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.
Linear, branch-chained triterpenes, including squalene (C30), botryococcene (C30), and their methylated derivatives (C31–C37), generated by the green alga Botryococcus braunii race B have received significant attention because of their utility as chemical and biofuel feedstocks. However, the slow growth habit of B. braunii makes it impractical as a production system. In this study, we evaluated the potential of generating high levels of botryococcene in tobacco (Nicotiana tabacum) plants by diverting carbon flux from the cytosolic mevalonate pathway or the plastidic methylerythritol phosphate pathway by the targeted overexpression of an avian farnesyl diphosphate synthase along with two versions of botryococcene synthases. Up to 544 µg g−1 fresh weight of botryococcene was achieved when this metabolism was directed to the chloroplasts, which is approximately 90 times greater than that accumulating in plants engineered for cytosolic production. To test if methylated triterpenes could be produced in tobacco, we also engineered triterpene methyltransferases (TMTs) from B. braunii into wild-type plants and transgenic lines selected for high-level triterpene accumulation. Up to 91% of the total triterpene contents could be converted to methylated forms (C31 and C32) by cotargeting the TMTs and triterpene biosynthesis to the chloroplasts, whereas only 4% to 14% of total triterpenes were methylated when this metabolism was directed to the cytoplasm. When the TMTs were overexpressed in the cytoplasm of wild-type plants, up to 72% of the total squalene was methylated, and total triterpene (C30+C31+C32) content was elevated 7-fold. Altogether, these results point to innate mechanisms controlling metabolite fluxes, including a homeostatic role for squalene.Terpenes and terpenoids represent a distinct class of natural products (Buckingham, 2003) that are derived from two universal five-carbon precursors: isopentenyl diphosphate (IPP) and dimethylallyl diphosphate (DMAPP). In eukaryotic fungi and animals, IPP and DMAPP are synthesized via the mevalonate (MVA) pathway, whereas in prokaryotes, they are synthesized via the methylerythritol phosphate (MEP) pathway. In higher plants, the pathways are present in separate compartments and are believed to operate independently. The MVA pathway in the cytoplasm is predominantly responsible for sesquiterpene (C15), triterpene (C30), and polyprenol (greater than C45) biosynthesis and associated with the endoplasmic reticulum (ER) system. The MEP pathway resides in plastids and is dedicated to monoterpenes (C10), diterpenes (C20), carotenoids (C40), and long-chain phytol biosynthesis. All these compounds are usually produced by plants for a variety of physiological (i.e. hormones, aliphatic membrane anchors, and maintaining membrane structure) and ecological (i.e. defense compounds and insect/animal attractants) roles (Kempinski et al., 2015). Terpenes are also important for various industrial applications, ranging from flavors and fragrances (Schwab et al., 2008) to medicines (Dewick, 2009; Niehaus et al., 2011; Shelar, 2011).The utility of terpenes as chemical and biofuel feedstocks has also received considerable attention recently. Isoprenoid-derived biofuels include farnesane (Renninger and McPhee, 2008; Rude and Schirmer, 2009), bisabolene (Peralta-Yahya et al., 2011), pinene dimers (Harvey et al., 2010), isopentenal (Withers et al., 2007), and botryococcene (Moldowan and Seifert, 1980; Hillen et al., 1982; Glikson et al., 1989; Mastalerz and Hower, 1996). The richness of branches within these hydrocarbon scaffolds correlate with their high-energy content, which enables them to serve as suitable alternatives to crude petroleum (Peralta-Yahya and Keasling, 2010). Indeed, some of them are already major contributors to current-day petroleum-based fuels. One of the best examples of this is the triterpene oil accumulating in the green alga Botryococcus braunii race B, which is considered a major progenitor to oil and coal shale deposits (Moldowan and Seifert, 1980). This alga has been well studied, and the major constituents of its prodigious hydrocarbon oil are a group of triterpenes including squalene (C30), organism-specific botryococcene (C30), methylated squalene (C31–C34), and methylated botryococcene (C31–C37; Metzger et al., 1988; Huang and Poulter, 1989; Okada et al., 1995), which can be readily converted into all classes of combustible fuels under hydrocracking conditions (Hillen et al., 1982).The unique biosynthetic mechanism for the triterpenes in B. braunii was recently described by Niehaus et al. (2011), and a series of novel squalene synthase-like genes were identified (Fig. 1). In short, squalene synthase-like enzyme, SSL-1, performs a head-to-head condensation of two farnesyl diphosphate (FPP) molecules into presqualene diphosphate, followed by a reductive rearrangement to yield squalene (C30) by the enzyme SSL-2, or is converted by SSL-3 to form botryococcene through a different reductive rearrangement (Niehaus et al., 2011). Methylated derivatives are the dominant triterpene species generated by B. braunii race B (Metzger, 1985; Metzger et al., 1988), and these derivatives are known to yield higher quality fuels due to their high energy content and the hydrocracking products derived by virtue of having more hydrocarbon branches. Triterpene methyltransferases (TMTs) that can methylate squalene and botryococcene have been successfully characterized by Niehaus et al. (2012). TRITERPENE METHYLTRANSFERASE1 (TMT-1) and TMT-2 prefer squalene C30 as their substrate for the production of monomethylated (C31) or dimethylated (C32) squalene, while TMT-3 prefers botryococcene as its substrate for the biosynthesis of monomethylated (C31) or dimethylated (C32) botryococcene (Fig. 1). These TMTs are believed to be insoluble enzymes; they exhibit large hydrophobic areas, and their activities were only observed in vitro using yeast microsomal preparations (no activity was observed when expressed in bacteria; Niehaus et al., 2012).Open in a separate windowFigure 1.Depiction of the catalytic roles of novel SSL and TMT enzymes in B. braunii race B and their putative contributions to the triterpene constituents (Niehaus et al., 2011; Niehaus et al., 2012). SSL-1 catalyzes the condensation of two farnesyl diphosphate (FPP) molecules to presqualene diphosphate (PSPP), which is converted to either squalene or botryococcene by SSL-2 or SSL-3, respectively. Squalene can also be synthesized directly from the condensation of two FPP molecules catalyzed by squalene synthase (SQS). TMT-1 and TMT-2 transfer the methyl donor group from S-adenosylmethionine (SAM) to squalene to form monomethylated and dimethylated squalene, whereas TMT-3 acts on botryococcene to form monomethylated and dimethylated botryococcene (Niehaus et al., 2012).Like the majority of identified methyltransferases, these TMTs utilize the methyl donor S-adenosyl methionine (SAM), which is ubiquitous in prokaryotes and eukaryotes (Scheer et al., 2011; Liscombe et al., 2012). In plants, SAM is one of the most abundant cofactors (Fontecave et al., 2004; Sauter et al., 2013) and is synthesized exclusively in the cytosol (Wallsgrove et al., 1983; Ravanel et al., 1998, 2004; Bouvier et al., 2006). While it is used predominantly as a methyl donor in the methylation reaction (Ravanel et al., 2004), it also serves as the primary precursor for the biosynthesis of ethylene (Wang et al., 2002b), polyamines (Kusano et al., 2008), and nicotianamine (Takahashi et al., 2003), which play a variety of important roles for plant growth and development (Huang et al., 2012; Sauter et al., 2013). The SAM present in organelles, like the chloroplast, appears to be imported from the cytosol by specific SAM/S-adenosylhomocysteine exchange transporters that reside on the envelope membranes of plastids (Ravanel et al., 2004; Bouvier et al., 2006). The imported SAM is involved in the biogenesis of Asp-derived amino acids (Curien et al., 1998; Jander and Joshi, 2009; Sauter et al., 2013) and serves as the methyl donor for the methylation of macromolecules, such as plastid DNA (Nishiyama et al., 2002; Ahlert et al., 2009) and proteins (Houtz et al., 1989; Niemi et al., 1990; Ying et al., 1999; Trievel et al., 2003; Alban et al., 2014), and small molecule metabolites, such as prenylipids (e.g. plastoquinone, tocopherol, chlorophylls, and phylloquinone; Bouvier et al., 2005, 2006; DellaPenna, 2005).Although plants and microbes are the natural sources for useful terpenes, most of them are produced in very