首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
Proper utilization of plant disease resistance genes requires a good understanding of their short- and long-term evolution. Here we present a comprehensive study of the long-term evolutionary history of nucleotide-binding site (NBS)-leucine-rich repeat (LRR) genes within and beyond the legume family. The small group of NBS-LRR genes with an amino-terminal RESISTANCE TO POWDERY MILDEW8 (RPW8)-like domain (referred to as RNL) was first revealed as a basal clade sister to both coiled-coil-NBS-LRR (CNL) and Toll/Interleukin1 receptor-NBS-LRR (TNL) clades. Using Arabidopsis (Arabidopsis thaliana) as an outgroup, this study explicitly recovered 31 ancestral NBS lineages (two RNL, 21 CNL, and eight TNL) that had existed in the rosid common ancestor and 119 ancestral lineages (nine RNL, 55 CNL, and 55 TNL) that had diverged in the legume common ancestor. It was shown that, during their evolution in the past 54 million years, approximately 94% (112 of 119) of the ancestral legume NBS lineages experienced deletions or significant expansions, while seven original lineages were maintained in a conservative manner. The NBS gene duplication pattern was further examined. The local tandem duplications dominated NBS gene gains in the total number of genes (more than 75%), which was not surprising. However, it was interesting from our study that ectopic duplications had created many novel NBS gene loci in individual legume genomes, which occurred at a significant frequency of 8% to 20% in different legume lineages. Finally, by surveying the legume microRNAs that can potentially regulate NBS genes, we found that the microRNA-NBS gene interaction also exhibited a gain-and-loss pattern during the legume evolution.To combat the constant challenges by pathogens, plants have evolved a sophisticated two-layered defense system, in which proteins encoded by disease RESISTANCE (R) genes serve to sense pathogen invasion signals and to elicit defense responses (Jones and Dangl, 2006; McDowell and Simon, 2006; Bent and Mackey, 2007). Over 140 R genes have been characterized from different flowering plants, which confer resistance to a large array of pathogens, including bacteria, fungi, oomycetes, viruses, and nematodes (Liu et al., 2007; Yang et al., 2013). Among these, about 80% belong to the NBS-LRR class, which encodes a central nucleotide-binding site (NBS) domain and a C-terminal leucine-rich repeat (LRR) domain. Based on whether their N termini are homologous to the Toll/Interleukin1 receptor (TIR), the angiosperm NBS-LRR genes are further divided into the TIR-NBS-LRR (TNL) subclass and the non-TIR-NBS-LRR (nTNL) subclass (Meyers et al., 1999; Bai et al., 2002; Cannon et al., 2002). The latter has been also called CC-NBS-LRR (CNL) subclass, since a coiled-coil (CC) structure is often detected at the N terminus (Meyers et al., 2003). Interestingly, a small group of nTNL genes have an N-terminal RPW8-like domain with a transmembrane region before the CC structure (Xiao et al., 2001). This group of RPW8-NBS-LRR (RNL) genes has been usually viewed as a specific lineage of CNLs (Bonardi et al., 2011; Collier et al., 2011); however, its real phylogenetic relationship with CNLs and TNLs requires further investigation.NBS-LRR genes have been surveyed in many sequenced genomes of flowering plants, including four monocots: rice (Oryza sativa), maize (Zea mays), sorghum (Sorghum bicolor), and Brachypodium distachyon; one basal eudicot: Nelumbo nucifera; two asterid species: potato (Solanum tuberosum) and tomato (Solanum lycopersicum); and 14 rosids: Vitis vinifera, Populus trichocarpa, Ricinus communis, Medicago truncatula, soybean (Glycine max), Lotus japonicus, Cucumis sativus, Cucumis melo, Citrullus lanatus, Gossypium raimondii, Carica papaya, Arabidopsis (Arabidopsis thaliana), Arabidopsis lyrata, and Brassica rapa (Bai et al., 2002; Meyers et al., 2003; Monosi et al., 2004; Zhou et al., 2004; Yang et al., 2006, 2008b; Ameline-Torregrosa et al., 2008; Mun et al., 2009; Porter et al., 2009; Chen et al., 2010; Li et al., 2010a, 2010b; Guo et al., 2011; Zhang et al., 2011; Jupe et al., 2012; Lozano et al., 2012; Luo et al., 2012; Tan and Wu, 2012; Andolfo et al., 2013; Jia et al., 2013; Lin et al., 2013; Wan et al., 2013; Wei et al., 2013; Wu et al., 2014). Variable numbers (from dozens to hundreds) of NBS-LRR genes were reported from these genomes, making one wonder: how did these genes evolve so variably during flowering plant speciation?Comparative genomic studies conducted in the available genome sequences of closely related species or subspecies revealed that a significant proportion of NBS-LRR genes are not shared. For example, 70 NBS-LRR genes between Arabidopsis and A. lyrata show the presence/absence of polymorphisms (Chen et al., 2010; Guo et al., 2011). Moreover, synteny analysis revealed that, among 363 NBS-LRR gene loci in indica (cv 93-11) and japonica (cv Nipponbare) rice, 124 loci exist in only one genome (Luo et al., 2012). Unequal crossover, homologous repair, and nonhomologous repair are the three ways that NBS-LRR gene deletions are caused in rice genomes (Luo et al., 2012).In many surveyed genomes, the majority of NBS-LRR genes are found in a clustered organization (physically close to each other), with the rest exhibited as singletons. Many clusters are homogenous, with only duplicated members occupying the same phylogenetic lineage, whereas heterogenous clusters comprise members from distantly related clades (Meyers et al., 2003). Leister (2004) defined three types of NBS gene duplications: local tandem, ectopic, and segmental duplications. Although a general agreement on the widespread occurrence of local tandem duplications can be reached by various genome survey studies, the relative importance of ectopic and segmental duplications has been seldom investigated since the earliest surveys of the Arabidopsis genome (Richly et al., 2002; Baumgarten et al., 2003; Meyers et al., 2003; McDowell and Simon, 2006).With more genomic data available in certain angiosperm families, NBS-LRR genes should be further investigated among phylogenetically distant species to fill the gaps in the