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 359 毫秒
1.
2.
3.
4.
5.
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   

6.
7.
Most proteins produced in the endoplasmic reticulum (ER) of eukaryotic cells fold via disulfide formation (oxidative folding). Oxidative folding is catalyzed by protein disulfide isomerase (PDI) and PDI-related ER protein thiol disulfide oxidoreductases (ER oxidoreductases). In yeast and mammals, ER oxidoreductin-1s (Ero1s) supply oxidizing equivalent to the active centers of PDI. In this study, we expressed recombinant soybean Ero1 (GmERO1a) and found that GmERO1a oxidized multiple soybean ER oxidoreductases, in contrast to mammalian Ero1s having a high specificity for PDI. One of these ER oxidoreductases, GmPDIM, associated in vivo and in vitro with GmPDIL-2, was unable to be oxidized by GmERO1a. We therefore pursued the possible cooperative oxidative folding by GmPDIM, GmERO1a, and GmPDIL-2 in vitro and found that GmPDIL-2 synergistically accelerated oxidative refolding. In this process, GmERO1a preferentially oxidized the active center in the a′ domain among the a, a′, and b domains of GmPDIM. A disulfide bond introduced into the active center of the a′ domain of GmPDIM was shown to be transferred to the active center of the a domain of GmPDIM and the a domain of GmPDIM directly oxidized the active centers of both the a or a′ domain of GmPDIL-2. Therefore, we propose that the relay of an oxidizing equivalent from one ER oxidoreductase to another may play an essential role in cooperative oxidative folding by multiple ER oxidoreductases in plants.In eukaryotes, many secretory and membrane proteins fold via disulfide bond formation in the endoplasmic reticulum (ER). Seed storage proteins of major crops, such as wheat, corn, rice, and beans, which are important protein sources for humans and domestic animals, are synthesized in the ER of the endosperm or cotyledon. A number of seed storage proteins fold by the formation of intramolecular disulfide bonds (oxidative folding) and are transported to and accumulate in protein bodies (Kermode and Bewley, 1999; Jolliffe et al., 2005). In contrast to normally folded proteins, misfolded and unfolded proteins are retained in the ER and degraded by an ER-associated degradation or vacuolar system (Smith et al., 2011; Pu and Bassham, 2013). Therefore, quick and efficient oxidative folding of nascent seed storage proteins is needed for their accumulation in protein bodies.During this process, protein disulfide isomerase (PDI; EC 5.3.4.1) and other ER protein thiol disulfide oxidoreductases (ER oxidoreductases) are thought to catalyze the formation and isomerization of disulfide bonds in nascent proteins (Hatahet and Ruddock, 2009; Feige and Hendershot, 2011; Lu and Holmgren, 2014). After phylogenetic analysis of the Arabidopsis genome, 10 classes of ER oxidoreductases (classes I–X) were identified (Houston et al., 2005). Among them, class I ER oxidoreductase, a plant PDI ortholog, has been studied in a wide variety of plants. Class I ER oxidoreductases have two catalytically active domains a and a′, containing active centers composed of Cys-Gly-His-Cys and two catalytically inactive domains b and b′. An Arabidopsis ortholog of class I ER oxidoreductases is required for proper seed development and regulates the timing of programmed cell death by chaperoning and inhibiting Cys proteases (Andème Ondzighi et al., 2008). OaPDI, a PDI from Oldenlandia affinis, a coffee family (Rubiaceae) plant, is involved in the folding of knotted circular proteins (Gruber et al., 2007). The rice ortholog (PDIL1-1) was suggested to be involved in the maturation of the major seed storage protein glutelin (Takemoto et al., 2002). Furthermore, rice PDIL1-1 plays a role in regulatory activities for various proteins that are essential for the synthesis of grain components as determined by analysis of a T-DNA insertion mutant (Satoh-Cruz et al., 2010).The oxidative refolding ability of class I ER oxidoreductases was confirmed in recombinant soybean (GmPDIL-1) and wheat proteins