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 593 毫秒
1.
BAX INHIBITOR-1 (BI-1) is a cell death suppressor widely conserved in plants and animals. Overexpression of BI-1 enhances tolerance to stress-induced cell death in plant cells, although the molecular mechanism behind this enhancement is unclear. We recently found that Arabidopsis (Arabidopsis thaliana) BI-1 is involved in the metabolism of sphingolipids, such as the synthesis of 2-hydroxy fatty acids, suggesting the involvement of sphingolipids in the cell death regulatory mechanism downstream of BI-1. Here, we show that BI-1 affects cell death-associated components localized in sphingolipid-enriched microdomains of the plasma membrane in rice (Oryza sativa) cells. The amount of 2-hydroxy fatty acid-containing glucosylceramide increased in the detergent-resistant membrane (DRM; a biochemical counterpart of plasma membrane microdomains) fraction obtained from BI-1-overexpressing rice cells. Comparative proteomics analysis showed quantitative changes of DRM proteins in BI-1-overexpressing cells. In particular, the protein abundance of FLOTILLIN HOMOLOG (FLOT) and HYPERSENSITIVE-INDUCED REACTION PROTEIN3 (HIR3) markedly decreased in DRM of BI-1-overexpressing cells. Loss-of-function analysis demonstrated that FLOT and HIR3 are required for cell death by oxidative stress and salicylic acid, suggesting that the decreased levels of these proteins directly contribute to the stress-tolerant phenotypes in BI-1-overexpressing rice cells. These findings provide a novel biological implication of plant membrane microdomains in stress-induced cell death, which is negatively modulated by BI-1 overexpression via decreasing the abundance of a set of key proteins involved in cell death.BAX INHIBITOR-1 (BI-1) is an endoplasmic reticulum (ER)-based cell death suppressor widely conserved in plants and animals (Xu and Reed, 1998; Kawai et al., 1999). In plants, BI-1 is considered a stress-associated factor, since its expression is stimulated by various stresses (Sanchez et al., 2000; Kawai-Yamada et al., 2001; Matsumura et al., 2003; Watanabe and Lam, 2006; Isbat et al., 2009). Although plants lack the homolog of animal BAX as an inducer of programmed cell death, loss of BI-1 expression results in a severe cell death phenotype under stress conditions, such as fumonisin B1-induced ER stress and disturbance of ion homeostasis (Watanabe and Lam, 2006; Ihara-Ohori et al., 2007). Conversely, plants overexpressing BI-1 exhibit tolerance to cell death induced by various stresses (Kawai-Yamada et al., 2001, 2004; Matsumura et al., 2003; Ihara-Ohori et al., 2007; Watanabe and Lam, 2008; Ishikawa et al., 2010). Moreover, BI-1 overexpression confers not only tolerance to oxidative stress-mediated cell death but also enhanced metabolic acclimation involved in energy and redox balance (Ishikawa et al., 2010). The results of these studies indicate that plant BI-1 is potentially useful for engineering stress-tolerant plants. However, little is known about the mode of action of BI-1 in the cell death regulatory pathway (Ishikawa et al., 2011). While overexpression systems sometimes include artificial or off-site effects, the observation that BI-1 overexpression improves stress tolerance suggests the importance of dissecting plants overexpressing it to further address the molecular basis of BI-1 function and cell death and stress tolerance management.As another approach to understand the molecular function of BI-1, screening of candidates interacting biochemically or functionally with BI-1 has been performed. First, Arabidopsis (Arabidopsis thaliana) BI-1 was confirmed to bind to calmodulin, like barley (Hordeum vulgare) MLO protein, a membrane-bound cell death regulator (Kim et al., 2002; Ihara-Ohori et al., 2007). Since the calmodulin-binding ability of BI-1 and MLO is necessary for their cell death-suppressing activity, Ca2+ signaling is critically involved in BI-1- and MLO-mediated cell death regulation (Kim et al., 2002; Kawai-Yamada et al., 2009). More recently, it was also demonstrated that the cell death suppression by BI-1 is mediated, at least in part, through fatty acid hydroxylase (FAH) in a Saccharomyces cerevisiae ectopic expression system (Nagano et al., 2009). In addition, Arabidopsis FAHs (AtFAH1 and AtFAH2) interact with BI-1 via cytochrome b5 at the ER, resulting in the accumulation of 2-hydroxy fatty acids (2-HFAs) in Arabidopsis plants overexpressing BI-1. 2-HFAs are typical components of the ceramide backbone of sphingolipids (Imai et al., 1995; Pata et al., 2010). Although many functions of plant sphingolipids remain to be elucidated, accumulating evidence clearly indicates that sphingolipids and their metabolism are closely involved in cell death regulation and various stress responses in plants (Ng et al., 2001; Liang et al., 2003; Townley et al., 2005; Chen et al., 2008, 2012; Wang et al., 2008; Saucedo-García et al., 2011; Dutilleul et al., 2012; Kӧnig et al., 2012; Nagano et al., 2012; Mortimer et al., 2013), implying that BI-1 plays a role in cell death regulation through sphingolipid metabolism. Sphingolipids are major components of membrane lipids and are at particularly high concentrations in membrane microdomains, known as lipid rafts in animal cells, which are essential for membrane-mediated signaling and act as a sorting platform for targeted protein traffic (Simons and Toomre, 2000; Staubach and Hanisch, 2011). In mammalian cells, sphingomyelin metabolism in lipid rafts plays a vital role in the initiation of apoptotic cell death (Milhas et al., 2010). Recent studies have demonstrated the presence of raft-like membrane microdomains in plant cells and a role for them in defense responses and targeted protein sorting (Peskan et al., 2000; Fujiwara et al., 2009; Minami et al., 2009; Melser et al., 2010; Markham et al., 2011).This study focused on membrane microdomains in relation to BI-1-mediated sphingolipid metabolism. Our findings indicated that BI-1 alters sphingolipid composition in membrane microdomains, and this is accompanied by dynamic changes in a number of detergent-resistant membrane (DRM) proteins involved in cell death regulation.  相似文献   

2.