understanding of their long-term evolutionary patterns. The legume family contains many economically important crop species, such as clover (Trifolium spp.), soybean, peanut (Arachis hypogaea), and common bean (Phaseolus vulgaris). Although these legumes are constantly threatened by various pathogens, only a few functional legume R genes have been characterized, and all of them belong to the NBS-LRR class (Ashfield et al., 2004; Hayes et al., 2004; Gao et al., 2005; Seo et al., 2006; Yang et al., 2008a; Meyer et al., 2009). Therefore, it would be interesting to investigate the NBS-LRR gene repertoire among different legume species. Here, we carried out a comprehensive analysis of NBS-LRR genes in four divergent legume genomes, M. truncatula, pigeon pea (Cajanus cajan), common bean, and soybean, which shared a common ancestor approximately 54 million years ago (MYA; Fig. 1; Lavin et al., 2005). Approximately 1,000 nTNL and 667 TNL subclass NBS-encoding genes were identified in our study. Their genomic distribution, organization modes, phylogenetic relationships, and syntenic patterns were examined to obtain insight into the long-term evolutionary patterns of NBS-LRR genes.Open in a separate windowFigure 1.The phylogenetic tree of four investigated legume species (M. truncatula, pigeon pea, common bean, and soybean). Two WGD events are indicated with triangles: one occurred approximately 59 MYA in the common ancestor of the four investigated legumes, and the other occurred approximately 13 MYA in the Glycine spp. lineage alone (Schmutz et al., 2010). The numbers at the branch nodes indicate divergence times (Lavin et al., 2005; Stefanovic et al., 2009). [See online article for color version of this figure.]  相似文献   

3.
The P6 protein of Cauliflower mosaic virus (CaMV) is responsible for the formation of inclusion bodies (IBs), which are the sites for viral gene expression, replication, and virion assembly. Moreover, recent evidence indicates that ectopically expressed P6 inclusion-like bodies (I-LBs) move in association with actin microfilaments. Because CaMV virions accumulate preferentially in P6 IBs, we hypothesized that P6 IBs have a role in delivering CaMV virions to the plasmodesmata. We have determined that the P6 protein interacts with a C2 calcium-dependent membrane-targeting protein (designated Arabidopsis [Arabidopsis thaliana] Soybean Response to Cold [AtSRC2.2]) in a yeast (Saccharomyces cerevisiae) two-hybrid screen and have confirmed this interaction through coimmunoprecipitation and colocalization assays in the CaMV host Nicotiana benthamiana. An AtSRC2.2 protein fused to red fluorescent protein (RFP) was localized to the plasma membrane and specifically associated with plasmodesmata. The AtSRC2.2-RFP fusion also colocalized with two proteins previously shown to associate with plasmodesmata: the host protein Plasmodesmata-Localized Protein1 (PDLP1) and the CaMV movement protein (MP). Because P6 I-LBs colocalized with AtSRC2.2 and the P6 protein had previously been shown to interact with CaMV MP, we investigated whether P6 I-LBs might also be associated with plasmodesmata. We examined the colocalization of P6-RFP I-LBs with PDLP1-green fluorescent protein (GFP) and aniline blue (a stain for callose normally observed at plasmodesmata) and found that P6-RFP I-LBs were associated with each of these markers. Furthermore, P6-RFP coimmunoprecipitated with PDLP1-GFP. Our evidence that a portion of P6-GFP I-LBs associate with AtSRC2.2 and PDLP1 at plasmodesmata supports a model in which P6 IBs function to transfer CaMV virions directly to MP at the plasmodesmata.Through the years, numerous studies have focused on the characterization of viral replication sites within the cell, as well as how plant virus movement proteins (MPs) modify the plasmodesmata to facilitate cell-to-cell movement (for review, see Benitez-Alfonso et al., 2010; Laliberté and Sanfaçon, 2010; Niehl and Heinlein, 2011; Ueki and Citovsky, 2011; Verchot, 2012). It is accepted that plant virus replication is associated with host membranes, and at some point, the viral genomic nucleic acid must be transferred from the site of replication in the cell to the plasmodesmata. This step could involve transport from a distant site within the cell, or alternatively, it may be that replication is coupled with transport at the entrance of the plasmodesmata (Tilsner et al., 2013). However, even with the latter model, there is ample evidence that the viral proteins necessary for replication or cell-to-cell movement utilize intracellular trafficking pathways within the cell to become positioned at the plasmodesma. These pathways may involve microfilaments, microtubules, or specific endomembranes that participate in macromolecular transport pathways, or combinations of these elements (Harries et al., 2010; Schoelz et al., 2011; Patarroyo et al., 2012; Peña and Heinlein, 2012; Tilsner and Oparka 2012; Liu and Nelson, 2013).The P6 protein of Cauliflower mosaic virus (CaMV) is one viral protein that had not been considered to play a role in viral movement until recently. P6 is the most abundant protein component of the amorphous, electron-dense inclusion bodies (IBs) present during virus infection (Odell and Howell, 1980; Shockey et al., 1980). Ectopic expression of P6 in Nicotiana benthamiana leaves resulted in the formation of inclusion-like bodies (I-LBs) that were capable of intracellular movement along actin microfilaments. Furthermore, treatment of Nicotiana edwardsonii leaves with latrunculin B abolished the formation of CaMV local lesions, suggesting that intact microfilaments are required for CaMV infection (Harries et al., 2009a). A subsequent paper showed that P6 physically interacts with Chloroplast Unusual Positioning1 (CHUP1), a plant protein localized to the chloroplast outer membrane that contributes to movement of chloroplasts on microfilaments in response to changes in light intensity (Oikawa et al., 2003, 2008; Angel et al., 2013). The implication was that P6 might hijack CHUP1 to facilitate movement of the P6 IBs on microfilaments. Silencing of CHUP1 in N. edwardsonii, a host for CaMV, slowed the rate of local lesion formation, suggesting that CHUP1 contributes