small amounts and often as complex mixtures. In contrast, B. braunii produces large quantities of triterpenes, but its slow growth makes it undesirable as a viable production platform (Niehaus et al., 2011). Nevertheless, metabolic engineering and synthetic biology offer many strategies to manipulate terpene metabolism in various biological systems to achieve high-value terpene production with high yield and high fidelity for particular practical applications (Nielsen and Keasling, 2011). Many successes have been achieved in engineering valuable terpenes in heterotrophic microbes, such as Escherichia coli (Nishiyama et al., 2002; Martin et al., 2003; Ajikumar et al., 2010) and Saccharomyces cerevisiae (Ro et al., 2006; Takahashi et al., 2007; Westfall et al., 2012; Zhuang and Chappell, 2015). The strategies developed in these efforts usually take advantage of specific microbe strains whose innate biosynthetic machinery is genetically modified to accumulate certain prenyldiphosphate precursors (e.g. IPP or FPP), which can be utilized by other introduced terpene synthase(s) for the production of the desired terpene(s). For example, greater than 900 mg L−1 bisabolene was produced when bisabolene synthase genes from plants were introduced into FPP-overproducing E. coli or S. cerevisiae strains (Peralta-Yahya et al., 2011). High levels of farnesane production for diesel fuels were also achieved by reductive hydrogenation of its precursor farnesene, which was generated from a genetically engineered yeast (e.g. Saccharomyces cerevisiae) strain using plant farnesene synthases (Renninger and McPhee, 2008; Ubersax and Platt, 2010). However, terpene production using microbial platforms is still dependent on exogenous feedstocks (i.e. sugars) and elaborate production facilities, both of which add significantly to their production costs.Compared with microbial systems, engineering terpene production in plant systems seems like an attractive target as well. This is because plants can take advantage of photosynthesis by using atmospheric CO2 as their carbon resource instead of relying on exogenous carbon feedstocks. Moreover, crop plants such as tobacco (Nicotiana tabacum) can generate a large amount of green tissues efficiently when grown for biomass production (Schillberg et al., 2003; Andrianov et al., 2010), making them a robust, sustainable, and scalable platform for large-scale terpene production. Nonetheless, compared with microbial platforms, there are only a few examples of elevating terpene production in bioengineered plants. This is due partly to higher plants being complex multicellular organisms, in which terpene metabolism generally utilizes more complex innate machinery that can be compartmentalized intracellularly and to cell/tissue specificities (Lange and Ahkami, 2013; Kempinski et al., 2015). Significant efforts have been made to overcome these obstacles to improve the production of valuable terpenes in plants, including monoterpenes (Lücker et al., 2004; Ohara et al., 2010; Lange et al., 2011), sesquiterpenes (Aharoni et al., 2003; Kappers et al., 2005; Wu et al., 2006; Davidovich-Rikanati et al., 2008), diterpenes (Besumbes et al., 2004; Anterola et al., 2009), and triterpenes (Inagaki et al., 2011; Wu et al., 2012). Among these, engineering terpene metabolism into a subcellular organelle, where the engineered enzymes/pathways can utilize unlimited/unregulated precursors as substrates, appears most successful. For example, Wu et al. (2006, 2012) expressed an avian farnesyl diphosphate synthase (FPS) with foreign sesquiterpene/triterpene synthases targeted to the plastid to divert the IPP/DMAPP pool from the plastidic MEP pathway to synthesize high levels of the novel sesquiterpenes patchoulol and amorpha-4,11-diene up to 30 µg g−1 fresh weight and the triterpene squalene up to 1,000 µg g−1 fresh weight. This strategy appears to be particularly robust because it avoids possible endogenous regulation of sesquiterpene and triterpene biosynthesis, which occurs normally in the cytoplasm, and relies upon more plastic precursor pools of IPP/DMAPP inherent in the plastid, which are primarily derived from the local CO2 fixation (Wright et al., 2014).The goal of this study was to evaluate the prospects for engineering advanced features of triterpene metabolism from B. braunii into tobacco and, thus, to probe the innate intricacies of isoprenoid metabolism in plants. In order to achieve this, we first introduced the key steps of botryococcene biosynthesis into specific subcellular compartments of tobacco cells under the direction of constitutive or trichome-specific promoters. The transgenic lines expressing the enzymes in the chloroplast were found to accumulate the highest levels of botryococcene. Triterpene methyltransferases were next introduced into the same intracellular compartments of selected high-triterpene-accumulating lines. A high yield of methylated triterpenes was also achieved in transgenic lines when the TMTs were targeted to the chloroplast. Through careful comparison of the levels of triterpenes and the methylated triterpene products in the various transgenic lines, we have also gained a deeper insight into the subcellular distribution of the triterpene products in these transgenic lines as well as a better understanding of methylation metabolism for specialized metabolites in particular compartments. These findings all contribute to our understanding of the regulatory elements that control carbon flux through the innate terpene biosynthetic pathways operating in plants.  相似文献   

3.
Trehalose is a nonreducing sugar used as a reserve carbohydrate and stress protectant in a variety of organisms. While higher plants typically do not accumulate high levels of trehalose, they encode large families of putative trehalose biosynthesis genes. Trehalose biosynthesis in plants involves a two-step reaction in which trehalose-6-phosphate (T6P) is synthesized from UDP-glucose and glucose-6-phosphate (catalyzed by T6P synthase [TPS]), and subsequently dephosphorylated to produce the disaccharide trehalose (catalyzed by T6P phosphatase [TPP]). In Arabidopsis (Arabidopsis thaliana), 11 genes encode proteins with both TPS- and TPP-like domains but only one of these (AtTPS1) appears to be an active (TPS) enzyme. In addition, plants contain a large family of smaller proteins with a conserved TPP domain. Here, we present an in-depth analysis of the 10 TPP genes and gene products in Arabidopsis (TPPA-TPPJ). Collinearity analysis revealed that all of these genes originate from whole-genome duplication events. Heterologous expression in yeast (Saccharomyces cerevisiae) showed that all encode active TPP enzymes with an essential role for some conserved residues in the catalytic domain. These results suggest that the TPP genes function in the regulation of T6P levels, with T6P emerging as a novel key regulator of growth and development in higher plants. Extensive gene expression analyses using a complete set of promoter-β-glucuronidase/green fluorescent protein reporter lines further uncovered cell- and tissue-specific expression patterns, conferring spatiotemporal control of trehalose metabolism. Consistently, phenotypic characterization of knockdown and overexpression lines of a single TPP, AtTPPG, points to unique properties of individual TPPs in Arabidopsis, and underlines the intimate connection between trehalose metabolism and abscisic acid signaling.The presence of trehalose in a wide variety of organisms and the existence of different biosynthesis pathways suggest a pivotal and ancient role for trehalose metabolism in nature. The most widely distributed metabolic pathway consists of two consecutive enzymatic reactions, with trehalose-6-phosphate (T6P) synthase (TPS) catalyzing the transfer of a glucosyl moiety from UDP-Glc to Glc-6-phosphate to produce T6P and UDP, and T6P phosphatase (TPP) catalyzing dephosphorylation of T6P to trehalose (Cabib and Leloir, 1958; Avonce et al., 2006). Apart from operating as a (reserve) carbon source and structural component in bacteria, fungi, and invertebrates, trehalose also functions as a major stress protectant of proteins and membranes during adverse conditions such as dehydration, high salinity, hypoxia, and nutrient starvation (Elbein et al., 2003). Trehalose accumulation is also observed in a few lower vascular resurrection plants (e.g. Selaginella lepidophylla). Until about a decade ago, higher vascular plants were believed to have lost the ability to produce trehalose, but with the emergence of more sensitive assays, genome sequencing, and the use of yeast (Saccharomyces