Lipid mixtures within artificial membranes undergo a separation into liquid-disordered and liquid-ordered phases. However, the existence of this segregation into microscopic liquid-ordered phases has been difficult to prove in living cells, and the precise organization of the plasma membrane into such phases has not been elucidated in plant cells. We developed a multispectral confocal microscopy approach to generate ratiometric images of the plasma membrane surface of Bright Yellow 2 tobacco (Nicotiana tabacum) suspension cells labeled with an environment sensitive fluorescent probe. This allowed the in vivo characterization of the global level of order of this membrane, by which we could demonstrate that an increase in its proportion of ordered phases transiently occurred in the early steps of the signaling triggered by cryptogein and flagellin, two elicitors of plant defense reactions. The use of fluorescence recovery after photobleaching revealed an increase in plasma membrane fluidity induced by cryptogein, but not by flagellin. Moreover, we characterized the spatial distribution of liquid-ordered phases on the membrane of living plant cells and monitored their variations induced by cryptogein elicitation. We analyze these results in the context of plant defense signaling, discuss their meaning within the framework of the “membrane raft” hypothesis, and propose a new mechanism of signaling platform formation in response to elicitor treatment.The adaptive capacity of biological membranes is a primary determinant of cell survival in fluctuating conditions. In particular, membrane physical properties are adjusted in the perception of and response to environmental modifications (including temperature, mechanical, and osmotic stresses) in various organisms (Los and Murata, 2004; Vígh et al., 2007; Verstraeten et al., 2010), including plants (Vaultier et al., 2006; Königshofer et al., 2008). Moreover, it has been shown that modifications of plasma membrane (PM) physical properties induced by pharmacological treatments can trigger signaling events in tobacco (Nicotiana tabacum) suspension cells (Bonneau et al., 2010). This reinforces the need to analyze the relationships between membrane organization and signaling in greater detail.Fluidity, a physical property of the PM, is a measure of the rotational and translational motions of molecules within the membrane, and consequently this reflects the level of lipid order in the bilayer. Lipid order is comprised of structure, microviscosity, and membrane phase; the latter feature includes lipid shape, packing, and curvature (Rilfors et al., 1984; van der Meer et al., 1984; Bloom et al., 1991). Lipid self-association induces a physical segregation into lipid bilayers, wherein a liquid-ordered (Lo) phase coexists with a liquid-disordered (Ld) phase (Veatch and Keller, 2005; Gaus et al., 2006; Klymchenko et al., 2009; Heberle et al., 2010). The Lo phase couples a high rotational mobility with a high conformational order in the lipid acyl chain, two physical properties that could be spatially resolved by fluorescence microscopy (Kubiak et al., 2011). Moreover, some observations indicate that Lo size or proportion could be controlled by temperature or cholesterol content (Roche et al., 2008; Orth et al., 2011).This preferential association of some lipids in complex mixtures has resulted in the “membrane raft” hypothesis within the cell biology field. This theory postulates the existence of small (20–200 nm), short-lived, sterol-, and sphingolipid-enriched Lo assemblies within the membrane. An important feature is that these aggregations are believed to coalesce, upon a biological stimulus, into larger structures whose dynamics can regulate many cellular processes (Simons and Ikonen, 1997; Pike, 2006; Lingwood and Simons, 2010; Simons and Gerl, 2010). An increased resistance to solubilization by detergents of Lo versus Ld phases has led researchers to consider that membrane fractions insoluble to nonionic detergents at low temperatures could contain the putative “raft” fractions. One caveat of this theory is that recovered detergent-insoluble membrane fractions (DIMs) only exist after detergent treatment and do not correspond to the native membrane structure (Lichtenberg et al., 2005). Nevertheless, their significant enrichment in sterols, sphingolipids, and specific subsets of proteins, some of which displaying a clustered distribution within the PM (Simons and Gerl, 2010), has encouraged their use as a biochemical counterpart of Lo microdomains existing in biological membranes.Plant DIMs with a lipid content similar to animal DIMs have been isolated from several species, including tobacco cells, and are enriched in proteins involved in signaling and stress responses (Mongrand et al., 2004; Borner et al., 2005; Morel et al., 2006; Lefebvre et al., 2007; Kierszniowska et al., 2009). Moreover, immunoelectron microscopy experiments have revealed that lateral segregation of lipids and proteins occurs at the nanoscale level at the tobacco PM, thus correlating detergent insolubility with membrane domain localization of presumptive raft proteins (Raffaele et al., 2009; Furt et al., 2010; Demir et al., 2013). Together, these data point to the