produced by an Escherichia coli expression system established from cDNAs (Kamauchi et al., 2008; Kimura et al., 2015).Class II and III ER oxidoreductases have an a–b–b′–a′ domain structure. Class II ER oxidoreductases have an acidic amino acid-rich sequence in the N-terminal region ahead of the a domain. Recombinant soybean (GmPDIL-2) and wheat class II ER oxidoreductases have oxidative refolding activities similar to that of class I (Kamauchi et al., 2008; Kimura et al., 2015). Class III ER oxidoreductases contain the nonclassical redox-center Cys-X-X-Ser/Cys motifs, as opposed to the more traditional CGHC sequence, in the a and a′ domains. Recombinant soybean (GmPDIL-3) and wheat proteins lack oxidative refolding activity in vitro (Iwasaki et al., 2009; Kimura et al., 2015). Class IV ER oxidoreductases are unique to plants and have an a–a′–ERp29 domain structure, which is homologous to the C-terminal domain of mammalian ERp29 (Demmer et al., 1997).Recombinant soybean class IV ER oxidoreductases (GmPDIS-1 and GmPDIS-2) and wheat class IV ER oxidoreductase possess an oxidative refolding activity that is weaker than that of classes I and II (Wadahama et al., 2007; Kimura et al., 2015). Class V ER oxidoreductases are plant orthologs of mammalian P5 and have an a–a′–b domain structure. A rice class V ER oxidoreductase, consisting of PDIL2 and PDIL3, plays an important role in the accumulation of the seed storage protein Cys-rich 10-kD prolamin (crP10; Onda et al., 2011). Recombinant soybean class V ER oxidoreductase, GmPDIM and wheat class V ER oxidoreductase possess an oxidative refolding activity similar to that of class IV (Wadahama et al., 2008; Kimura et al., 2015). In the soybean, GmPDIL-1, GmPDIL-2, GmPDIM, GmPDIS-1, and GmPDIS-2 were found to associate transiently with a seed storage precursor protein, proglycinin, in the ER of the cotyledon by coimmunoprecipitation experiments, suggesting that multiple ER oxidoreductases are involved in the folding of the nascent proglycinin.The disulfide bond in the active center of ER oxidoreductases is reduced as a result of catalyzing disulfide bond formation in an unfolded protein. The reduced active center of PDI was discovered to be oxidized again by ER oxidoreductin-1 (Ero1p) in yeast (Frand and Kaiser, 1998; Pollard et al., 1998). Ero1p orthologs are present universally in eukaryotes. Yeast and flies have a single copy of the ERO1 gene, which is essential for survival (Frand and Kaiser, 1998; Pollard et al., 1998; Tien et al., 2008). Mammals have two genes encoding Ero1-α (Cabibbo et al., 2000) and Ero1-β (Pagani et al., 2000) that function as major disulfide donors to nascent proteins in the ER, but are not critical for survival (Zito et al., 2010). Domain a of yeast PDI is the most favored substrate of yeast Ero1p (Vitu et al., 2010), whereas a′ of human PDI is specifically oxidized by human Ero1-α (Chambers et al., 2010) and Ero1-β (Wang et al., 2011). Electrons from Cys residues of the active centers of PDI are transferred to oxygen by Ero1 (Tu and Weissman, 2004; Sevier and Kaiser, 2008). The reaction mechanisms of yeast Ero1p and human Ero1s have been intensively investigated; their regulation by PDI has been extensively studied as well (Tavender and Bulleid, 2010; Araki and Inaba, 2012; Benham et al., 2013; Ramming et al., 2015). Only rice Ero1 (OsERO1) has been identified as a plant ortholog of Ero1p (Onda et al., 2009). OsERO1 is necessary for disulfide bond formation in rice endosperm. The formation of native disulfide bonds in the major seed storage protein proglutelin was demonstrated to depend upon OsERO1 by RNAi knockdown experiments. However, no plant protein thiol disulfide oxidoreductases that are oxidized by a plant Ero1 ortholog have been identified to date.In this study, we show that multiple soybean ER oxidoreductases can be activated by a soybean Ero1 ortholog (GmERO1a). In addition, we propose a synergistic mechanism by which GmPDIM and GmPDIL-2 cooperatively fold unfolded proteins using oxidizing equivalents provided by GmERO1 in vitro.  相似文献   

8.