cerevisiae) mutant complementation, minute amounts of trehalose and T6P, and functional plant enzyme orthologs were found (Goddijn et al., 1997; Vogel et al., 1998; Lunn et al., 2006). In addition, heterologous expression and disruption of trehalose metabolism in plants conferred pleiotropic effects, ranging from altered stress tolerance, leaf morphology, and developmental timing to embryo lethality (Holmström et al., 1996; Goddijn et al., 1997; Romero et al., 1997; Eastmond et al., 2002; Schluepmann et al., 2003; Avonce et al., 2004; Satoh-Nagasawa et al., 2006; Miranda et al., 2007; Chary et al., 2008), pointing to an important regulatory function. The intermediate T6P has been highlighted as a novel signal for carbohydrate status (for review, see Paul, 2008), positively correlating with Suc levels, redox-regulated ADP-Glc pyrophosphorylase activity, and starch biosynthesis (Lunn et al., 2006). Recently, it was reported that T6P inhibits the activity of the SnRK1 protein kinase to activate energy-consuming biosynthetic processes in growing tissue (Zhang et al., 2009) and that it is required for the onset of leaf senescence (Wingler et al., 2012).In most bacterial and eukaryotic species, the TPS and TPP activities are found on separate proteins. Recent phylogenetic and biochemical analyses showed that some archaea and bacteria, such as Cytophaga hutchinsonii, express proteins that have both active TPS and TPP domains resulting from gene fusion, suggesting that such prokaryotic bifunctional proteins are the evolutionary ancestors of the large eukaryotic trehalose biosynthesis enzymes in which one or both domains have subsequently lost their catalytic activity (Avonce et al., 2010). The yeast TPP enzyme Tps2, for example, harbors an inactive N-terminal TPS domain and an active C-terminal TPP domain. In contrast to the single TPS and TPP genes in most microorganisms, the genomes of higher plants encode a remarkably large family of putative trehalose biosynthesis enzyme homologs. These are commonly classified in three distinct subgroups, according to their similarity to the microbial TPS and TPP proteins and/or presence of specific motifs (e.g. conserved phosphatase boxes; Thaller et al., 1998; Leyman et al., 2001; Eastmond et al., 2003). Even primitive plants such as the alga Ostreococcus tauri and the moss Physcomitrella patens already contain members of each of these gene families, pointing to the early establishment and conservation of these proteins in plant evolution (Lunn, 2007; Avonce et al., 2010). In Arabidopsis (Arabidopsis thaliana), the class I TPS proteins (AtTPS1-4) show most similarity to the yeast TPS Tps1, but also have a C-terminal domain with limited similarity to TPPs. However, only one of these, AtTPS1, appears to have heterologous enzymatic TPS activity in yeast (Blázquez et al., 1998; Vandesteene et al., 2010). Strikingly, AtTPS1 is the only class I enzyme with an N-terminal extension that seems to operate as an autoinhibitory domain (Van Dijck et al., 2002). The class II TPS proteins (AtTPS5-11) are similar bipartite proteins with a TPS-like domain but a more conserved TPP domain. They appear to lack both heterologous TPS and TPP activity (Ramon et al., 2009). The high level of conservation of putative substrate-binding residues in class I and class II proteins, however, suggests that substrates might still bind (Avonce et al., 2006; Lunn, 2007; Ramon et al., 2009; Vandesteene et al., 2010). Together with the specific expression patterns of the class I genes (van Dijken et al., 2004; Geelen et al., 2007; Vandesteene et al., 2010) and the extensive expression regulation of all class II members by plant carbon status (Baena-González et al., 2007; Usadel et al., 2008; Ramon et al., 2009), this suggests tissue-specific regulatory functions for these proteins in metabolic regulation of plant growth and development. Finally, Arabidopsis also harbors a family of 10 smaller proteins (AtTPPA-J; 320–385 amino acids) with limited similarity to the class I and class II proteins (795–942 amino acids). Like class II proteins, they contain the phosphatase box consensus sequences, characteristic of the l-2-haloacid dehalogenase (HAD) super family of enzymes, which includes a wide range of phosphatases and hydrolases (Thaller et al., 1998). It has been suggested that the origin of these plant TPP genes is different from the origin of the class I and II genes (Avonce et al., 2010) and that plants recruited the TPP genes after their divergence from fungi, most probably from proteobacteria or actinobacteria. Consistently, homologous TPP proteins are present in proteobacteria such as Rhodopherax ferrireducens (Avonce et al., 2010). To date, only a few of these single-domain plant TPP proteins have been subject to biochemical characterization, e.g. TPPA and TPPB from Arabidopsis (Vogel et al., 1998), OsTPP1 and OsTPP2 from rice (Oryza sativa; Pramanik and Imai, 2005; Shima et al., 2007), and RAMOSA3 (RA3) from maize (Zea mays; Satoh-Nagasawa et al., 2006).The phenotypic alterations observed in plants fed with trehalose or genetically modified in trehalose biosynthesis, suggest a pivotal role for trehalose metabolism in integrating the metabolic status with growth and development. Disruption of the only known active TPS enzyme in Arabidopsis (AtTPS1) results in embryo lethality (Eastmond et al., 2002) and, when rescued to bridge embryogenesis, causes a strong disruption of vegetative and generative development and abscisic acid (ABA) hypersensitivity (van Dijken et al., 2004; Gómez et al., 2010). Overexpressing AtTPS1 on the other hand renders seedlings sugar and ABA insensitive (Avonce et al., 2004, 2005). These observations strongly link trehalose metabolism with ABA signaling. Interestingly, a mutation of a TPP gene in maize, RA3, results in a distinct phenotype, with incorrect axillary meristem identity and determinacy in both male and female inflorescences (Satoh-Nagasawa et al., 2006). Arabidopsis plants with overall increased T6P levels, such as OtsA (Escherichia coli TPS) overexpression plants, similarly show increased inflorescence branching (Schluepmann et al., 2003; van Dijken et al., 2004).To better understand why higher plants harbor such a large number of putative TPP proteins, we have made a comprehensive study of the 10 Arabidopsis TPP genes and gene products, combining phylogenetic approaches and yeast growth complementation assays, together with a detailed analysis of all 10 TPP gene expression profiles in Arabidopsis, and a more detailed single AtTPP mutant phenotypic analysis.  相似文献   

4.
5.
6.
Necrotrophic and biotrophic pathogens are resisted by different plant defenses. While necrotrophic pathogens are sensitive to jasmonic acid (JA)-dependent resistance, biotrophic pathogens are resisted by salicylic acid (SA)- and reactive oxygen species (ROS)-dependent resistance. Although many pathogens switch from biotrophy to necrotrophy during infection, little is known about the signals triggering this transition. This study is based on the observation that the early colonization pattern and symptom development by the ascomycete pathogen Plectosphaerella cucumerina (P. cucumerina) vary between inoculation methods. Using the Arabidopsis (Arabidopsis thaliana) defense response as a proxy for infection strategy, we examined whether P. cucumerina alternates between hemibiotrophic and necrotrophic lifestyles, depending on initial spore density and distribution on the leaf surface. Untargeted metabolome analysis revealed profound differences in metabolic defense signatures upon different inoculation methods. Quantification of JA and SA, marker gene expression, and cell death confirmed that infection from high spore densities activates JA-dependent defenses with excessive cell death, while infection from low spore densities induces SA-dependent defenses with lower levels of cell death. Phenotyping of Arabidopsis mutants in JA, SA, and ROS signaling confirmed that P. cucumerina is differentially resisted by JA- and SA/ROS-dependent defenses, depending on initial spore density and distribution on the leaf. Furthermore, in situ staining for early callose deposition at the infection sites revealed that necrotrophy by P. cucumerina is associated with elevated host defense. We conclude that P. cucumerina adapts to early-acting plant defenses by switching from a hemibiotrophic to a necrotrophic infection program, thereby gaining an advantage of immunity-related cell death in the host.Plant pathogens are often classified as necrotrophic or biotrophic, depending on their infection strategy (Glazebrook, 