existence of specialized lipid domains in plants. Concomitantly, the presence of sterol-rich Lo membrane domains was observed in vivo at the tip of the growing pollen tube in Picea meyeri, using both filipin and the fluorescent probe 1-[2-hydroxy-3-(N,N-dimethyl-N-hydroxyethyl)ammoniopropyl]-4-[β-[2-(di-n-butylamino)-6-napthyl]vinyl] pyridinium dibromide (di-4-ANEPPDHQ; Liu et al., 2009). This observation argues in favor of a sterol-dependent organization of ordered domains at the plant PM surface. In addition, the combined use of fluorescent lipid analogs and the environmental dye laurdan has revealed different lipid phases that emerge in the PM of Arabidopsis (Arabidopsis thaliana) protoplasts during restoration of the cell wall (Blachutzik et al., 2012). Despite these details, necessary data concerning the presence and in vivo characterization of Lo domains at a micrometer to nanometer scale are still lacking.The importance of a more refined resolution for observing Lo domains was proposed in several recent reviews (Bagatolli, 2006; Duggan et al., 2008; García-Sáez and Schwille, 2010; Owen et al., 2010a; Stöckl and Herrmann, 2010; Klenerman et al., 2011). Although the physical properties of biological membranes have been studied in situ by various techniques, including two-channel ratiometric microscopy (Owen et al., 2010c) and microscopy imaging of partitioning of fluorescent lipids and proteins (Rosetti et al., 2010) or environmentally sensitive probes (Parasassi et al., 1990; Jin et al., 2006), membrane segregation into microscopic Lo- and Ld-like phases has been difficult to observe in living cells. Furthermore, only a few studies have demonstrated that a microscopic phase separation involving an ordered phase similar to the Lo domain of model membranes could occur in biomembranes using PM giant vesicles (Baumgart et al., 2007; Lingwood et al., 2008; Sengupta et al., 2008). A potentially powerful approach for imaging small ordered membrane domains relies on environment-sensitive probes coupled with fluorescence spectroscopy (Gaus et al., 2003, 2006; Oncul et al., 2010). In particular, analysis of the fluorescence of the di-4-ANEPPDHQ probe, which exhibits an emission shift independent of local chemical composition under different lipid packing conditions (Jin et al., 2005; Demchenko et al., 2009; Dinic et al., 2011), recently enabled the imaging of plant membrane domains at the micrometer scale (Liu et al., 2009). The relevance of this approach has been confirmed by mapping membrane domains using generalized anisotropy-based images of di-4-ANEPPDHQ-stained T cell immunological synapses (Owen et al., 2010c), together with the characterization of membrane organization of nonadherent cells (such as living zebrafish embryo tissues) labeled with this dye (Owen et al., 2012a).The function of dynamic PM compartmentalization in the detection and transduction of environmental signals in plant cells has only recently begun to emerge, along with a crucial role for sterols in this organization (for review, see Zappel and Panstruga, 2008; Mongrand et al., 2010; Simon-Plas et al., 2011). These observations make it indispensable to align how the surface membrane of living cells might reorganize during signaling with the membrane raft hypothesis. To investigate possible modifications of membrane organization during the initial steps of plant defense signaling, tobacco cells were treated with two well-described elicitors of defense reaction, cryptogein, a small protein able to trigger an hypersensitive reaction (HR) and an acquired resistance in tobacco plants (Ponchet et al., 1999; Garcia Brugger et al., 2006) together with a widely described signaling cascade in tobacco suspension cells, and flg22 (a 22-amino acid peptide corresponding to a conserved domain of bacterial flagellin). The latter peptide is also a potent elicitor in plants, yet it does not induce an HR type of necrosis (Gomez-Gomez and Boller, 2002; Chinchilla et al., 2007). The study of cryptogein response reveals that the earliest steps of the signal transduction pathway mainly involve PM activities (Ponchet et al., 1999; Garcia-Brugger et al., 2006). How the PM is laterally organized and possibly reorganized in response to this stress so it can efficiently trigger a signaling cascade remains unknown.Here, we have developed a confocal multispectral microscopy approach to generate in vivo ratiometric pictures of large areas of the tobacco cell PM labeled with di-4-ANEPPDHQ, allowing the in vivo characterization of the global level of order of this membrane. Although an increase in the proportion of ordered phase within the membrane transiently occurred in the early steps of the cryptogein and flg22 signaling cascades, the fluorescence recovery after photobleaching (FRAP) technique revealed an increase in PM fluidity induced by cryptogein, but not by flagellin. Moreover, we characterized the spatial distribution of Lo phases on the membrane of living plant cells and monitored the variations induced by cryptogein elicitation. The results are discussed within the framework of the “membrane raft” hypothesis, in which we propose a new mechanism of signaling platform formation in the context of plant defense.  相似文献   

3.