9.
10.
Effective grain filling is one of the key determinants of grain setting in rice (Oryza sativa). Grain setting defect1 (GSD1), which encodes a putative remorin protein, was found to affect grain setting in rice. Investigation of the phenotype of a transfer DNA insertion mutant (gsd1-Dominant) with enhanced GSD1 expression revealed abnormalities including a reduced grain setting rate, accumulation of carbohydrates in leaves, and lower soluble sugar content in the phloem exudates. GSD1 was found to be specifically expressed in the plasma membrane and plasmodesmata (PD) of phloem companion cells. Experimental evidence suggests that the phenotype of the gsd1-Dominant mutant is caused by defects in the grain-filling process as a result of the impaired transport of carbohydrates from the photosynthetic site to the phloem. GSD1 functioned in affecting PD conductance by interacting with rice ACTIN1 in association with the PD callose binding protein1. Together, our results suggest that GSD1 may play a role in regulating photoassimilate translocation through the symplastic pathway to impact grain setting in rice.Grain filling, a key determinant of grain yield in rice (Oryza sativa), hinges on the successful translocation of photoassimilates from the leaves to the fertilized reproductive organs through the phloem transport system. Symplastic phloem loading, which is one of the main pathways responsible for the transport of photoassimilates in rice, is mediated by plasmodesmata (PD) that connect phloem companion cells with sieve elements and surrounding parenchyma cells (Kaneko et al., 1980; Chonan et al., 1981; Eom et al., 2012). PD are transverse cell wall channels structured with the cytoplasmic sleeve and the modified endoplasmic reticulum desmotubule between neighboring cells (Maule, 2008). A number of proteins affect the structure and functional performance of the PD, which in turn impacts the cell-to-cell transport of small and large molecules through the PD during plant growth, development, and defense (Cilia and Jackson, 2004; Sagi et al., 2005; Lucas et al., 2009; Simpson et al., 2009; Stonebloom et al., 2009). For example, actin and myosin, which link the desmotubule to the plasma membrane (PM) at the neck region of PD, are believed to play a role in regulating PD permeability by controlling PD aperture (White et al., 1994; Ding et al., 1996; Reichelt et al., 1999). Callose deposition can also impact the size of the PD aperture at the neck region (Radford et al., 1998; Levy et al., 2007) and callose synthase genes such as Glucan Synthase-Like7 (GSL7, also named CalS7), GSL8, and GSL12 have been shown to play a role in regulating symplastic trafficking (Guseman et al., 2010; Barratt et al., 2011; Vatén et al., 2011; Xie et al., 2011). Other proteins that have been shown to impact the structure and function of the PD include glycosylphosphatidylinositol (GPI)-anchored proteins, PD callose binding protein1 (PDCB1), which is also associated with callose deposition (Simpson et al., 2009), and LYSIN MOTIF DOMAIN-CONTAINING GLYCOSYLPHOSPHATIDYLINOSITOL-ANCHORED PROTEIN2, which limits the molecular flux through the PD by chitin perception (Faulkner et al., 2013). Changes in PD permeability can have major consequences for the translocation of photoassimilates needed for grain filling in rice. However, the genes and molecular mechanisms underlying the symplastic transport of photoassimilates remain poorly characterized.Remorins are a diverse family of plant-specific proteins with conserved C-terminal sequences and highly variable N-terminal sequences. Remorins can be classified into six distinct phylogenetic groups (Raffaele et al., 2007). The functions of most remorins are unknown, but some members of the family have been shown to be involved in immune response through controlling the cell-to-cell spread of microbes. StREM1.3, a remorin that is located in PM