to intracellular movement of CaMV (Angel et al., 2013).In addition to its role in intracellular trafficking, the P6 protein has been shown to have at least four other distinct functions in the viral infection cycle. P6-containing IBs induced during virus infection are likely virion factories, as they are the primary site for CaMV protein synthesis, genome replication, and assembly of virions (Hohn and Fütterer, 1997). Second, P6 interacts with host ribosomes to facilitate reinitiation of translation of genes on the polycistronic 35S viral RNA, a process called translational transactivation (Bonneville et al., 1989; Park et al., 2001; Ryabova et al., 2002). The translational transactivator region of P6 (Fig. 1) defines the essential sequences required for translational transactivation (DeTapia et al., 1993). Third, P6 is an important pathogenicity determinant. P6 functions as an avirulence determinant in some solanaceous and cruciferous species (Daubert et al., 1984; Schoelz et al., 1986; Hapiak et al., 2008) and is a chlorosis symptom determinant in susceptible hosts (Daubert et al., 1984; Baughman et al., 1988; Goldberg et al., 1991; Cecchini et al., 1997). Finally, P6 has the capacity to compromise host defenses, as it is a suppressor of RNA silencing and cell death (Love et al., 2007; Haas et al., 2008), and it modulates signaling by salicylic acid, jasmonic acid, ethylene, and auxin (Geri et al., 2004; Love et al., 2012; Laird et al., 2013). Domain D1 of P6 has been shown to be necessary but not sufficient for suppression of silencing and salicylic acid-mediated defenses (Laird et al., 2013).Open in a separate windowFigure 1.CaMV and host constructs used for confocal microscopy or coimmunoprecipitation (co-IP). A, Structure of CaMV P6 and Arabidopsis (Arabidopsis thaliana) Soybean Response to Cold (AtSRC2.2) proteins. The functions of P6 domains D1 to D4 tested for interaction with AtSRC2.2 are indicated by the shaded boxes. The Mini TAV is the minimal region for the translational transactivation function. The NLSa sequence corresponds to the nuclear localization signal of influenza virus. The NLS sequence corresponds to the nuclear localization signal of human ribosomal protein L22. B, Structure of P6 (Angel et al., 2013), AtSRC2.2, PDLP (Thomas et al., 2008), and CaMV MP fusions developed for confocal microscopy and/or co-IP. aa, Amino acid.Because P6-containing IBs are the site for virion accumulation and they are capable of movement, they may be responsible for delivering virions to the CaMV MP located at the plasmodesmata (for review, see Schoelz et al., 2011). The vast majority of CaMV virions accumulate in association with P6-containing IBs. Furthermore, P6 physically interacts with the CaMV capsid and MP, as well as the two proteins necessary for aphid transmission, P2 and P3 (Himmelbach et al., 1996; Ryabova et al., 2002; Hapiak et al., 2008; Lutz et al., 2012). Recent studies have indicated that P6 IBs serve as a reservoir for virions, in which the virions may be rapidly transferred to P2 electron-lucent IBs for acquisition by aphids (Bak et al., 2013). It stands to reason that P6 IBs may also serve as a reservoir for CaMV virions to be transferred to the CaMV MP in the plasmodesmata.CaMV virions move from cell to cell through plasmodesmata modified into tubules through the function of its MP (Perbal et al., 1993; Kasteel et al., 1996). However, studies have suggested that CaMV virions do not appear to directly interact with the MP. Instead, the MP interacts with the CaMV P3 protein (also known as the virion-associated protein [VAP]), which forms a trimeric structure that is anchored into the virions (Leclerc et al., 1998; Leclerc et al., 2001). Electron microscopy studies have indicated that MP and VAP colocalize with virions only at the entrance to or within the plasmodesmata, and it has been suggested that the VAP/virion complex travels to the plasmodesmata independently from the MP (Stavolone et al., 2005). Consequently, there is a need for a second CaMV protein such as P6 to fulfill the role of delivery of virions to the plasmodesmata (Schoelz et al., 2011).Additional studies have shown that the CaMV MP is incorporated into vesicles and is trafficked on the endomembrane system to reach the plasmodesma (Carluccio et al., 2014). These authors suggest that the CaMV MP is recycled in a vesicular transport pathway between plasmodesmata and early endosome compartments. The CaMV MP interacts with µA-Adaptin (Carluccio et al., 2014) and Movement Protein-Interacting7 (Huang et al., 2001), two proteins shown to have a role in vesicular trafficking. Once the MP arrives at plasmodesmata, it interacts with the Plasmodesmata-Localized Protein (PDLP) proteins, which comprise a family of eight proteins associated with plasmodesmata (Amari et al., 2010). In addition to its interaction with CaMV MP, PDLP1 interacts with the 2B protein of Grapevine fan leaf virus (GFLV) at the base of tubules formed by the 2B protein. Furthermore, an Arabidopsis transfer DNA (T-DNA) mutant line in which three PDLP genes had been knocked out (pdlp1-pdlp2-pdlp3) responded to GFLV and CaMV inoculation with a delayed infection (Amari et al., 2010). This has led to the suggestion that the PDLPs might act as receptors for the MPs of the tubule-forming viruses such as GFLV and CaMV (Amari et al., 2010, 2011).To better understand the function of the P6 protein during CaMV intracellular movement, we have utilized a yeast (Saccharomyces cerevisiae) two-hybrid assay to identify host proteins that interact with CaMV P6. We show that P6 physically interacts with a C2-calcium-dependent protein (designated AtSRC2.2). AtSRC2.2 is a membrane-bound protein that is capable of forming punctate spots associated with plasmodesmata. The localization of AtSRC2.2 with plasmodesmata led to an analysis of interactions between P6 I-LBs, AtSRC2.2, PDLP1, and the CaMV MP and also revealed that a portion of P6 I-LBs are found adjacent to plasmodesmata. These results provide further evidence for a model in which P6 IBs are capable of delivery of virions to plasmodesmata for their transit to other host cells.  相似文献   

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.