2005; Nishimura and Dangl, 2010). Necrotrophic pathogens kill living host cells and use the decayed plant tissue as a substrate to colonize the plant, whereas biotrophic pathogens parasitize living plant cells by employing effector molecules that suppress the host immune system (Pel and Pieterse, 2013). Despite this binary classification, the majority of pathogenic microbes employ a hemibiotrophic infection strategy, which is characterized by an initial biotrophic phase followed by a necrotrophic infection strategy at later stages of infection (Perfect and Green, 2001). The pathogenic fungi Magnaporthe grisea, Sclerotinia sclerotiorum, and Mycosphaerella graminicola, the oomycete Phytophthora infestans, and the bacterial pathogen Pseudomonas syringae are examples of hemibiotrophic plant pathogens (Perfect and Green, 2001; Koeck et al., 2011; van Kan et al., 2014; Kabbage et al., 2015).Despite considerable progress in our understanding of plant resistance to necrotrophic and biotrophic pathogens (Glazebrook, 2005; Mengiste, 2012; Lai and Mengiste, 2013), recent debate highlights the dynamic and complex interplay between plant-pathogenic microbes and their hosts, which is raising concerns about the use of infection strategies as a static tool to classify plant pathogens. For instance, the fungal genus Botrytis is often labeled as an archetypal necrotroph, even though there is evidence that it can behave as an endophytic fungus with a biotrophic lifestyle (van Kan et al., 2014). The rice blast fungus Magnaporthe oryzae, which is often classified as a hemibiotrophic leaf pathogen (Perfect and Green, 2001; Koeck et al., 2011), can adopt a purely biotrophic lifestyle when infecting root tissues (Marcel et al., 2010). It remains unclear which signals are responsible for the switch from biotrophy to necrotrophy and whether these signals rely solely on the physiological state of the pathogen, or whether host-derived signals play a role as well (Kabbage et al., 2015).The plant hormones salicylic acid (SA) and jasmonic acid (JA) play a central role in the activation of plant defenses (Glazebrook, 2005; Pieterse et al., 2009, 2012). The first evidence that biotrophic and necrotrophic pathogens are resisted by different immune responses came from Thomma et al. (1998), who demonstrated that Arabidopsis (Arabidopsis thaliana) genotypes impaired in SA signaling show enhanced susceptibility to the biotrophic pathogen Hyaloperonospora arabidopsidis (formerly known as Peronospora parastitica), while JA-insensitive genotypes were more susceptible to the necrotrophic fungus Alternaria brassicicola. In subsequent years, the differential effectiveness of SA- and JA-dependent defense mechanisms has been confirmed in different plant-pathogen interactions, while additional plant hormones, such as ethylene, abscisic acid (ABA), auxins, and cytokinins, have emerged as regulators of SA- and JA-dependent defenses (Bari and Jones, 2009; Cao et al., 2011; Pieterse et al., 2012). Moreover, SA- and JA-dependent defense pathways have been shown to act antagonistically on each other, which allows plants to prioritize an appropriate defense response to attack by biotrophic pathogens, necrotrophic pathogens, or herbivores (Koornneef and Pieterse, 2008; Pieterse et al., 2009; Verhage et al., 2010).In addition to plant hormones, reactive oxygen species (ROS) play an important regulatory role in plant defenses (Torres et al., 2006; Lehmann et al., 2015). Within minutes after the perception of pathogen-associated molecular patterns, NADPH oxidases and apoplastic peroxidases generate early ROS bursts (Torres et al., 2002; Daudi et al., 2012; O’Brien et al., 2012), which activate downstream defense signaling cascades (Apel and Hirt, 2004; Torres et al., 2006; Miller et al., 2009; Mittler et al., 2011; Lehmann et al., 2015). ROS play an important regulatory role in the deposition of callose (Luna et al., 2011; Pastor et al., 2013) and can also stimulate SA-dependent defenses (Chaouch et al., 2010; Yun and Chen, 2011; Wang et al., 2014; Mammarella et al., 2015). However, the spread of SA-induced apoptosis during hyperstimulation of the plant immune system is contained by the ROS-generating NADPH oxidase RBOHD (Torres et al., 2005), presumably to allow for the sufficient generation of SA-dependent defense signals from living cells that are adjacent to apoptotic cells. Nitric oxide (NO) plays an additional role in the regulation of SA/ROS-dependent defense (Trapet et al., 2015). This gaseous molecule can stimulate ROS production and cell death in the absence of SA while preventing excessive ROS production at high cellular SA levels via S-nitrosylation of RBOHD (Yun et al., 2011). Recently, it was shown that pathogen-induced accumulation of NO and ROS promotes the production of azelaic acid, a lipid derivative that primes distal plants for SA-dependent defenses (Wang et al., 2014). Hence, NO, ROS, and SA are intertwined in a complex regulatory network to mount local and systemic resistance against biotrophic pathogens. Interestingly, pathogens with a necrotrophic lifestyle can benefit from ROS/SA-dependent defenses and associated cell death (Govrin and Levine, 2000). For instance, Kabbage et al. (2013) demonstrated that S. sclerotiorum utilizes oxalic acid to repress oxidative defense signaling during initial biotrophic colonization, but it stimulates apoptosis at later stages to advance necrotrophic colonization. Moreover, SA-induced repression of JA-dependent resistance not only benefits necrotrophic pathogens but also hemibiotrophic pathogens after having switched from biotrophy to necrotrophy (Glazebrook, 2005; Pieterse et al., 2009, 2012).Plectosphaerella cucumerina ((P. cucumerina, anamorph Plectosporum tabacinum) anamorph Plectosporum tabacinum) is a filamentous ascomycete fungus that can survive saprophytically in soil by decomposing plant material (Palm et al., 1995). The fungus can cause sudden death and blight disease in a variety of crops (Chen et al., 1999; Harrington et al., 2000). Because P. cucumerina can infect Arabidopsis leaves, the P. cucumerina-Arabidopsis interaction has emerged as a popular model system in which to study plant defense reactions to necrotrophic fungi (Berrocal-Lobo et al., 2002; Ton and Mauch-Mani, 2004; Carlucci et al., 2012; Ramos et al., 2013). Various studies have shown that Arabidopsis deploys a wide range of inducible defense strategies against P. cucumerina, including JA-, SA-, ABA-, and auxin-dependent defenses, glucosinolates (Tierens et al., 2001; Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014), callose deposition (García-Andrade et al., 2011; Gamir et al., 2012, 2014; Sánchez-Vallet et al., 2012), and ROS (Tierens et al., 2002; Sánchez-Vallet et al., 2010; Barna et al., 2012; Gamir et al., 2012, 2014; Pastor et al., 2014). Recent metabolomics studies have revealed large-scale metabolic changes in P. cucumerina-infected Arabidopsis, presumably to mobilize chemical defenses (Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014). Furthermore, various chemical agents have been reported to induce resistance against P. cucumerina. These chemicals include β-amino-butyric acid, which primes callose deposition and SA-dependent defenses, benzothiadiazole (BTH or Bion; Görlach et al., 1996; Ton and Mauch-Mani, 2004), which activates SA-related defenses (Lawton et al., 1996; Ton and Mauch-Mani, 2004; Gamir et al., 2014; Luna et al., 2014), JA (Ton and Mauch-Mani, 2004), and ABA, which primes ROS and callose deposition (Ton and Mauch-Mani, 2004; Pastor et al., 2013). However, among all these studies, there is increasing controversy about the exact signaling pathways and defense responses contributing to plant resistance against P. cucumerina. While it is clear that JA and ethylene contribute to basal resistance against the fungus, the exact roles of SA, ABA, and ROS in P. cucumerina resistance vary between studies (Thomma et al., 1998; Ton and Mauch-Mani, 2004; Sánchez-Vallet et al., 2012; Gamir et al., 2014).This study is based on the observation that the disease phenotype during P. cucumerina infection differs according to the inoculation method used. We provide evidence that the fungus follows a hemibiotrophic infection strategy when infecting from relatively low spore densities on the leaf surface. By contrast, when challenged by localized host defense to relatively high spore densities, the fungus switches to a necrotrophic infection program. Our study has uncovered a novel strategy by which plant-pathogenic fungi can take advantage of the early immune response in the host plant.  相似文献   

7.
8.
9.