rafts and the PD, was shown to impair the cell-to-cell movement of a plant virus X by binding to Triple Gene Block protein1 (Raffaele et al., 2009). Medicago truncatula symbiotic remorin1 (MtSYMREM1), a remorin located at the PM in Medicago truncatula, was shown to facilitate infection and the release of rhizobial bacteria into the host cytoplasm (Lefebvre et al., 2010). Overexpression of LjSYMREM1, the ortholog of MtSYMREM1 in Lotus japonicus, resulted in increased root nodulation (Lefebvre et al., 2010; Tóth et al., 2012). Although a potential association between remorins and PD permeability has been proposed (Raffaele et al., 2009), the diversity observed across remorins, plus the fact that remorin mutants generated through different approaches fail to show obvious phenotypes (Reymond et al., 1996; Bariola et al., 2004), have made it challenging to characterize the function of remorins in cell-to-cell transport.In this study, we identified a rice transfer DNA (T-DNA) insertion mutant (grain setting defect1-Dominant [gsd1-D]), with a grain setting-deficient phenotype caused by overexpression of GSD1, a remorin gene with unknown function. GSD1 is expressed specifically in phloem companion cells and is localized in the PD and PM. We provide evidence to show that overexpression of GSD1 leads to deficient grain setting in rice, likely as a consequence of reduced sugar transport resulting from decreased PD permeability in phloem companion cells.  相似文献   

11.
12.
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   

13.
14.
The actin cytoskeleton is a major regulator of cell morphogenesis and responses to biotic and abiotic stimuli. The organization and activities of the cytoskeleton are choreographed by hundreds of accessory proteins. Many actin-binding proteins are thought to be stimulus-response regulators that bind to signaling phospholipids and change their activity upon lipid binding. Whether these proteins associate with and/or are regulated by signaling lipids in plant cells remains poorly understood. Heterodimeric capping protein (CP) is a conserved and ubiquitous regulator of actin dynamics. It binds to the barbed end of filaments with high affinity and modulates filament assembly and disassembly reactions in vitro. Direct interaction of CP with phospholipids, including phosphatidic acid, results in uncapping of filament ends in vitro. Live-cell imaging and reverse-genetic analyses of cp mutants in Arabidopsis (Arabidopsis thaliana) recently provided compelling support for a model in which CP activity is negatively regulated by phosphatidic acid in vivo. Here, we used complementary biochemical, subcellular fractionation, and immunofluorescence microscopy approaches to elucidate CP-membrane association. We found that CP is moderately abundant in Arabidopsis tissues and present in a microsomal membrane fraction. Sucrose density gradient separation and immunoblotting with known compartment markers were used to demonstrate that CP is enriched on membrane-bound organelles such as the endoplasmic reticulum and Golgi. This association could facilitate cross talk between the actin cytoskeleton and a wide spectrum of essential cellular functions such as organelle motility and signal transduction.The cellular levels of membrane-associated lipids undergo dynamic changes in response to developmental and environmental stimuli. Different species of phospholipids target specific proteins and this often affects the activity and/or subcellular localization of these lipid-binding proteins. One such membrane lipid, phosphatidic acid (PA), serves as a second messenger and regulates multiple developmental processes in plants, including seedling development, root hair growth and pattern formation, pollen tube growth, leaf senescence, and fruit ripening. PA levels also change during various stress responses, including high salinity and dehydration, pathogen attack, and