7.
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   

8.
Plant resistance to phytopathogenic microorganisms mainly relies on the activation of an innate immune response usually launched after recognition by the plant cells of microbe-associated molecular patterns. The plant hormones, salicylic acid (SA), jasmonic acid, and ethylene have emerged as key players in the signaling networks involved in plant immunity. Rhamnolipids (RLs) are glycolipids produced by bacteria and are involved in surface motility and biofilm development. Here we report that RLs trigger an immune response in Arabidopsis (Arabidopsis thaliana) characterized by signaling molecules accumulation and defense gene activation. This immune response participates to resistance against the hemibiotrophic bacterium Pseudomonas syringae pv tomato, the biotrophic oomycete Hyaloperonospora arabidopsidis, and the necrotrophic fungus Botrytis cinerea. We show that RL-mediated resistance involves different signaling pathways that depend on the type of pathogen. Ethylene is involved in RL-induced resistance to H. arabidopsidis and to P. syringae pv tomato whereas jasmonic acid is essential for the resistance to B. cinerea. SA participates to the restriction of all pathogens. We also show evidence that SA-dependent plant defenses are potentiated by RLs following challenge by B. cinerea or P. syringae pv tomato. These results highlight a central role for SA in RL-mediated resistance. In addition to the activation of plant defense responses, antimicrobial properties of RLs are thought to participate in the protection against the fungus and the oomycete. Our data highlight the intricate mechanisms involved in plant protection triggered by a new type of molecule that can be perceived by plant cells and that can also act directly onto pathogens.In their environment, plants are challenged by potentially pathogenic microorganisms. In response, they express a set of defense mechanisms including preformed structural and chemical barriers, as well as an innate immune response quickly activated after microorganism perception (Boller and Felix, 2009). Plant innate immunity is triggered after recognition by pattern recognition receptors of conserved pathogen- or microbe-associated molecular patterns (PAMPs or MAMPs, respectively) or by plant endogenous molecules released by pathogen invasion and called danger-associated molecular patterns (Boller and Felix, 2009; Dodds and Rathjen, 2010). This first step of recognition leads to the activation of MAMP-triggered immunity (MTI). Successful pathogens can secrete effectors that interfere or suppress MTI, resulting in effector-triggered susceptibility. A second level of perception involves the direct or indirect recognition by specific receptors of pathogen effectors leading to effector-triggered immunity (ETI; Boller and Felix, 2009; Dodds and Rathjen, 2010). Whereas MTI and ETI are thought to involve common signaling network, ETI is usually quantitatively stronger than MTI and associated with more sustained and robust immune responses (Katagiri and Tsuda, 2010; Tsuda and Katagiri, 2010).The plant hormones, salicylic acid (SA), jasmonic acid (JA), and ethylene (ET) have emerged as key players in the signaling networks involved in MTI and ETI (Robert-Seilaniantz et al., 2007; Tsuda et al., 2009; Katagiri and Tsuda, 2010; Mersmann et al., 2010; Tsuda and Katagiri, 2010; Robert-Seilaniantz et al., 2011). Interactions between these signal molecules allow the plant to activate and/or modulate an appropriate spectrum of responses, depending on the pathogen lifestyle, necrotroph or biotroph (Glazebrook, 2005; Koornneef and Pieterse, 2008). It is assumed that JA and ET signaling pathways are important for resistance to necrotrophic fungi including Botrytis cinerea and Alternaria brassicicola (Thomma et al., 2001; Ferrari et al., 2003; Glazebrook, 2005). Infection of Arabidopsis (Arabidopsis thaliana) with B. cinerea causes the induction of the JA/ET responsive gene PLANT DEFENSIN1.2 (PDF1.2; Penninckx et al., 1996; Zimmerli et al., 2001). Induction of PDF1.2 by B. cinerea is blocked in ethylene-insensitive2 (ein2) and coronatine-insensitive1 (coi1) mutants that are respectively defective in ET and JA signal transduction pathways. Moreover, ein2 and coi1 plants are highly susceptible to B. cinerea infection (Thomma et al., 1998; Thomma et al., 1999). JA/ET-dependent responses do not seem to be usually induced during resistance to biotrophs, but they can be effective if they are stimulated prior to pathogen challenge (Glazebrook, 2005). Plants impaired in SA signaling are highly susceptible to biotrophic and hemibiotrophic pathogens. Following pathogen infection, SA hydroxylase (NahG), enhanced disease susceptibility5 (eds5), or SA induction-deficient2 (sid2) plants are unable to accumulate high SA levels and they display heightened susceptibility to Pseudomonas syringae pv tomato (Pst), Hyaloperonospora arabidopsidis, or Erysiphe orontii (Delaney et al., 1994; Lawton et al., 1995; Wildermuth et al., 2001; Nawrath et al., 2002; Vlot et al., 2009). Mutants that are insensitive to SA, such as nonexpressor of PATHOGENESIS-RELATED (PR) genes1 (npr1), have enhanced susceptibility to these pathogens (Cao et al., 1994; Glazebrook et al., 1996; Shah et al., 1997; Dong, 2004). According to some reports, plant defense against necrotrophs also involves SA. Arabidopsis plants expressing the nahG gene and infected with B. cinerea show larger lesions compared with wild-type plants (Govrin and Levine, 2002). In tobacco (Nicotiana tabacum), acidic isoforms of PR3 and PR5 gene that are specifically induced by SA (Ménard et al., 2004) are up-regulated after challenge by B. cinerea (El Oirdi et al., 2010). Resistance to some necrotrophs like Fusarium graminearum involves both SA and JA signaling pathways (Makandar et al., 2010). It is assumed that SA and JA signaling can be antagonistic (Bostock, 2005; Koornneef and Pieterse, 2008; Pieterse et al., 2009; Thaler et al., 2012). In Arabidopsis, SA inhibits JA-dependent resistance against A. brassicicola or B. cinerea (Spoel et al., 2007; Koornneef et al., 2008). Recent studies demonstrated that ET modulates the NPR1-mediated antagonism between SA and JA (Leon-Reyes et al., 2009; Leon-Reyes et al., 2010a) and suppression by SA of JA-responsive gene expression is targeted at a position downstream of the JA biosynthesis pathway (Leon-Reyes et al., 2010b). Synergistic effects of SA- and JA-dependent signaling are also well documented (Schenk et al., 2000; van Wees et al., 2000; Mur et al., 2006) and induction of some defense responses after pathogen challenge requires intact JA, ET, and SA signaling pathways (Campbell et al., 2003).Isolated MAMPs trigger defense responses that also require the activation of SA, JA, and ET signaling pathways (Tsuda et al., 2009; Katagiri and Tsuda, 2010). For instance, treatment with the flagellin peptide flg22 induces many SA-related genes including SID2, EDS5, NPR1, and PR1 (Ferrari et al., 2007; Denoux et al., 2008), causes SA accumulation (Tsuda et al., 2008; Wang et al., 2009), and activates ET signaling (Bethke et al., 2009; Mersmann et al., 2010). Local application of lipopolysaccharides elevates the level of SA (Mishina and Zeier, 2007). The oomycete Pep13 peptide induces defense responses in potato (Solanum tuberosum) that require both SA and JA (Halim et al., 2009). Although signaling networks induced by isolated MAMPs are well documented, the contribution of SA, JA, and ET in MAMP- or PAMP-induced resistance to biotrophs and necrotrophs is poorly understood.Rhamnolipids (RLs) are glycolipids produced by various bacteria species including some Pseudomonas and Burkholderia species. They are essential for bacterial surface motility and biofilm development (Vatsa et al., 2010; Chrzanowski et al., 2012). RLs are potent stimulators of animal immunity (Vatsa et al., 2010). They have recently been shown to elicit plant defense responses and to induce resistance against B. cinerea in grapevine (Vitis vinifera; Varnier et al., 2009). They also participate to biocontrol activity of the plant beneficial bacteria Pseudomonas aeruginosa PNA1 against oomycetes (Perneel et al., 2008). However, the signaling pathways used by RLs to stimulate plant innate immunity are not known. To gain more insights into RL-induced MTI, we investigated RL-triggered defense responses and resistance to the necrotrophic fungus B. cinerea, the biotroph oomycete H. arabidopsidis, and the hemibiotroph bacterium Pst in Arabidopsis. Our results show that RLs trigger an innate immune response in Arabidopsis that protects the plant against these different lifestyle pathogens. We demonstrate that RL-mediated resistance involves separated signaling sectors that depend on the type of pathogen. In plants challenged by RLs, SA has a central role and participates to the restriction of the three pathogens. ET is fully involved in RL-induced resistance to the biotrophic oomycete and to the hemibiotrophic bacterium whereas JA is essential for the resistance to the necrotrophic fungus.  相似文献   

9.
10.
11.
12.