4.
5.
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   

6.
The endoplasmic reticulum (ER) is a ubiquitous organelle that plays roles in secretory protein production, folding, quality control, and lipid biosynthesis. The cortical ER in plants is pleomorphic and structured as a tubular network capable of morphing into flat cisternae, mainly at three-way junctions, and back to tubules. Plant reticulon family proteins (RTNLB) tubulate the ER by dimerization and oligomerization, creating localized ER membrane tensions that result in membrane curvature. Some RTNLB ER-shaping proteins are present in the plasmodesmata (PD) proteome and may contribute to the formation of the desmotubule, the axial ER-derived structure that traverses primary PD. Here, we investigate the binding partners of two PD-resident reticulon proteins, RTNLB3 and RTNLB6, that are located in primary PD at cytokinesis in tobacco (Nicotiana tabacum). Coimmunoprecipitation of green fluorescent protein-tagged RTNLB3 and RTNLB6 followed by mass spectrometry detected a high percentage of known PD-localized proteins as well as plasma membrane proteins with putative membrane-anchoring roles. Förster resonance energy transfer by fluorescence lifetime imaging microscopy assays revealed a highly significant interaction of the detected PD proteins with the bait RTNLB proteins. Our data suggest that RTNLB proteins, in addition to a role in ER modeling, may play important roles in linking the cortical ER to the plasma membrane.The endoplasmic reticulum (ER) is a multifunctional organelle (Hawes et al., 2015) and is the site of secretory protein production, folding, and quality control (Brandizzi et al., 2003) and lipid biosynthesis (Wallis and Browse, 2010), but it is also involved in many other aspects of day-to-day plant life, including auxin regulation (Friml and Jones, 2010) and oil and protein body formation (Huang, 1996; Herman, 2008). The cortical ER network displays a remarkable polygonal arrangement of motile tubules that are capable of morphing into small cisternae, mainly at the three-way junctions of the ER network (Sparkes et al., 2009). The cortical ER network of plants has been shown to play multiple roles in protein trafficking (Palade, 1975; Vitale and Denecke, 1999) and pathogen responses (for review, see Pattison and Amtmann, 2009; Beck et al., 2012).In plants, the protein family of reticulons (RTNLBs) contributes significantly to tubulation of the ER (Tolley et al., 2008, 2010; Chen et al., 2012). RTNLBs are integral ER membrane proteins that feature a C-terminal reticulon homology domain (RHD) that contains two major hydrophobic regions. These regions form two V-shaped transmembrane wedges joined together via a cytosolic loop, with the C and N termini of the protein facing the cytosol. RTNLBs can dimerize or oligomerize, creating localized tensions in the ER membrane, inducing varying degrees of membrane curvature (Sparkes et al., 2010). Hence, RTNLBs are considered to be essential in maintaining the tubular ER network.The ability of RTNLBs to constrict membranes is of interest in the context of cell plate development and the formation of primary plasmodesmata (PD; Knox et al., 2015). PD formation involves extensive remodeling of the cortical ER into tightly furled tubules to form the desmotubules, axial structures that run through the PD pore (Overall and Blackman, 1996; Ehlers and Kollmann, 2001). At only 15 nm in diameter, the desmotubule is one of the most constricted membrane structures found in nature, with no animal counterparts (Tilsner et al., 2011). PD are membrane-rich structures characterized by a close association of the plasma membrane (PM) with the ER. The forces that model the ER into desmotubules, however, are poorly understood. RTNLBs are excellent candidates for this process and can constrict fluorescent protein-labeled ER membranes into extremely fine tubules (Sparkes et al., 2010). We showed recently that two of the RTNLBs present in the PD proteome, RTNLB3 and RTNLB6 (Fernandez-Calvino et al., 2011), are present in primary PD at cytokinesis (Knox et al., 2015). However, nothing is known of the proteins that interact with RTNLBs identified in the PD proteome or that may link RTNLBs to the PM. To date, the only protein shown to bind to plant RTNLBs is RHD3-LIKE2, the plant homolog of the ER tubule fusion protein ATLASTIN (Lee et al., 2013).Here, we used a dual approach to identify interacting partners of RTNLB3 and RTNLB6 (Fernandez-Calvino et al., 2011; Knox et al.., 2015). First, we used GFP immunoprecipitation assays coupled to mass spectrometry (MS) to identify proteins potentially binding