cold tolerance (Testerink and Munnik, 2005, 2011; Wang, 2005; Li et al., 2009). In mammalian cells, PA is critical for vesicle trafficking events, such as vesicle budding from the Golgi apparatus, vesicle transport, exocytosis, endocytosis, and vesicle fusion (Liscovitch et al., 2000; Freyberg et al., 2003; Jenkins and Frohman, 2005).The actin cytoskeleton and a plethora of actin-binding proteins (ABPs) are well-known targets and transducers of lipid signaling (Drøbak et al., 2004; Saarikangas et al., 2010; Pleskot et al., 2013). For example, several ABPs have the ability to bind phosphoinositide lipids, such as phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2]. The severing or actin filament depolymerizing proteins such as villin, cofilin, and profilin are inhibited when bound to PtdIns(4,5)P2. One ABP appears to be strongly regulated by another phospholipid; human gelsolin binds to lysophosphatidic acid and its filament severing and barbed-end capping activities are inhibited by this biologically active lipid (Meerschaert et al., 1998). Gelsolin is not, however, regulated by PA (Meerschaert et al., 1998), nor are profilin (Lassing and Lindberg, 1985), α-actinin (Fraley et al., 2003), or chicken CapZ (Schafer et al., 1996).The heterodimeric capping protein (CP) from Arabidopsis (Arabidopsis thaliana) also binds to and its activity is inhibited by phospholipids, including both PtdIns(4,5)P2 and PA (Huang et al., 2003, 2006). PA and phospholipase D activity have been implicated in the actin-dependent tip growth of root hairs and pollen tubes (Ohashi et al., 2003; Potocký et al., 2003; Samaj et al., 2004; Monteiro et al., 2005a; Pleskot et al., 2010). Exogenous application of PA causes an elevation of actin filament levels in suspension cells, pollen, and Arabidopsis epidermal cells (Lee et al., 2003; Potocký et al., 2003; Huang et al., 2006; Li et al., 2012; Pleskot et al., 2013). Capping protein (CP) binds to the barbed end of actin filaments with high (nanomolar) affinity, dissociates quite slowly, and prevents the addition of actin subunits at this end (Huang et al., 2003, 2006; Kim et al., 2007). In the presence of phospholipids, AtCP is not able to bind to the barbed end of actin filaments (Huang et al., 2003, 2006). Furthermore, capped filament ends are uncapped by the addition of PA, allowing actin assembly from a pool of profilin-actin (Huang et al., 2006). Collectively, these data lead to a simple model whereby CP, working in concert with profilin-actin, serves to maintain tight regulation of actin assembly at filament barbed ends (Huang et al., 2006; Blanchoin et al., 2010; Henty-Ridilla et al., 2013; Pleskot et al., 2013). Furthermore, the availability of CP for filament ends can be modulated by fluxes in signaling lipids. Genetic evidence for this model was recently obtained by analyzing the dynamic behavior of actin filament ends in living Arabidopsis epidermal cells after treatment with exogenous PA (Li et al., 2012). Specifically, changes in the architecture of cortical actin arrays and dynamics of individual actin filaments that are induced by PA treatment were found to be attenuated in cp mutant cells (Li et al., 2012; Pleskot et al., 2013).Structural characterization of chicken CapZ demonstrates that the α- and β-subunits of the heterodimer form a compact structure resembling a mushroom with pseudo-two-fold rotational symmetry (Yamashita et al., 2003). Actin- and phospholipid-binding sites are conserved on the C-terminal regions, sometimes referred to as tentacles, which comprise amphipathic α-helices (Cooper and Sept, 2008; Pleskot et al., 2012). Coarse-grained molecular dynamics (CG-MD) simulations recently revealed the mechanism of chicken and AtCP association with membranes (Pleskot et al., 2012). AtCP interacts specifically with lipid bilayers through interactions between PA and the amphipathic helix of the α-subunit