The multifunctional movement protein (MP) of Tomato mosaic tobamovirus (ToMV) is involved in viral cell-to-cell movement, symptom development, and resistance gene recognition. However, it remains to be elucidated how ToMV MP plays such diverse roles in plants. Here, we show that ToMV MP interacts with the Rubisco small subunit (RbCS) of Nicotiana benthamiana in vitro and in vivo. In susceptible N. benthamiana plants, silencing of NbRbCS enabled ToMV to induce necrosis in inoculated leaves, thus enhancing virus local infectivity. However, the development of systemic viral symptoms was delayed. In transgenic N. benthamiana plants harboring Tobacco mosaic virus resistance-22 (Tm-22), which mediates extreme resistance to ToMV, silencing of NbRbCS compromised Tm-22-dependent resistance. ToMV was able to establish efficient local infection but was not able to move systemically. These findings suggest that NbRbCS plays a vital role in tobamovirus movement and plant antiviral defenses.Plant viruses use at least one movement protein (MP) to facilitate viral spread between plant cells via plasmodesmata (PD; Lucas and Gilbertson, 1994; Ghoshroy et al., 1997). Among viral MPs, the MP of tobamoviruses, such as Tobacco mosaic virus (TMV) and its close relative Tomato mosaic virus (ToMV), is the best characterized. TMV MP specifically accumulates in PD and modifies the plasmodesmatal size exclusion limit in mature source leaves or tissues (Wolf et al., 1989; Deom et al., 1990; Ding et al., 1992). TMV MP and viral genomic RNA form a mobile ribonucleoprotein complex that is essential for cell-to-cell movement of viral infection (Watanabe et al., 1984; Deom et al., 1987; Citovsky et al., 1990, 1992; Kiselyova et al., 2001; Kawakami et al., 2004; Waigmann et al., 2007). TMV MP also enhances intercellular RNA silencing (Vogler et al., 2008) and affects viral symptom development, host range, and host susceptibility to virus (Dardick et al., 2000; Bazzini et al., 2007). Furthermore, ToMV MP is identified as an avirulence factor that is recognized by tomato (Solanum lycopersicum) resistance proteins Tobacco mosaic virus resistance-2 (Tm-2) and Tm-22 (Meshi et al., 1989; Lanfermeijer et al., 2004). Indeed, tomato Tm-22 confers extreme resistance against TMV and ToMV in tomato plants and even in heterologous tobacco (Nicotiana tabacum) plants (Lanfermeijer et al., 2003, 2004).To date, several host factors that interact with TMV MP have been identified. These TMV MP-binding host factors include cell wall-associated proteins such as pectin methylesterase (Chen et al., 2000), calreticulin (Meshi et al., 1989), ANK1 (Ueki et al., 2010), and the cellular DnaJ-like protein MPIP1 (Shimizu et al., 2009). Many cytoskeletal components such as actin filaments (McLean et al., 1995), microtubules (Heinlein et al., 1995), and the microtubule-associated proteins MPB2C (Kragler et al., 2003) and EB1a (Brandner et al., 2008) also interact with TMV MP. Most of these factors are involved in TMV cell-to-cell movement.Rubisco catalyzes the first step of CO2 assimilation in photosynthesis and photorespiration. The Rubisco holoenzyme is a heteropolymer consisting of eight large subunits (RbCLs) and eight small subunits (RbCSs). RbCL was reported to interact with the coat protein of Potato virus Y (Feki et al., 2005). Both RbCS and RbCL were reported to interact with the P3 proteins encoded by several potyviruses, including Shallot yellow stripe virus, Onion yellow dwarf virus, Soybean mosaic virus, and Turnip mosaic virus (Lin et al., 2011). Proteomic analysis of the plant-virus interactome revealed that RbCS participates in the formation of virus complexes of Rice yellow mottle virus (Brizard et al., 2006). However, the biological function of Rubisco in viral infection remains unknown.In this study, we show that RbCS plays an essential role in virus movement, host susceptibility, and Tm-22-mediated extreme resistance in the ToMV-host plant interaction.  相似文献   

13.
14.
15.
16.
The predominant structure of the hemicellulose xyloglucan (XyG) found in the cell walls of dicots is a fucogalactoXyG with an XXXG core motif, whereas in the Poaceae (grasses and cereals), the structure of XyG is less xylosylated (XXGGn core motif) and lacks fucosyl residues. However, specialized tissues of rice (Oryza sativa) also contain fucogalactoXyG. Orthologous genes of the fucogalactoXyG biosynthetic machinery of Arabidopsis (Arabidopsis thaliana) are present in the rice genome. Expression of these rice genes, including fucosyl-, galactosyl-, and acetyltransferases, in the corresponding Arabidopsis mutants confirmed their activity and substrate specificity, indicating that plants in the Poaceae family have the ability to synthesize fucogalactoXyG in vivo. The data presented here provide support for a functional conservation of XyG structure in higher plants.The plant cell wall protects and structurally supports plant cells. The wall consists of a variety of polymers, including polysaccharides, the polyphenol lignin, and glycoproteins. One of the major polysaccharides present in the primary walls (i.e. walls of growing cells) in dicots is xyloglucan (XyG), which consists of a β-1,4-glucan backbone with xylosyl substituents. XyG binds noncovalently to cellulose microfibrils and thereby, is thought to act as a spacer molecule, hindering cellulose microfibrils to aggregate (Carpita and Gibeaut, 1993; Pauly et al., 1999a; Bootten et al., 2004; Cosgrove, 2005; Hayashi and Kaida, 2011; Park and Cosgrove, 2012).The side-chain substitutions on XyG can be structurally diverse depending on plant species, tissue