to RTNLB3 and RTNLB6. Second, from the proteins we identified, we conducted a detailed Förster resonance energy transfer by fluorescence lifetime imaging microscopy (FRET-FLIM) analysis to confirm prey-bait interactions in vivo.The application of time-resolved fluorescence spectroscopy to imaging biological systems has allowed the design and implementation of fluorescence lifetime imaging microscopy (FLIM). The technique allows measuring and determining the space map of picosecond fluorescence decay at each pixel of the image through confocal single and multiphoton excitation. The general fluorescence or Förster resonance energy transfer (FRET) to determine the colocalization of two color chromophores can now be improved to determine physical interactions using FRET-FLIM and protein pairs tagged with appropriate GFP fluorophores and monomeric red fluorescent protein (mRFP). FRET-FLIM measures the reduction in the excited-state lifetime of GFP (donor) fluorescence in the presence of an acceptor fluorophore (e.g. mRFP) that is independent of the problems associated with steady-state intensity measurements. The observation of such a reduction is an indication that the two proteins are within a distance of 1 to 10 nm, thus indicating a direct physical interaction between the two protein fusions (Osterrieder et al., 2009; Sparkes et al., 2010; Schoberer and Botchway, 2014). It was shown previously that a reduction of as little as approximately 200 ps in the excited-state lifetime of the GFP-labeled protein represents quenching through a protein-protein interaction (Stubbs et al., 2005).Our interaction data identified a large percentage (40%) of ER proteins, including other RTNLB family members. However, we also found a relatively large number (25%) of proteins present in the published PD proteome (Fernandez-Calvino et al., 2011) and a surprisingly high proportion (35%) of PM proteins. Of the PD-resident proteins we identified, a significant number were shown previously to be targets of viral movement proteins (MPs) or proteins present within lipid rafts, consistent with the view that PD are lipid-rich microdomains (Bayer et al., 2014). Additional proteins identified suggested roles for RTNLBs in transport and pathogen defense. We suggest that RTNLBs may play key roles in anchoring and/or signaling between the cortical ER and PM.  相似文献   

7.
8.
9.
10.
11.
12.
In plants, K transporter (KT)/high affinity K transporter (HAK)/K uptake permease (KUP) is the largest potassium (K) transporter family; however, few of the members have had their physiological functions characterized in planta. Here, we studied OsHAK5 of the KT/HAK/KUP family in rice (Oryza sativa). We determined its cellular and tissue localization and analyzed its functions in rice using both OsHAK5 knockout mutants and overexpression lines in three genetic backgrounds. A β-glucuronidase reporter driven by the OsHAK5 native promoter indicated OsHAK5 expression in various tissue organs from root to seed, abundantly in root epidermis and stele, the vascular tissues, and mesophyll cells. Net K influx rate in roots and K transport from roots to aerial parts were severely impaired by OsHAK5 knockout but increased by OsHAK5 overexpression in 0.1 and 0.3 mm K external solution. The contribution of OsHAK5 to K mobilization within the rice plant was confirmed further by the change of K concentration in the xylem sap and K distribution in the transgenic lines when K was removed completely from the external solution. Overexpression of OsHAK5 increased the K-sodium concentration ratio in the shoots and salt stress tolerance (shoot growth), while knockout of OsHAK5 decreased the K-sodium concentration ratio in the shoots, resulting in sensitivity to salt stress. Taken together, these results demonstrate that OsHAK5 plays a major role in K acquisition by roots faced with low external K and in K upward transport from roots to shoots in K-deficient rice plants.Potassium (K) is one of the three most important macronutrients and the most abundant cation in plants. As a major osmoticum in the vacuole, K drives the generation of turgor pressure, enabling cell expansion. In the vascular tissue, K is an important participant in the generation of root pressure (for review, see Wegner, 2014 [including his new hypothesis]). In the phloem, K is critical for the transport of photoassimilates from source to sink (Marschner, 1996; Deeken et al., 2002; Gajdanowicz et al., 2011). In addition, enhancing K absorption and decreasing sodium (Na) accumulation is a major strategy of glycophytes in salt stress tolerance (Maathuis and Amtmann, 1999; Munns and Tester, 