Fumarylacetoacetate hydrolase (FAH) hydrolyzes fumarylacetoacetate to fumarate and acetoacetate, the final step in the tyrosine (Tyr) degradation pathway that is essential to animals. Deficiency of FAH in animals results in an inborn lethal disorder. However, the role for the Tyr degradation pathway in plants remains to be elucidated. In this study, we isolated an Arabidopsis (Arabidopsis thaliana) short-day sensitive cell death1 (sscd1) mutant that displays a spontaneous cell death phenotype under short-day conditions. The SSCD1 gene was cloned via a map-based cloning approach and found to encode an Arabidopsis putative FAH. The spontaneous cell death phenotype of the sscd1 mutant was completely eliminated by further knockout of the gene encoding the putative homogentisate dioxygenase, which catalyzes homogentisate into maleylacetoacetate (the antepenultimate step) in the Tyr degradation pathway. Furthermore, treatment of Arabidopsis wild-type seedlings with succinylacetone, an abnormal metabolite caused by loss of FAH in the Tyr degradation pathway, mimicked the sscd1 cell death phenotype. These results demonstrate that disruption of FAH leads to cell death in Arabidopsis and suggest that the Tyr degradation pathway is essential for plant survival under short-day conditions.Programmed cell death (PCD) has been defined as a sequence of genetically regulated events that lead to the elimination of specific cells, tissues, or whole organs (Lockshin and Zakeri, 2004). In plants, PCD is essential for developmental processes and defense responses (Dangl et al., 1996; Greenberg, 1996; Durrant et al., 2007). One well-characterized example of plant PCD is the hypersensitive response occurring during incompatible plant-pathogen interactions (Lam, 2004), which results in cell death to form visible lesions at the site of infection by an avirulent pathogen and consequently limits the pathogen spread (Morel and Dangl, 1997).To date, a large number of mutants that display spontaneous cell death lesions have been identified in barley (Hordeum vulgare), maize (Zea mays), rice (Oryza sativa), and Arabidopsis (Arabidopsis thaliana; Marchetti et al., 1983; Wolter et al., 1993; Dietrich et al., 1994; Gray et al., 1997). Because lesions form in the absence of pathogen infection, these mutants have been collectively termed as lesion-mimic mutants. Many genes with regulatory roles in PCD and defense responses, including LESION SIMULATING DISEASE1, ACCELERATED CELL DEATH11, and VASCULAR ASSOCIATED DEATH1, have been cloned and characterized (Dietrich et al., 1997; Brodersen et al., 2002; Lorrain et al., 2004).The appearance of spontaneous cell death lesions in some lesion-mimic mutants is dependent on photoperiod. For example, the Arabidopsis mutant lesion simulating disease1 and myoinositol-1-phosphate synthase1 show lesions under long days (LD; Dietrich et al., 1994; Meng et al., 2009), whereas the lesion simulating disease2, lesion initiation1, enhancing RPW8-mediated HR-like cell death1, and lag one homolog1 display lesions under short days (SD; Dietrich et al., 1994; Ishikawa et al., 2003; Wang et al., 2008; Ternes et al., 2011).Blockage of some metabolic pathways in plants may cause cell death and result in lesion formation. For example, the lesion-mimic phenotypes in the Arabidopsis mutants lesion initiation2 and accelerated cell death2 and the maize mutant lesion mimic22 result from an impairment of porphyrin metabolism (Hu et al., 1998; Ishikawa et al., 2001; Mach et al., 2001). Deficiency in fatty acid, sphingolipid, and myoinositol metabolism also causes cell death in Arabidopsis (Mou et al., 2000; Liang et al., 2003; Wang et al., 2008; Meng et al., 2009; Donahue et al., 2010; Berkey et al., 2012).Tyr degradation is an essential five-step pathway in animals (Lindblad et al., 1977). First, Tyr aminotransferase catalyzes the conversion of Tyr into 4-hydroxyphenylpyruvate, which is further transformed into homogentisate by 4-hydroxyphenylpyruvate dioxygenase. Through the sequential action of homogentisate dioxygenase (HGO), maleylacetoacetate isomerase (MAAI), and fumarylacetoacetate hydrolase (FAH), homogentisate is catalyzed to generate fumarate and acetoacetate (Lindblad et al., 1977). Blockage of this pathway in animals results in metabolic disorder diseases (Lindblad et al., 1977; Ruppert et al., 1992; Grompe et al., 1993). For example, human FAH deficiency causes hereditary tyrosinemia type I (HT1), an inborn lethal disease (St-Louis and Tanguay, 1997). Although the homologous genes putatively encoding these enzymes exist in plants (Dixon et al., 2000; Lopukhina et al., 2001; Dixon and Edwards, 2006), it is unclear whether this pathway is essential for plant growth and development.In this study, we report the isolation and characterization of a recessive short-day sensitive cell death1 (sscd1) mutant in Arabidopsis. Map-based cloning of the corresponding gene revealed that SSCD1 encodes the Arabidopsis putative FAH. Further knockout of the gene encoding the Arabidopsis putative HGO completely eliminated the spontaneous cell death phenotype in the sscd1 mutant. Furthermore, we found that treatment of Arabidopsis wild-type seedlings with succinylacetone, an abnormal metabolite caused by loss of FAH in the Tyr degradation pathway (Lindblad et al., 1977), is able to mimic the sscd1 cell death phenotype. These results demonstrate that disruption of FAH leads to cell death in Arabidopsis and suggest that the Tyr degradation pathway is essential for plant survival under SD.  相似文献   

10.
Cytosolic Ca2+ in guard cells plays an important role in stomatal movement responses to environmental stimuli. These cytosolic Ca2+ increases result from Ca2+ influx through Ca2+-permeable channels in the plasma membrane and Ca2+ release from intracellular organelles in guard cells. However, the genes encoding defined plasma membrane Ca2+-permeable channel activity remain unknown in guard cells and, with some exceptions, largely unknown in higher plant cells. Here, we report the identification of two Arabidopsis (Arabidopsis thaliana) cation channel genes, CNGC5 and CNGC6, that are highly expressed in guard cells. Cytosolic application of cyclic GMP (cGMP) and extracellularly applied membrane-permeable 8-Bromoguanosine 3′,5′-cyclic monophosphate-cGMP both activated hyperpolarization-induced inward-conducting currents in wild-type guard cells using Mg2+ as the main charge carrier. The cGMP-activated currents were strongly blocked by lanthanum and gadolinium and also conducted Ba2+, Ca2+, and Na+ ions. cngc5 cngc6 double mutant guard cells exhibited dramatically impaired cGMP-activated currents. In contrast, mutations in CNGC1, CNGC2, and CNGC20 did not disrupt these cGMP-activated currents. The yellow fluorescent protein-CNGC5 and yellow fluorescent protein-CNGC6 proteins localize in the cell periphery. Cyclic AMP activated modest inward currents in both wild-type and cngc5cngc6 mutant guard cells. Moreover, cngc5 cngc6 double mutant guard cells exhibited functional abscisic acid (ABA)-activated hyperpolarization-dependent Ca2+-permeable cation channel currents, intact ABA-induced stomatal closing responses, and whole-plant stomatal conductance responses to darkness and changes in CO2 concentration. Furthermore, cGMP-activated currents remained intact in the growth controlled by abscisic acid2 and abscisic acid insensitive1 mutants. This research demonstrates that the CNGC5 and CNGC6 genes encode unique cGMP-activated nonselective Ca2+-permeable cation channels in the plasma membrane of Arabidopsis guard cells.Plants lose water via transpiration and take in CO2 for photosynthesis through stomatal pores. Each stomatal pore is surrounded by two guard cells, and stomatal movements are driven by the change of turgor pressure in guard cells. The intracellular second messenger Ca2+ functions in guard cell signal transduction (Schroeder and Hagiwara, 1989; McAinsh et al., 1990; Webb et al., 1996; Grabov and Blatt, 1998; Allen et al., 1999; MacRobbie, 2000; Mori et al., 2006; Young et al., 2006; Siegel et al., 2009; Chen et al., 2010; Hubbard et al., 2012). Plasma membrane ion channel activity and gene expression in guard cells are finely regulated by the intracellular free calcium concentration ([Ca2+]cyt; Schroeder and Hagiwara, 1989; Webb et al., 2001; Allen et al., 2002; Siegel et al., 2009; Kim et al., 2010; Stange et al., 2010). Ca2+-dependent protein kinases (CPKs) function as targets of the cytosolic Ca2+ signal, and several members of the CPK family have been shown to function in stimulus-induced stomatal closing, including the Arabidopsis (Arabidopsis thaliana) CPK3, CPK4, CPK6, CPK10, and CPK11 proteins (Mori et al., 2006; Zhu et al., 2007; Zou et al., 2010; Brandt et al., 2012; Hubbard et al., 2012). Further research found that several CPKs could activate the S-type anion channel SLAC1 in Xenopus laevis oocytes, including CPK21, CPK23, and CPK6 (Geiger et al., 2010; Brandt et al., 2012). At the same time, the Ca2+-independent protein kinase Open Stomata1 mediates stomatal closing and activates the S-type anion channel SLAC1 (Mustilli et al., 2002; Yoshida et al., 2002; Geiger et al., 2009; Lee et al., 2009; Xue et al., 2011), indicating that both Ca2+-dependent and Ca2+-independent pathways function in guard cells.Multiple essential factors of guard cell abscisic acid (ABA) signal transduction function in the regulation of Ca2+-permeable channels and [Ca2+]cyt elevations, including Abscisic Acid Insensitive1 (ABI1), ABI2, Enhanced Response to Abscisic Acid1 (ERA1), the NADPH oxidases AtrbohD and AtrbohF, the Guard Cell Hydrogen Peroxide-Resistant1 (GHR1) receptor kinase, as well as the Ca2+-activated CPK6 protein kinase (Pei et al., 1998; Allen et al., 1999, 2002; Kwak et al., 2003; Miao et al., 2006; Mori et al., 2006; Hua et al., 2012). [Ca2+]cyt increases result from both Ca2+ release from intracellular Ca2+ stores (McAinsh et al., 1992) and Ca2+ influx across the plasma membrane (Hamilton et al., 2000; Pei et al., 2000; Murata et al., 2001; Kwak et al., 2003; Hua et al., 2012). Electrophysiological analyses have characterized nonselective Ca2+-permeable channel activity in the plasma membrane of guard cells (Schroeder and Hagiwara, 1990; Hamilton et al., 2000; Pei et al., 2000; Murata et al., 2001; Köhler and Blatt, 2002; Miao et al., 2006; Mori et al., 2006; Suh et al., 2007; Vahisalu et al., 2008; Hua et al., 2012). However, the genetic identities of Ca2+-permeable channels in the plasma membrane of guard cells have remained unknown despite over two decades of research on these channel activities.The Arabidopsis genome includes 20 genes encoding cyclic nucleotide-gated channel (CNGC) homologs and 20 genes encoding homologs to animal Glu receptor channels (Lacombe et al., 2001; Kaplan et al., 2007; Ward et al., 2009), which have been proposed to function in plant cells as cation channels (Schuurink et al., 1998; Arazi et al., 1999; Köhler et al., 1999). Recent research has demonstrated functions of specific Glu receptor channels in mediating Ca2+ channel activity (Michard et al., 2011; Vincill et al., 2012). Previous studies have shown cAMP activation of nonselective cation currents in guard cells (Lemtiri-Chlieh and Berkowitz, 2004; Ali et al., 2007). However, only a few studies have shown the disappearance of a defined plasma membrane Ca2+ channel activity in plants upon mutation of candidate Ca2+ channel genes (Ali et al., 2007; Michard et al., 2011; Laohavisit et al., 2012; Vincill et al., 2012). Some CNGCs have been found to be involved in cation nutrient intake, including monovalent cation intake (Guo et al., 2010; Caballero et al., 2012), salt tolerance (Guo et al., 2008; Kugler et al., 2009), programmed cell death and pathogen responses (Clough et al., 2000; Balagué et al., 2003; Urquhart et al., 2007; Abdel-Hamid et al., 2013), thermal sensing (Finka et al., 2012; Gao et al., 2012), and pollen tube growth (Chang et al., 2007; Frietsch et al., 2007; Tunc-Ozdemir et al., 2013a, 2013b). Direct in vivo disappearance of Ca2+ channel activity in cngc disruption mutants has been demonstrated in only a few cases thus far (Ali et al., 2007; Gao et al., 2012). In this research, we show that CNGC5 and CNGC6 are required for a cyclic GMP (cGMP)-activated nonselective Ca2+-permeable cation channel activity in the plasma membrane of Arabidopsis guard cells.  相似文献   

11.