tentacle. Extensive polar contacts between lipid headgroups and basic residues on CP (including K278, which is unique to plant CP), as well as partial embedding of nonpolar groups into the lipid bilayer, are observed (Pleskot et al., 2012). Moreover, a glutathione S-transferase fusion protein containing the C-terminal 38 amino acids from capping protein α subunit (CPA) is sufficient to bind PA-containing liposomes in vitro (Pleskot et al., 2012). Collectively, these findings lead us to predict that AtCP will behave like a membrane-associated protein in plant cells.Additional evidence from animal and microbial cells supports the association of CP with biological membranes. In Acanthamoeba castellanii, CP is localized primarily to the hyaline ectoplasm in a region of the cytoplasm just under the plasma membrane that contains a high concentration of actin filaments (Cooper et al., 1984). Localization of CP with regions rich in actin filaments and with membranes was supported by subcellular fractionation experiments, in which CP was associated with a crude membrane fraction that included plasma membrane (Cooper et al., 1984). Further evidence demonstrates that CP localizes to cortical actin patches at sites of new cell wall growth in budding yeast (Saccharomyces cerevisiae), including the site of bud emergence. By contrast, CP did not colocalize with actin cables in S. cerevisiae (Amatruda and Cooper, 1992). CP may localize to these sites by direct interactions with membrane lipids, through binding the ends of actin filaments, or by association with another protein different from actin. In support of this hypothesis, GFP-CP fusion proteins demonstrate that sites of actin assembling in living cells contain both CP and the actin-related protein2/3 (Arp2/3) complex, and CP is located in two types of structures: (1) motile regions of the cell periphery, which reflect movement of the edge of the lamella during extension and ruffling; and (2) dynamic spots within the lamella (Schafer et al., 1998). CP has been colocalized to the F-actin patches in fission yeast (Schizosaccharomyces pombe; Kovar et al., 2005), which promotes Arp2/3-dependent nucleation and branching and limits the extent of filament elongation (Akin and Mullins, 2008). These findings lend additional support for a model whereby CP cooperates with the Arp2/3 complex to regulate actin dynamics (Nakano and Mabuchi, 2006). Activities and localization of other plant ABPs are linked to membranes. Membrane association has been linked to the assembly status of the ARP2/3 complex, an actin filament nucleator, in Arabidopsis (Kotchoni et al., 2009). SPIKE1 (SPK1), a Rho of plants (Rop)-guanine nucleotide exchange factor (GEF) and peripheral membrane protein, maintains the homeostasis of the early secretory pathway and signal integration during morphogenesis through specialized domains in the endoplasmic reticulum (ER; Zhang et al., 2010). Furthermore, Nck-associated protein1 (NAP1), a component of the suppressor of cAMP receptor/WASP-family verprolin homology protein (SCAR/WAVE) complex, strongly associates with membranes and is particularly enriched in ER membranes (Zhang et al., 2013a). Finally, a superfamily of plant ABPs, called NETWORKED proteins, was recently discovered; these link the actin cytoskeleton to various cellular membranes (Deeks et al., 2012; Hawkins et al., 2014; Wang et al., 2014).In this work, we demonstrate that CP is a membrane-associated protein in Arabidopsis. To our knowledge, this is the first direct evidence for CP-membrane association in plants. This interaction likely targets CP to cellular compartments such as the ER and Golgi. This unique location may allow CP to remodel the actin cytoskeleton in the vicinity of endomembrane compartments and/or to respond rapidly to fluxes in signaling lipids.  相似文献   

15.
16.
17.
18.
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   

19.
20.