2008; Shabala and Cuin, 2008).Plants acquire K through K-permeable proteins at the root surface. Since available K concentration in the soil may vary by 100-fold, plants have developed multiple K uptake systems for adapting to this variability (Epstein et al., 1963; Grabov, 2007; Maathuis, 2009). In a classic K uptake experiment in barley (Hordeum vulgare), root K absorption has been described as a high-affinity and low-affinity biphasic transport process (Epstein et al., 1963). It is generally assumed that the low-affinity transport system (LATS) in the roots mediates K uptake in the millimolar range and that the activity of this system is insensitive to external K concentration (Maathuis and Sanders, 1997; Chérel et al., 2014). In contrast, the high-affinity transport system (HATS) was rapidly up-regulated when the supply of exogenous K was halted (Glass, 1976; Glass and Dunlop, 1978).The membrane transporters for K flux identified in plants are generally classified into three channels and three transporter families based on phylogenetic analysis (Mäser et al., 2001; Véry and Sentenac, 2003; Lebaudy et al., 2007; Alemán et al., 2011). For K uptake, it was predicted that, under most circumstances, K transporters function as HATS, while K-permeable channels mediate LATS (Maathuis and Sanders, 1997). However, a root-expressed K channel in Arabidopsis (Arabidopsis thaliana), Arabidopsis K Transporter1 (AKT1), mediates K absorption over a wide range of external K concentrations (Sentenac et al., 1992; Lagarde et al., 1996; Hirsch et al., 1998; Spalding et al., 1999), while evidence is accumulating that many K transporters, including members of the K transporter (KT)/high affinity K transporter (HAK)/K uptake permease (KUP) family, are low-affinity K transporters (Quintero and Blatt, 1997; Senn et al., 2001), implying that functions of plant K channels and transporters overlap at different K concentration ranges.Out of the three families of K transporters, cation proton antiporter (CPA), high affinity K/Na transporter (HKT), and KT/HAK/KUP, CPA was characterized as a K+(Na+)/H+ antiporter, HKT may cotransport Na and K or transport Na only (Rubio et al., 1995; Uozumi et al., 2000), while KT/HAK/KUP were predicted to be H+-coupled K+ symporters (Mäser et al., 2001; Lebaudy et al., 2007). KT/HAK/KUP were named by different researchers who first identified and cloned them (Quintero and Blatt, 1997; Santa-María et al., 1997). In plants, the KT/HAK/KUP family is the largest K transporter family, including 13 members in Arabidopsis and 27 members in the rice (Oryza sativa) genome (Rubio et al., 2000; Mäser et al., 2001; Bañuelos et al., 2002; Gupta et al., 2008). Sequence alignments show that genes of this family share relatively low homology to each other. The KT/HAK/KUP family was divided into four major clusters (Rubio et al., 2000; Gupta et al., 2008), and in cluster I and II, they were further separated into A and B groups. Genes of cluster I or II likely exist in all plants, cluster III is composed of genes from both Arabidopsis and rice, while cluster IV includes only four rice genes (Grabov, 2007; Gupta et al., 2008).The functions of KT/HAK/KUP were studied mostly in heterologous expression systems. Transporters of cluster I, such as AtHAK5, HvHAK1, OsHAK1, and OsHAK5, are localized in the plasma membrane (Kim et al., 1998; Bañuelos et al., 2002; Gierth et al., 2005) and exhibit high-affinity K uptake in the yeast Saccharomyces cerevisiae (Santa-María et al., 1997; Fu and Luan, 1998; Rubio et al., 2000) and in Escherichia coli (Horie et al., 2011). Transporters of cluster II, like AtKUP4 (TINY ROOT HAIRS1, TRH1), HvHAK2, OsHAK2, OsHAK7, and OsHAK10, could not complement the K uptake-deficient yeast (Saccharomyces cerevisiae) but were able to mediate K fluxes in a bacterial mutant; they might be tonoplast transporters (Senn et al., 2001; Bañuelos et al., 2002; Rodríguez-Navarro and Rubio, 2006). The function of transporters in clusters III and IV is even less known (Grabov, 2007).Existing data suggest that some KT/HAK/KUP transporters also may respond to salinity stress (Maathuis, 2009). The cluster I transporters of HvHAK1 mediate Na influx (Santa-María et al., 1997), while AtHAK5 expression is inhibited by Na (Rubio et al., 2000; Nieves-Cordones et al., 2010). Expression of OsHAK5 in tobacco (Nicotiana tabacum) BY2 cells enhanced the salt tolerance of these cells by accumulating more K without affecting their Na content (Horie et al., 2011).There are only scarce reports on the physiological function of KT/HAK/KUP in planta. In Arabidopsis, mutation of AtKUP2 (SHORT HYPOCOTYL3) resulted in a