In rice (Oryza sativa) roots, lysigenous aerenchyma, which is created by programmed cell death and lysis of cortical cells, is constitutively formed under aerobic conditions, and its formation is further induced under oxygen-deficient conditions. Ethylene is involved in the induction of aerenchyma formation. reduced culm number1 (rcn1) is a rice mutant in which the gene encoding the ATP-binding cassette transporter RCN1/OsABCG5 is defective. Here, we report that the induction of aerenchyma formation was reduced in roots of rcn1 grown in stagnant deoxygenated nutrient solution (i.e. under stagnant conditions, which mimic oxygen-deficient conditions in waterlogged soils). 1-Aminocyclopropane-1-carboxylic acid synthase (ACS) is a key enzyme in ethylene biosynthesis. Stagnant conditions hardly induced the expression of ACS1 in rcn1 roots, resulting in low ethylene production in the roots. Accumulation of saturated very-long-chain fatty acids (VLCFAs) of 24, 26, and 28 carbons was reduced in rcn1 roots. Exogenously supplied VLCFA (26 carbons) increased the expression level of ACS1 and induced aerenchyma formation in rcn1 roots. Moreover, in rice lines in which the gene encoding a fatty acid elongase, CUT1-LIKE (CUT1L; a homolog of the gene encoding Arabidopsis CUT1, which is required for cuticular wax production), was silenced, both ACS1 expression and aerenchyma formation were reduced. Interestingly, the expression of ACS1, CUT1L, and RCN1/OsABCG5 was induced predominantly in the outer part of roots under stagnant conditions. These results suggest that, in rice under oxygen-deficient conditions, VLCFAs increase ethylene production by promoting 1-aminocyclopropane-1-carboxylic acid biosynthesis in the outer part of roots, which, in turn, induces aerenchyma formation in the root cortex.Aerenchyma formation is a morphological adaptation of plants to complete submergence and waterlogging of the soil, and facilitates internal gas diffusion (Armstrong, 1979; Jackson and Armstrong, 1999; Colmer, 2003; Voesenek et al., 2006; Bailey-Serres and Voesenek, 2008; Licausi and Perata, 2009; Sauter, 2013; Voesenek and Bailey-Serres, 2015). To adapt to waterlogging in soil, rice (Oryza sativa) develops lysigenous aerenchyma in shoots (Matsukura et al., 2000; Colmer and Pedersen, 2008; Steffens et al., 2011) and roots (Jackson et al., 1985b; Justin and Armstrong, 1991; Kawai et al., 1998), which is formed by programmed cell death and subsequent lysis of some cortical cells (Jackson and Armstrong, 1999; Evans, 2004; Yamauchi et al., 2013). In rice roots, lysigenous aerenchyma is constitutively formed under aerobic conditions (Jackson et al., 1985b), and its formation is further induced under oxygen-deficient conditions (Colmer et al., 2006; Shiono et al., 2011). The former and latter are designated constitutive and inducible lysigenous aerenchyma formation, respectively (Colmer and Voesenek, 2009). The gaseous plant hormone ethylene regulates adaptive growth responses of plants to submergence (Voesenek and Blom, 1989; Voesenek et al., 1993; Visser et al., 1996a,b; Lorbiecke and Sauter, 1999; Hattori et al., 2009; Steffens and Sauter, 2009; van Veen et al., 2013). Ethylene also induces lysigenous aerenchyma formation in roots of some gramineous plants (Drew et al., 2000; Shiono et al., 2008). The treatment of roots with ethylene or its precursor (1-aminocyclopropane-1-carboxylic acid [ACC]) stimulates aerenchyma formation in rice (Justin and Armstrong, 1991; Colmer et al., 2006; Yukiyoshi and Karahara, 2014), maize (Zea mays; Drew et al., 1981; Jackson et al., 1985a; Takahashi et al., 2015), and wheat (Triticum aestivum; Yamauchi et al., 2014a,b). Moreover, treatment of roots with inhibitors of ethylene action or ethylene biosynthesis effectively blocks aerenchyma formation under hypoxic conditions in maize (Drew et al., 1981; Konings, 1982; Jackson et al., 1985a; Rajhi et al., 2011).Ethylene biosynthesis is accomplished by two main successive enzymatic reactions: conversion of S-adenosyl-Met to ACC by 1-aminocyclopropane-1-carboxylic acid synthase (ACS), and conversion of ACC to ethylene by 1-aminocyclopropane-1-carboxylic acid oxidase (ACO; Yang and Hoffman, 1984). The activities of both enzymes are enhanced during aerenchyma formation under hypoxic conditions in maize root (He et al., 1996). Since the ACC content in roots of maize is increased by oxygen deficiency and is strongly correlated with ethylene production (Atwell et al., 1988), ACC biosynthesis is essential for ethylene production during aerenchyma formation in roots. In fact, exogenously supplied ACC induced ethylene production in roots of maize (Drew et al., 1979; Konings, 1982; Atwell et al., 1988) and wheat (Yamauchi et al., 2014b), even under aerobic conditions. Ethylene production in plants is inversely related to oxygen concentration (Yang and Hoffman, 1984). Under anoxic conditions, the oxidation of ACC to ethylene by ACO, which requires oxygen, is almost completely repressed (Yip et al., 1988; Tonutti and Ramina, 1991). Indeed, anoxic conditions stimulate neither ethylene production nor aerenchyma formation in maize adventitious roots (Drew et al., 1979). Therefore, it is unlikely that the root tissues forming inducible aerenchyma are anoxic, and that the ACO-mediated step is repressed. Moreover, aerenchyma is constitutively formed in rice roots even under aerobic conditions (Jackson et al., 1985b), and thus, after the onset of waterlogging, oxygen can be immediately supplied to the apical regions of roots through the constitutively formed aerenchyma.Very-long-chain fatty acids (VLCFAs; ≥20 carbons) are major constituents of sphingolipids, cuticular waxes, and suberin in plants (Franke and Schreiber, 2007; Kunst and Samuels, 2009). In addition to their structural functions, VLCFAs directly or indirectly participate in several physiological processes (Zheng et al., 2005; Reina-Pinto et al., 2009; Roudier et al., 2010; Ito et al., 2011; Nobusawa et al., 2013; Tsuda et al., 2013), including the regulation of ethylene biosynthesis (Qin et al., 2007). During fiber cell elongation in cotton ovules, ethylene biosynthesis is enhanced by treatment with saturated VLCFAs, especially 24-carbon fatty acids, and is suppressed by an inhibitor of VLCFA biosynthesis (Qin et al., 2007). The first rate-limiting step in VLCFA biosynthesis is condensation of acyl-CoA with malonyl-CoA by β-ketoacyl-CoA synthase (KCS; Joubès et al., 2008). KCS enzymes are thought to determine the substrate and tissue specificities of fatty acid elongation (Joubès et al., 2008). The Arabidopsis (Arabidopsis thaliana) genome has 21 KCS genes (Joubès et al., 2008). In the Arabidopsis cut1 mutant, which has a defect in the gene encoding CUT1 that is required for cuticular wax production (i.e. one of the KCS genes), the expression of AtACO genes and growth of root cells were reduced when compared with the wild type (Qin et al., 2007). Furthermore, expression of the AtACO genes was rescued by exogenously supplied saturated VLCFAs (Qin et al., 2007). These observations imply that VLCFAs or their derivatives work as regulatory factors for gene expression during some physiological processes in plants.reduced culm number1 (rcn1) was first identified as a rice mutant with a low tillering rate in a paddy field (Takamure and Kinoshita, 1985; Yasuno et al., 2007). The rcn1 (rcn1-2) mutant has a single nucleotide substitution in the gene encoding a member of the ATP-binding cassette (ABC) transporter subfamily G, RCN1/OsABCG5, causing an Ala-684Pro substitution (Yasuno et al., 2009). The mutation results in several mutant phenotypes, although the substrates of RCN1/OsABCG5 have not been determined (Ureshi et al., 2012; Funabiki et al., 2013; Matsuda et al., 2014). We previously found that the rcn1 mutant has abnormal root morphology, such as shorter root length and brownish appearance of roots, under stagnant (deoxygenated) conditions (which mimics oxygen-deficient conditions in waterlogged soils). We also found that the rcn1 mutant accumulates less of the major suberin monomers originating from VLCFAs in the outer part of adventitious roots, and this results in a reduction of a functional apoplastic barrier in the root hypodermis (Shiono et al., 2014a).The objective of this study was to elucidate the molecular basis of inducible aerenchyma formation. To this end, we examined lysigenous aerenchyma formation and ACC, ethylene, and VLCFA accumulation and their biosyntheses in rcn1 roots. Based on the results of these studies, we propose that VLCFAs are involved in inducible aerenchyma formation through the enhancement of ethylene biosynthesis in rice roots.  相似文献   

