movement have identified many molecular factors involved in signal transduction and/or motility systems as well as those involved in the photoreceptor system for chloroplast photorelocation movement (and thus, nuclear photorelocation movement; for review, see Suetsugu and Wada, 2012; Wada and Suetsugu, 2013; Kong and Wada, 2014). CHLOROPLAST UNUSUAL POSITIONING1 (CHUP1; Oikawa et al., 2003) and KINESIN-LIKE PROTEIN FOR ACTIN-BASED CHLOROPLAST MOVEMENT (KAC; Suetsugu et al., 2010b) are key factors for generating and/or maintaining cp-actin filaments. Both proteins are highly conserved in land plants and essential for the movement and attachment of chloroplasts to the plasma membrane in Arabidopsis (Oikawa et al., 2003, 2008; Suetsugu et al., 2010b), A. capillus-veneris (Suetsugu et al., 2012), and P. patens (Suetsugu et al., 2012; Usami et al., 2012). CHUP1 is localized on the chloroplast outer membrane and binds to globular and filamentous actins and profilin in vitro (Oikawa et al., 2003, 2008; Schmidt von Braun and Schleiff, 2008). Although KAC is a kinesin-like protein, it lacks microtubule-dependent motor activity but has filamentous actin binding activity (Suetsugu et al., 2010b). An actin-bundling protein THRUMIN1 (THRUM1) is required for efficient chloroplast photorelocation movement (Whippo et al., 2011) and interacts with cp-actin filaments (Kong et al., 2013a). chup1 and kac mutant plants were shown to lack detectable cp-actin filaments (Kadota et al., 2009; Suetsugu et al., 2010b; Ichikawa et al., 2011; Kong et al., 2013a). Similarly, cp-actin filaments were rarely detected in thrum1 mutant plants (Kong et al., 2013a), indicating that THRUM1 also plays an important role in maintaining cp-actin filaments.Other proteins J-DOMAIN PROTEIN REQUIRED FOR CHLOROPLAST ACCUMULATION RESPONSE1 (JAC1; Suetsugu et al., 2005), WEAK CHLOROPLAST MOVEMENT UNDER BLUE LIGHT1 (WEB1; Kodama et al., 2010), and PLASTID MOVEMENT IMPAIRED2 (PMI2; Luesse et al., 2006; Kodama et al., 2010) are involved in the light regulation of cp-actin filaments and chloroplast photorelocation movement. JAC1 is an auxilin-like J-domain protein that mediates the chloroplast accumulation response through its J-domain function (Suetsugu et al., 2005; Takano et al., 2010). WEB1 and PMI2 are coiled-coil proteins that interact with each other (Kodama et al., 2010). Although web1 and pmi2 were partially defective in the avoidance response, the jac1 mutation completely suppressed the phenotype of web1 and pmi2, suggesting that the WEB1/PMI2 complex suppresses JAC1 function (i.e. the accumulation response) under strong light conditions (Kodama et al., 2010). Both web1 and pmi2 showed impaired disappearance of cp-actin filaments in response to strong blue light (Kodama et al., 2010). However, the exact molecular functions of these proteins are unknown.In this study, we characterized mutant plants deficient in the PMI1 gene and two homologous genes PLASTID MOVEMENT IMPAIRED1-RELATED1 (PMIR1) and PMIR2. PMI1 was identified through molecular genetic analyses of pmi1 mutants that showed severe defects in chloroplast accumulation and avoidance responses (DeBlasio et al., 2005). PMI1 is a plant-specific C2-domain protein (DeBlasio et al., 2005; Zhang and Aravind, 2010), but its roles and those of PMIRs in cp-actin-mediated chloroplast and nuclear photorelocation movements remained unclear. Thus, we analyzed chloroplast and nuclear photorelocation movements in the single, double, and triple mutants of pmi1, pmir1, and pmir2.  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