type, and developmental stage of the tissue (Pauly et al., 2001; Hoffman et al., 2005; Peña et al., 2008; Hsieh and Harris, 2009, 2012; Lampugnani et al., 2013; Schultink et al., 2014). A one-letter code nomenclature has been established to specify the XyG side-chain substitutions (Fry et al., 1993; Tuomivaara et al., 20145). According to this nomenclature, an unsubstituted glucosyl residue is indicated by a G, whereas a glucosyl residue substituted with a xylosyl moiety is shown as an X. In most dicots, the xylosyl residue can be further substituted with a galactosyl residue (L), which in turn, can be further decorated with a fucosyl residue (F) and/or an acetyl group (F/L). In some species, the xylosyl residue can be substituted with an arabinosyl moiety (S), and the backbone glucosyl residue can be O-acetylated (G; Jia et al., 2003; Hoffman et al., 2005).Numerous genes have been identified in Arabidopsis (Arabidopsis thaliana) that are involved in fucogalactoXyG biosynthesis (Fig. 1; Pauly et al., 2013; Schultink et al., 2014). The glucan backbone is thought to be synthesized by cellulose synthase-like C (CSLC) family proteins, such as AtCSLC4, as shown by in vitro activity data (Cocuron et al., 2007). Several xylosyltransferases (XXTs) from glycosyl transferase family 34 (GT34) are thought to be responsible for XyG xylosylation. Five of these XXTs in Arabidopsis seem to have XXT activity on XyG in vitro (Faik et al., 2002; Zabotina et al., 2008; Vuttipongchaikij et al., 2012; Mansoori et al., 2015). MURUS3 (MUR3) represents a galactosyltransferase that transfers galactosyl moieties specifically to xylosyl residues adjacent to an unsubstituted glucosyl residue on an XXXG unit, converting it to XXLG, whereas Xyloglucan L-side chain galactosyl Transferase2 (XLT2) was identified as another galactosyltransferase transferring a galactosyl moiety specifically to the second xylosyl residue, resulting in XLXG (Madson et al., 2003; Jensen et al., 2012). Both MUR3 and XLT2 belong to GT47 (Li et al., 2004). MUR2/FUCOSYLTRANSFERASE1 (FUT1) from GT37 was found to harbor fucosyltransferase activity, transferring Fuc from GDP-Fuc to a galactosyl residue adjacent to the unsubstituted glucosyl residue (i.e. onto XXLG but not onto XLXG; Perrin et al., 1999; Vanzin et al., 2002). O-acetylation of the galactosyl residue is mediated by Altered Xyloglucan4 (AXY4) and AXY4L, both of which belong to the Trichome Birefringence-Like (TBL) protein family (Bischoff et al., 2010; Gille et al., 2011; Gille and Pauly, 2012).Open in a separate windowFigure 1.Schematic structures of two types of XyGs and known biosynthetic proteins in Arabidopsis (Hsieh and Harris, 2009; Pauly et al., 2013). The corresponding one-letter code for XyG is shown below the pictograms (Fry et al., 1993; Tuomivaara et al., 2015).XyG found throughout land plants exhibits structural diversity with respect to side-chain substitution patterns (Schultink et al., 2014). Most dicots, such as Arabidopsis, and the noncommelinoid monocots possess a fucogalactoXyG of the XXXG-type XyG structure as shown in Figure 1. However, plant species in the Solanaceae and Poaceae as well as the moss Physcomitrella patens contain a different XyG structure with a reduced level of xylosylation, resulting in an XXGGn core motif (York et al., 1996; Kato et al., 2004; Gibeaut et al., 2005; Jia et al., 2005; Peña et al., 2008; Hsieh and Harris, 2009). In addition, the glucan backbone can be O-acetylated in plants of Solanaceae and Poaceae families (Gibeaut et al., 2005; Jia et al., 2005). XyG from Solanaceae with an XXGG core motif can be further arabinosylated and/or galactosylated (Jia et al., 2005). No XyGs with an XXGGn motif backbone have been reported to be fucosylated.The function of structural diversity of XyG substitutions, such as fucosylation and/or altered xylosylation pattern, remains enigmatic. Removing the terminal fucosyl or acetyl moieties in the corresponding Arabidopsis mutants does not lead to any change in plant growth and development (Vanzin et al., 2002; Gille et al., 2011). However, removing galactosyl residues as well as fucosyl and acetyl moieties in the Arabidopsis xlt2 mur3.1 double mutant results in a dwarfed plant (Jensen et al., 2012; Kong et al., 2015). Replacing the galactosyl moiety with an arabinofuranosyl residue by, for example, expressing a tomato (Solanum lycopersicum) arabinosyltransferase in the Arabidopsis xlt2 mur3.1 mutant rescues the growth phenotype and restores wall biomechanics, indicating that galactosylation and arabinosylation in XyG have an equivalent function (Schultink et al., 2013). Recently, fucosylated XyG structures were found in the pollen tubes of tobacco (Nicotiana alata) and tomato, indicating that fucogalactoXyG is likely also present in other Solanaceae plants, albeit restricted to specific tissues (Lampugnani et al., 2013; Dardelle et al., 2015). Although there is circumstantial evidence that fucogalactoXyG is present in cell suspension cultures of rice (Oryza sativa) and cell suspension cultures of fescue (Festuca arundinaceae; McDougall and Fry, 1994; Peña et al., 2008), fucogalactoXyG has not been found in any physiologically relevant plant tissues of members of the Poaceae (Kato et al., 1982; Watanabe et al., 1984; Gibeaut et al., 2005; Hsieh and Harris, 2009; Brennan and Harris, 2011). Here, we provide chemical and genetic evidence that fucogalactoXyG is, indeed, present in plant tissues of a grass (rice) and prove that the rice genome harbors the genes that could be part of the synthetic machinery necessary to produce fucogalactoXyG.  相似文献   

17.