Like many other viruses, Tobacco mosaic virus replicates in association with the endoplasmic reticulum (ER) and exploits this membrane network for intercellular spread through plasmodesmata (PD), a process depending on virus-encoded movement protein (MP). The movement process involves interactions of MP with the ER and the cytoskeleton as well as its targeting to PD. Later in the infection cycle, the MP further accumulates and localizes to ER-associated inclusions, the viral factories, and along microtubules before it is finally degraded. Although these patterns of MP accumulation have been described in great detail, the underlying mechanisms that control MP fate and function during infection are not known. Here, we identify CELL-DIVISION-CYCLE protein48 (CDC48), a conserved chaperone controlling protein fate in yeast (Saccharomyces cerevisiae) and animal cells by extracting protein substrates from membranes or complexes, as a cellular factor regulating MP accumulation patterns in plant cells. We demonstrate that Arabidopsis (Arabidopsis thaliana) CDC48 is induced upon infection, interacts with MP in ER inclusions dependent on the MP N terminus, and promotes degradation of the protein. We further provide evidence that CDC48 extracts MP from ER inclusions to the cytosol, where it subsequently accumulates on and stabilizes microtubules. We show that virus movement is impaired upon overexpression of CDC48, suggesting that CDC48 further functions in controlling virus movement by removal of MP from the ER transport pathway and by promoting interference of MP with microtubule dynamics. CDC48 acts also in response to other proteins expressed in the ER, thus suggesting a general role of CDC48 in ER membrane maintenance upon ER stress.Plant viruses are obligate intracellular pathogens that replicate in association with host membranes (Laliberté and Sanfaçon, 2010) and subvert host intra- and intercellular trafficking pathways to achieve cell-to-cell and systemic spread (Harries and Ding, 2011; Niehl and Heinlein, 2011). In the case of the well-studied Tobacco mosaic virus (TMV), viral replication factories form on membranes of the endoplasmic reticulum (ER; Heinlein et al., 1995, 1998). As the plant ER is continuous between cells through plasmodesmata (PD; Ding et al., 1992), this membrane network provides a direct pathway for the spread of replicated virus from the replication sites in infected cells into the ER network of noninfected cells. The spread of plant viruses depends on virus-encoded movement proteins (MPs; Deom et al., 1987; Lucas, 2006). The MP of TMV facilitates the cell-to-cell passage of the infectious particle by forming a ribonucleoprotein complex with the viral RNA (Citovsky et al., 1990) and by increasing the size exclusion limit of PD (Wolf et al., 1989).During the course of infection, as well as when ectopically expressed, the MP associates with PD, the ER/actin network, and microtubules (Heinlein et al., 1995, 1998; Reichel and Beachy, 1998; Wright et al., 2007; Sambade et al., 2008; Hofmann et al., 2009; Boutant et al., 2010; Peña and Heinlein, 2012; Supplemental Fig. S1). Shortly after infection of a new cell, the MP localizes to small, mobile, ER-associated particles proposed to play a role in PD targeting of the viral RNA (Boyko et al., 2007; Sambade et al., 2008). Similar small, mobile MP particles are observed early upon ectopic expression of the protein. These particles colocalize with RNA and undergo stop-and-go movements in association with the ER (Sambade et al., 2008). The particle movements pause at microtubule proximal sites and their detachment requires microtubule polymerization (Sambade et al., 2008). These observations suggest that the interaction with the microtubule system plays a critical role in the maturation and ER-mediated delivery of infectious viral RNA particles to PD during early infection stages. Consistently, tobacco (Nicotiana tabacum) mutants with reduced microtubule dynamics exhibit reduced TMV movement (Ouko et al., 2010). Following virus movement, the previously infected cell further accumulates MP at the ER, a process that coincides with the formation of large ER inclusions that contain viral replicase and viral RNA in addition to MP and likely function as virus factories (Heinlein et al., 1998; Más and Beachy, 1999). In mature form, these inclusions may represent the so-called viroplasms or X-bodies described in the classical literature (Bawden and Sheffield, 1939; Esau and Cronshaw, 1967; Hills et al., 1987). Their formation is associated with rearrangements of the ER membrane and likely mediated by the accumulated MP since the inclusions diminish and reconstitute a native ER structure when MP becomes degraded by the 26S proteasome (Reichel and Beachy, 1998, 2000). Transfected cells