12.
Plant metabolism is characterized by a unique complexity on the cellular, tissue, and organ levels. On a whole-plant scale, changing source and sink relations accompanying plant development add another level of complexity to metabolism. With the aim of achieving a spatiotemporal resolution of source-sink interactions in crop plant metabolism, a multiscale metabolic modeling (MMM) approach was applied that integrates static organ-specific models with a whole-plant dynamic model. Allowing for a dynamic flux balance analysis on a whole-plant scale, the MMM approach was used to decipher the metabolic behavior of source and sink organs during the generative phase of the barley (Hordeum vulgare) plant. It reveals a sink-to-source shift of the barley stem caused by the senescence-related decrease in leaf source capacity, which is not sufficient to meet the nutrient requirements of sink organs such as the growing seed. The MMM platform represents a novel approach for the in silico analysis of metabolism on a whole-plant level, allowing for a systemic, spatiotemporally resolved understanding of metabolic processes involved in carbon partitioning, thus providing a novel tool for studying yield stability and crop improvement.Plants are of vital significance as a source of food (Grusak and DellaPenna, 1999; Rogalski and Carrer, 2011), feed (Lu et al., 2011), energy (Tilman et al., 2006; Parmar et al., 2011), and feedstocks for the chemical industry (Metzger and Bornscheuer, 2006; Kinghorn et al., 2011). Given the close connection between plant metabolism and the usability of plant products, there is a growing interest in understanding and predicting the behavior and regulation of plant metabolic processes. In order to increase crop quality and yield, there is a need for methods guiding the rational redesign of the plant metabolic network (Schwender, 2009).Mathematical modeling of plant metabolism offers new approaches to understand, predict, and modify complex plant metabolic processes. In plant research, the issue of metabolic modeling is constantly gaining attention, and different modeling approaches applied to plant metabolism exist, ranging from highly detailed quantitative to less complex qualitative approaches (for review, see Giersch, 2000; Morgan and Rhodes, 2002; Poolman et al., 2004; Rios-Estepa and Lange, 2007).A widely used modeling approach is flux balance analysis (FBA), which allows the prediction of metabolic capabilities and steady-state fluxes under different environmental and genetic backgrounds using (non)linear optimization (Orth et al., 2010). Assuming steady-state conditions, FBA has the advantage of not requiring the knowledge of kinetic parameters and, therefore, can be applied to model detailed, large-scale systems. In recent years, the FBA approach has been applied to several different plant species, such as maize (Zea mays; Dal’Molin et al., 2010; Saha et al., 2011), barley (Hordeum vulgare; Grafahrend-Belau et al., 2009b; Melkus et al., 2011; Rolletschek et al., 2011), rice (Oryza sativa; Lakshmanan et al., 2013), Arabidopsis (Arabidopsis thaliana; Poolman et al., 2009; de Oliveira Dal’Molin et al., 2010; Radrich et al., 2010; Williams et al., 2010; Mintz-Oron et al., 2012; Cheung et al., 2013), and rapeseed (Brassica napus; Hay and Schwender, 2011a, 2011b; Pilalis et al., 2011), as well as algae (Boyle and Morgan, 2009; Cogne et al., 2011; Dal’Molin et al., 2011) and photoautotrophic bacteria (Knoop et al., 2010; Montagud et al., 2010; Boyle and Morgan, 2011). These models have been used to study different aspects of metabolism, including the prediction of optimal metabolic yields and energy efficiencies (Dal’Molin et al., 2010; Boyle and Morgan, 2011), changes in flux under different environmental and genetic backgrounds (Grafahrend-Belau et al., 2009b; Dal’Molin et al., 2010; Melkus et al., 2011), and nonintuitive metabolic pathways that merit subsequent experimental investigations (Poolman et al., 2009; Knoop et al., 2010; Rolletschek et al., 2011). Although FBA of plant metabolic models was shown to be capable of reproducing experimentally determined flux distributions (Williams et al., 2010; Hay and Schwender, 2011b) and generating new insights into metabolic behavior, capacities, and efficiencies (Sweetlove and Ratcliffe, 2011), challenges remain to advance the utility and predictive power of the models.Given that many plant metabolic functions are based on interactions between different subcellular compartments, cell types, tissues, and organs, the reconstruction of organ-specific models and the integration of these models into interacting multiorgan and/or whole-plant models is a prerequisite to get insight into complex plant metabolic processes organized on a whole-plant scale (e.g. source-sink interactions). Almost all FBA models of plant metabolism are restricted to one cell type (Boyle and Morgan, 2009; Knoop et al., 2010; Montagud et al., 2010; Cogne et al., 2011; Dal’Molin et al., 2011), one tissue or one organ (Grafahrend-Belau et al., 2009b; Hay and Schwender, 2011a, 2011b; Pilalis et al., 2011; Mintz-Oron et al., 2012), and only one model exists taking into account the interaction between two cell types by specifying the interaction between mesophyll and bundle sheath cells in C4 photosynthesis (Dal’Molin et al., 2010). So far, no model representing metabolism at the whole-plant scale exists.Considering whole-plant metabolism raises the problem of taking into account temporal and environmental changes in metabolism during plant development and growth. Although classical static FBA is unable to predict the dynamics of metabolic processes, as the network analysis is based on steady-state solutions, time-dependent processes can be taken into account by extending the classical static FBA to a dynamic flux balance analysis (dFBA), as proposed by Mahadevan et al. (2002). The static (SOA) and dynamic optimization approaches introduced in this work provide a framework for analyzing the transience of metabolism by integrating kinetic expressions to dynamically constrain exchange fluxes. Due to the requirement of knowing or estimating a large number of kinetic parameters, so far dFBA has only been applied to a plant metabolic model once, to study the photosynthetic metabolism in the chloroplasts of C3 plants by a simplified model of five biochemical reactions (Luo et al., 2009). Integrating a dynamic model into a static FBA model is an alternative approach to perform dFBA.In this study, a multiscale metabolic modeling (MMM) approach was applied with the aim of achieving a spatiotemporal resolution of cereal crop plant metabolism. To provide a framework for the in silico analysis of the metabolic dynamics of barley on a whole-plant scale, the MMM approach integrates a static multiorgan FBA model and a dynamic whole-plant multiscale functional plant model (FPM) to perform dFBA. The performance of the novel whole-plant MMM approach was tested by studying source-sink interactions during the seed developmental phase of barley plants.  相似文献   

13.