short hypocotyl, small leaves, and a short flowering stem (Elumalai et al., 2002), while a loss-of-function mutation of AtKUP4 (TRH1) resulted in short root hairs and a loss of gravity response in the root (Rigas et al., 2001; Desbrosses et al., 2003; Ahn et al., 2004). AtHAK5 is the only system currently known to mediate K uptake at concentrations below 0.01 mm (Rubio et al., 2010) and provides a cesium uptake pathway (Qi et al., 2008). AtHAK5 and AtAKT1 are the two major physiologically relevant molecular entities mediating K uptake into roots in the range between 0.01 and 0.05 mm (Pyo et al., 2010; Rubio et al., 2010). AtAKT1 may contribute to K uptake within the K concentrations that belong to the high-affinity system described by Epstein et al. (1963).Among all 27 members of the KT/HAK/KUP family in rice, OsHAK1, OsHAK5, OsHAK19, and OsHAK20 were grouped in cluster IB (Gupta et al., 2008). These four rice HAK members share 50.9% to 53.4% amino acid identity with AtHAK5. OsHAK1 was expressed in the whole plant, with maximum expression in roots, and was up-regulated by K deficiency; it mediated high-affinity K uptake in yeast (Bañuelos et al., 2002). In this study, we examined the tissue-specific localization and the physiological functions of OsHAK5 in response to variation in K supply and to salt stress in rice. By comparing K uptake and translocation in OsHAK5 knockout (KO) mutants and in OsHAK5-overexpressing lines with those in their respective wild-type lines supplied with different K concentrations, we found that OsHAK5 not only mediates high-affinity K acquisition but also participates in root-to-shoot K transport as well as in K-regulated salt tolerance.  相似文献   

13.
14.
15.
We have established an efficient transient expression system with several vacuolar reporters to study the roles of endosomal sorting complex required for transport (ESCRT)-III subunits in regulating the formation of intraluminal vesicles of prevacuolar compartments (PVCs)/multivesicular bodies (MVBs) in plant cells. By measuring the distributions of reporters on/within the membrane of PVC/MVB or tonoplast, we have identified dominant negative mutants of ESCRT-III subunits that affect membrane protein degradation from both secretory and endocytic pathways. In addition, induced expression of these mutants resulted in reduction in luminal vesicles of PVC/MVB, along with increased detection of membrane-attaching vesicles inside the PVC/MVB. Transgenic Arabidopsis (Arabidopsis thaliana) plants with induced expression of ESCRT-III dominant negative mutants also displayed severe cotyledon developmental defects with reduced cell size, loss of the central vacuole, and abnormal chloroplast development in mesophyll cells, pointing out an essential role of the ESCRT-III complex in postembryonic development in plants. Finally, membrane dissociation of ESCRT-III components is important for their biological functions and is regulated by direct interaction among Vacuolar Protein Sorting-Associated Protein20-1 (VPS20.1), Sucrose Nonfermenting7-1, VPS2.1, and the adenosine triphosphatase VPS4/SUPPRESSOR OF K+ TRANSPORT GROWTH DEFECT1.Endomembrane trafficking in plant cells is complicated such that secretory, endocytic, and recycling pathways are usually integrated with each other at the post-Golgi compartments, among which, the trans-Golgi network (TGN) and prevacuolar compartment (PVC)/multivesicular body (MVB) are best studied (Tse et al., 2004; Lam et al., 2007a, 2007b; Müller et al., 2007; Foresti and Denecke, 2008; Hwang, 2008; Otegui and Spitzer, 2008; Robinson et al., 2008; Richter et al., 2009; Ding et al., 2012; Gao et al., 2014). Following the endocytic trafficking of a lipophilic dye, FM4-64, the TGN and PVC/MVB are sequentially labeled and thus are defined as the early and late endosome, respectively, in plant cells (Lam et al., 2007a; Chow et al., 2008). While the TGN is a tubular vesicular-like structure that may include several different microdomains and fit its biological function as a sorting station (Chow et al., 2008; Kang et al., 2011), the PVC/MVB is 200 to 500 nm in size with multiple luminal vesicles of approximately 40 nm (Tse et al., 2004). Membrane cargoes destined for degradation are sequestered into these tiny luminal vesicles and delivered to the lumen of the lytic vacuole (LV) via direct fusion between the PVC/MVB and the LV (Spitzer et al., 2009; Viotti et al., 2010; Cai et al., 2012). Therefore, the PVC/MVB functions between the TGN and LV as an intermediate organelle and decides the fate of membrane cargoes in the LV.In yeast (Saccharomyces cerevisiae), carboxypeptidase S (CPS) is synthesized