accumulate MP in similar inclusions as those formed during infection, indicating that accumulated MP is indeed necessary and sufficient to form inclusions in association with the ER (Reichel and Beachy, 1998; Supplemental Fig. S1). Following accumulation of MP in virus factories, the infected cells accumulate the MP also along microtubules (Heinlein et al., 1998). The accumulation of MP in virus factories and on microtubules in cells behind the leading front of infection is dispensable for virus movement (Heinlein et al., 1998; Boyko et al., 2000a). At these late infection stages, the virus factories may enable the virus to produce high virion titers (Laliberté and Sanfaçon, 2010; Tilsner et al., 2012), and the subsequent accumulation along microtubules may play a role in withdrawing MP from the cell-to-cell communication pathway (Curin et al., 2007) and in stockpiling MP prior to degradation (Padgett et al., 1996; Gillespie et al., 2002).The molecular mechanisms that guide the MP to the ER and subsequently to microtubules during infection are not known. The MP is a hydrophobic protein that behaves like a membrane-integral or tightly membrane-associated protein in differential fractionation experiments and contains two predicted transmembrane domains (Reichel and Beachy, 1998; Brill et al., 2000, 2004) involved in ER association (Fujiki et al., 2006). The association with microtubules depends on MP amino acids 1 to 213 required for MP function (Kahn et al., 1998; Boyko et al., 2000b,Boyko et al., 2000c, 2002; Kotlizky et al., 2001). Moreover, certain amino acid exchange mutations known to affect the function of MP in virus movement in a temperature-sensitive manner also affect the ability of MP to interact with microtubules (Boyko et al., 2007,Boyko et al., 2000b). Interestingly, these mutations cluster together in a short domain of 25 amino acids showing a structural similarity with the M-loop of tubulin involved in tubulin-tubulin interactions (Boyko et al., 2000b; Waigmann et al., 2007). Importantly, this M-loop similarity domain overlaps with the predicted transmembrane domain (Brill et al., 2000, 2004) thus suggesting that the association of MP with membranes or microtubules is an alternative event that may depend on specific posttranslational modifications or specific folds of MP. However, although the different subcellular localizations of MPs during the course of infection indicate directional transport of MP from the ER to microtubules and may indicate different folds and functions of the protein when associated with these different subcellular components, the mechanism that controls the subcellular localization and, thus, the fate and function of MP is not known.Here, we identify CELL-DIVISION-CYCLE protein48 (CDC48), named p97/VCP (Valosin-containing protein) in mammals and Cdc48p in yeast (Saccharomyces cerevisiae), as a cellular factor regulating MP subcellular accumulation patterns. CDC48 functions are well characterized in mammalian and yeast systems but remain poorly investigated in plants. Yeast and mammalian CDC48s are essential, conserved chaperones involved in diverse cellular processes by controlling protein fate through extraction of substrates from membranes or complexes (Tsai et al., 2002; Meusser et al., 2005; Römisch, 2005; Rumpf and Jentsch, 2006; Schrader et al., 2009; Eisele et al., 2010; Meyer et al., 2012; Yamanaka et al., 2012). We show that virus infection leads to the induction of Arabidopsis (Arabidopsis thaliana) CDC48 isoforms and demonstrate a function of CDC48 in ER maintenance upon ER stress conditions. We further demonstrate that CDC48 interacts with MP and that CDC48 activity is required for MP degradation. Interaction of CDC48 with MP depends on the MP N terminus, which is required for degradation of the protein, for PD localization and microtubule accumulation of MP, and for function of MP in cell-to-cell transport of the viral RNA. Overexpressed CDC48 shifts MP subcellular localization from ER inclusions to microtubules, suggesting that CDC48 extracts the MP from ER-associated inclusions, where it accumulates in midstages of infection, to the cytosol, where it accumulates along microtubules during late infection stages. Moreover, overexpression of active, but not inactive, CDC48 inhibits virus movement. Our data demonstrate that a CDC48-dependent pathway leading to the clearance of ER-associated protein inclusions exists in plants, that plant viral MPs are substrates for this pathway, and that this pathway determines viral protein fate during infection. We suggest that CDC48-mediated extraction of MP from the ER is part of a plant defense response to remove MP from the ER, the compartment the virus uses for replication and movement.  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