Tomato (Solanum lycopersicum), like other Solanaceous species, accumulates high levels of antioxidant caffeoylquinic acids, which are strong bioactive molecules and protect plants against biotic and abiotic stresses. Among these compounds, the monocaffeoylquinic acids (e.g. chlorogenic acid [CGA]) and the dicaffeoylquinic acids (diCQAs) have been found to possess marked antioxidative properties. Thus, they are of therapeutic interest both as phytonutrients in foods and as pharmaceuticals. Strategies to increase diCQA content in plants have been hampered by the modest understanding of their biosynthesis and whether the same pathway exists in different plant species. Incubation of CGA with crude extracts of tomato fruits led to the formation of two new products, which were identified by liquid chromatography-mass spectrometry as diCQAs. This chlorogenate:chlorogenate transferase activity was partially purified from ripe fruit. The final protein fraction resulted in 388-fold enrichment of activity and was subjected to trypsin digestion and mass spectrometric sequencing: a hydroxycinnamoyl-Coenzyme A:quinate hydroxycinnamoyl transferase (HQT) was selected as a candidate protein. Assay of recombinant HQT protein expressed in Escherichia coli confirmed its ability to synthesize diCQAs in vitro. This second activity (chlorogenate:chlorogenate transferase) of HQT had a low pH optimum and a high Km for its substrate, CGA. High concentrations of CGA and relatively low pH occur in the vacuoles of plant cells. Transient assays demonstrated that tomato HQT localizes to the vacuole as well as to the cytoplasm of plant cells, supporting the idea that in this species, the enzyme catalyzes different reactions in two subcellular compartments.The importance of plant-based foods in preventing or reducing the risk of chronic disease has been widely demonstrated (Martin et al., 2011, 2013). In addition to vitamins, a large number of other nutrients in plant-based foods promote health and reduce the risk of chronic diseases; these are often referred to as phytonutrients. The presence of phytonutrients in fruit and vegetables is of significant nutritional and therapeutic importance, as many have been found to possess strong antioxidant activity (Rice-Evans et al., 1997). Phenolics are the most widespread dietary antioxidants and caffeoylquinic acids, such as chlorogenic acid (CGA), dicaffeoylquinic acids (diCQAs), and tricaffeoylquinic acids (triCQAs), play important roles in promoting health (Clifford, 1999; Niggeweg et al., 2004). CGA limits low density lipid oxidation (Meyer et al., 1998), diCQAs possess antihepatotoxic activity (Choi et al., 2005), and triCQAs reduce the blood Glc levels of diabetic rats (Islam, 2006). diCQA derivatives have been shown to protect humans from various kinds of diseases; diCQAs suppress melanogenesis effectively (Kaul and Khanduja, 1998), show anti-inflammatory activity in vitro (Peluso et al., 1995), and exhibit a selective inhibition of HIV replication (McDougall et al., 1998). The physiological effects of caffeoylquinic acid derivatives with multiple caffeoyl groups are generally greater than those of monocaffeoylquinic acids, perhaps because the antioxidant activity is largely determined by the number of hydroxyl groups present on the aromatic rings (Wang et al., 2003; Islam, 2006). Furthermore, both diCQAs and triCQAs may function as inhibitors of the activity of HIV integrase, which catalyzes the insertion of viral DNA into the genome of host cells (McDougall et al., 1998; Slanina et al., 2001; Gu et al., 2007).CGA is the major soluble phenolic in Solanaceous crops (Clifford, 1999) and the major antioxidant in the average U.S. diet (Luo et al., 2008), while different isomers of diCQAs have been identified in many crops such as coffee (Coffea canephora), globe artichoke (Cynara cardunculus), tomato (Solanum lycopersicum), lettuce (Lactuca sativa), and sweet potato (Ipomoea batatas; Clifford, 1999; Islam, 2006; Moco et al., 2006, 2007; Moglia et al., 2008). In tomato, CGA accounts for 75% and 35% of the total phenolics in mature green and ripe fruit, respectively, amounting to 2 to 40 mg 100 g–1 dry weight (DW), although levels decline after ripening and during postharvest storage (Slimestad and Verheul, 2009). diCQAs and triCQAs also accumulate in tomato fruit (diCQAs, approximately 2 mg 100 g–1 DW; and triCQAs, 1–2 mg 100 g–1 DW; Chanforan et al., 2012).Three pathways (Villegas and Kojima, 1986; Hoffmann et al., 2003; Niggeweg et al., 2004) have been proposed for the synthesis of CGA: (1) the direct pathway involving caffeoyl-CoA transesterification with quinic acid by hydroxycinnamoyl-Coenzyme A:quinate hydroxycinnamoyl transferase (HQT; Niggeweg et al., 2004; Comino et al., 2009; Menin et al., 2010; Sonnante et al., 2010); (2) the route by which p-coumaroyl-CoA is first transesterified with quinic acid via hydroxycinnamoyl-Coenzyme A transferase (HCT) acyltransferase (Hoffmann et al., 2003; Comino et al., 2007), followed by the hydroxylation of p-coumaroyl quinate to 5-caffeoylquinic acid, catalyzed by C3′H (p-coumaroyl-3-hydroxylase; Schoch et al., 2001; Mahesh et al., 2007; Moglia et al., 2009); and (3) the use of caffeoyl-glucoside as the acyl-donor (Villegas and Kojima, 1986). In tomato, the synthesis of CGA involves transesterification of caffeoyl-CoA with quinic acid by HQT (Niggeweg et al., 2004).To date, it is not clear whether diCQAs are derived directly from the monocaffeoylquinic acids (such as CGA) through a second acyltransferase reaction involving an acyl-CoA or not, although their structural similarity provides good a priori evidence supporting this hypothesis. Recently the in vitro synthesis of 3,5-diCQA from CGA and CoA by HCT from coffee has been reported (Lallemand et al., 2012). By contrast, in sweet potato, an enzyme that catalyzes the transfer of the caffeoyl moiety of CGA to another molecule of CGA, leading to the synthesis of isochlorogenate (3,5-di-O-caffeoylquinate), has been described, but the corresponding gene has not been identified (Villegas and Kojima, 1986).We report a chlorogenate:chlorogenate transferase (CCT) activity leading to the synthesis of diCQAs in tomato fruits and describe how alternative catalysis, by a single enzyme, leads to the production of both CGA and diCQA in different cellular compartments.  相似文献   

18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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