as a type II integral membrane protein and sorted from the Golgi to the lumen of the vacuole (Spormann et al., 1992). Genetic analyses on the trafficking of CPS have led to the identification of approximately 17 class E genes (Piper et al., 1995; Babst et al., 1997, 2002a, 2002b; Odorizzi et al., 1998; Katzmann et al., 2001) that constitute the core endosomal sorting complex required for transport (ESCRT) machinery. The evolutionarily conserved ESCRT complex consists of several functionally different subcomplexes, ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III and the ESCRT-III-associated/Vacuolar Protein Sorting4 (VPS4) complex. Together, they form a complex protein-protein interaction network that coordinates sorting of cargoes and inward budding of the membrane on the MVB (Hurley and Hanson, 2010; Henne et al., 2011). Cargo proteins carrying ubiquitin signals are thought to be passed from one ESCRT subcomplex to the next, starting with their recognition by ESCRT-0 (Bilodeau et al., 2002, 2003; Hislop and von Zastrow, 2011; Le Bras et al., 2011; Shields and Piper, 2011; Urbé, 2011). ESCRT-0 recruits the ESCRT-I complex, a heterotetramer of VPS23, VPS28, VPS37, and MVB12, from the cytosol to the endosomal membrane (Katzmann et al., 2001, 2003). The C terminus of VPS28 interacts with the N terminus of VPS36, a member of the ESCRT-II complex (Kostelansky et al., 2006; Teo et al., 2006). Then, cargoes passed from ESCRT-I and ESCRT-II are concentrated in certain membrane domains of the endosome by ESCRT-III, which includes four coiled-coil proteins and is sufficient to induce the membrane invagination (Babst et al., 2002b; Saksena et al., 2009; Wollert et al., 2009). Finally, the ESCRT components are disassociated from the membrane by the adenosine triphosphatase (ATPase) associated with diverse cellular activities (AAA) VPS4/SUPPRESSOR OF K+ TRANSPORT GROWTH DEFECT1 (SKD1) before releasing the internal vesicles (Babst et al., 1997, 1998).Putative homologs of ESCRT-I–ESCRT-III and ESCRT-III-associated components have been identified in plants, except for ESCRT-0, which is only present in Opisthokonta (Winter and Hauser, 2006; Leung et al., 2008; Schellmann and Pimpl, 2009). To date, only a few plant ESCRT components have been studied in detail. The Arabidopsis (Arabidopsis thaliana) AAA ATPase SKD1 localized to the PVC/MVB and showed ATPase activity that was regulated by Lysosomal Trafficking Regulator-Interacting Protein5, a plant homolog of Vps Twenty Associated1 Protein (Haas et al., 2007). Expression of the dominant negative form of SKD1 caused an increase in the size of the MVB and a reduction in the number of internal vesicles (Haas et al., 2007). This protein also contributes to the maintenance of the central vacuole and might be associated with cell cycle regulation, as leaf trichomes expressing its dominant negative mutant form lost the central vacuole and frequently contained multiple nuclei (Shahriari et al., 2010). Double null mutants of CHARGED MULTIVESICULAR BODY PROTEIN, chmp1achmp1b, displayed severe growth defects and were seedling lethal. This may be due to the mislocalization of plasma membrane (PM) proteins, including those involved in auxin transport such as PINFORMED1, PINFORMED2, and AUXIN-RESISTANT1, from the vacuolar degradation pathway to the tonoplast of the LV (Spitzer et al., 2009).Plant ESCRT components usually contain several homologs, with the possibility of functional redundancy. Single mutants of individual ESCRT components may not result in an obvious phenotype, whereas knockout of all homologs of an ESCRT component by generating double or triple mutants may be lethal to the plant. As a first step to carry out systematic analysis on each ESCRT complex in plant cells, here, we established an efficient analysis system to monitor the localization changes of four vacuolar reporters that accumulate either in the lumen (LRR84A-GFP, EMP12-GFP, and aleurain-GFP) or on the tonoplast (GFP-VIT1) of the LV and identified several ESCRT-III dominant negative mutants. We reported that ESCRT-III subunits were involved in the release of PVC/MVB’s internal vesicles from the limiting membrane and were required for membrane protein degradation from secretory and endocytic pathways. In addition, transgenic Arabidopsis plants with induced expression of ESCRT-III dominant negative mutants showed severe cotyledon developmental defects. We also showed that membrane dissociation of ESCRT-III subunits was regulated by direct interaction with SKD1.  相似文献   

16.
17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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