14.
15.
16.
Terpenoid volatiles are isoprene compounds that are emitted by plants to communicate with the environment. In addition to their function in repelling herbivores and attracting carnivorous predators in green tissues, the presumed primary function of terpenoid volatiles released from mature fruits is the attraction of seed-dispersing animals. Mature oranges (Citrus sinensis) primarily accumulate terpenes in peel oil glands, with d-limonene accounting for approximately 97% of the total volatile terpenes. In a previous report, we showed that down-regulation of a d-limonene synthase gene alters monoterpene levels in orange antisense (AS) fruits, leading to resistance against Penicillium digitatum infection. A global gene expression analysis of AS versus empty vector (EV) transgenic fruits revealed that the down-regulation of d-limonene up-regulated genes involved in the innate immune response. Basal levels of jasmonic acid were substantially higher in the EV compared with AS oranges. Upon fungal challenge, salicylic acid levels were triggered in EV samples, while jasmonic acid metabolism and signaling were drastically increased in AS orange peels. In nature, d-limonene levels increase in orange fruit once the seeds are fully viable. The inverse correlation between the increase in d-limonene content and the decrease in the defense response suggests that d-limonene promotes infection by microorganisms that are likely involved in facilitating access to the pulp for seed-dispersing frugivores.Plants are sessile organisms that produce and emit a vast array of volatile organic compounds (VOCs) to communicate between parts of the same plant and with other plants. It is generally accepted that the original role of these compounds in nature is related to defense functions (Degenhardt et al., 2003). Most VOCs are terpenoids, fatty acid degradation compounds, phenylpropanoids, and amino acid-derived products. Among these, terpenoids are likely to be the most abundant and expensive to produce (Gershenzon, 1994). Terpenoids are isoprenoid-derived compounds synthesized through the condensation of C5 isoprene units, a process that is catalyzed by a wide diversity of terpene synthases using geranyl diphosphate (GDP), farnesyl diphosphate (FDP), and geranylgeranyl diphosphate (GGDP) as substrates. These reactions give rise to the C5 hemiterpenes, the C10 monoterpenes, the C15 sesquiterpenes, and the C20 diterpenes (Dudareva et al., 2006).In green tissues, volatile terpenoid synthesis is either induced upon wounding or occurs constitutively; terpenes can be then stored in specific organs or tissues where they would be most effective in defense responses, such as leaf trichomes, resin ducts and lacticifers, pockets near the epidermis, or secretory cavities in Citrus spp. (Langenheim, 1994; Turner et al., 2000; Trapp and Croteau, 2001; Voo et al., 2012). Genetic engineering experiments have demonstrated that specific terpenoid compounds emitted by leaves can intoxicate, repel, or deter herbivores (Aharoni et al., 2003; Wu et al., 2006), or they may attract the natural predators and parasitoids of damaging herbivores to protect plants from further damage (Kappers et al., 2005; Schnee et al., 2006). These terpenoids are naturally found in complex mixtures, and it has been proposed that they can act synergistically, as in conifer resin, for simultaneous protection against pests and pathogens (Phillips and Croteau, 1999). Although fatty acid degradation products (such as jasmonates) and phenylpropanoids (such as salicylates) as well as their volatile and nonvolatile precursors are clearly involved in many induced defense responses against pests and pathogens (Glazebrook, 2005), much less is known regarding the participation of terpenoid volatiles in the defense against microorganisms in plants and about the possible interactions of these terpenoids with phytohormones.In contrast to their function in leaves, when released from flowers and mature fruits, the main function of terpenoid volatiles is in the attraction of pollinators (Pichersky and Gershenzon, 2002; Kessler et al., 2008; Junker and Blüthgen, 2010; Schiestl, 2010) and seed-dispersing animals (Lomáscolo et al., 2010; Rodríguez et al., 2011b), respectively. Fruit maturation and ripening are usually associated with large increases in the synthesis and accumulation of specific flavored volatiles, which are proposed to function as signals for seed dispersal (Auldridge et al., 2006; Goff and Klee, 2006; Rodríguez et al., 2013).Upon wounding, plant responses to biotic stresses are orchestrated locally and systemically by signaling molecules. Among these molecules, the jasmonates regulate defenses against arthropod herbivores and necrotroph fungal pathogens as well as biotrophic pathogens, such as some mildews (Ellis and Turner, 2001; Stintzi et al., 2001; Kessler et al., 2004; Li et al., 2005; Wasternack, 2007; Browse and Howe, 2008). In addition to jasmonates, molecules such as salicylic acid (SA) and ethylene appear to regulate distinct defense pathways and are major synergistic (Mur et al., 2006) or antagonistic (De Vos et al., 2005) regulators of plant innate immunity. Plants produce a specific blend of these alarm signals after pathogen or pest attacks, and the production of these molecules varies greatly in quantity, composition, and timing. These signals activate differential sets of defense-related genes that eventually determine the nature of the defense response against the attacker (Reymond and Farmer, 1998; Rojo et al., 2003; De Vos et al., 2005). All genes that encode enzymes involved in the biosynthesis of jasmonates are jasmonic acid (JA) inducible (Wasternack, 2006), indicating that JA biosynthesis is regulated by positive feedback. The precursor for the biosynthesis of JA is α-linolenic acid. The activity of the 13-lipoxygenase (LOX), allene oxide synthase (AOS), and allene oxide cyclase (AOC) enzymes converts α-linolenic acid to cis-(+)-12-oxophytodienoic acid (OPDA). OPDA REDUCTASE3 catalyzes the reduction of OPDA (and dinor-OPDA) to oxo-pentenyl-cycloheptane-octanoic acid, which, in turn, undergoes three rounds of β-oxidation leading to jasmonyl-CoA formation. Jasmonyl-CoA is then cleaved by a putative thioesterase yielding (+)-7-iso-JA, which equilibrates to the more stable (−)-JA (Wasternack and Kombrink, 2010).The exogenous application of jasmonates on plants and the existence of mutant and/or transgenic plants altered in JA biosynthesis or signaling have led to altered susceptibility or resistance to pathogens. Impaired JA biosynthesis or signaling is generally associated with decreased levels of defensive compounds, including VOCs, and reduced plant biomass and/or fitness under insect attack (Howe et al., 1996; Halitschke and Baldwin, 2004). For example, Arabidopsis (Arabidopsis thaliana) mutants defective in JA perception (e.g. coronatine-insensitive1 [coi1]) or biosynthesis (e.g. aos and defective in anther dehiscence1) are susceptible to pathogen infections (Feys et al., 1994; Xie et al., 1998; Park et al., 2002; Turner et al., 2002). In contrast, mutants (e.g. constitutive expression of vegetative storage protein1 and Arabidopsis Ser/Thr phosphatase of type 2C1) with constitutive or wound-induced activation of the JA pathway exhibit enhanced resistance to fungal pathogens and pests and phenotypes characteristic of JA-treated plants (Ellis and Turner, 2001; Ellis et al., 2002; Schweighofer et al., 2007).Sweet orange (Citrus sinensis) is a perennial tree species that is exposed to recurrent biotic and abiotic challenges during its decades of growth in orchards. Orange fruits undergo a nonclimacteric maturation process in which the biochemistry, physiology, and structure of the organ are altered to complete the release of mature seeds. These changes typically include fruit growth and texture modification; color change through the degradation of chlorophylls and a parallel induction of carotenogenesis in the peel (flavedo) and pulp; flavonoid accumulation in the pulp; increases and decreases in the sugar and acid contents, respectively; and global accumulation and selective emission of volatile terpenoids (Spiegel-Roy and Goldschmidt, 1996). In nature, d-limonene accumulates gradually in the oil glands of the peel during fruit development and reaches its maximum level shortly before the breaker stage, followed by a steady decline during maturation (Attaway et al., 1967; Kekelidze et al., 1989; Rodríguez et al., 2011b). The high amount of d-limonene that accumulates in orange peels has a tremendous metabolic cost, suggesting an important biological role for this terpene and other related compounds in the interactions between fruits and the biotic environment.Previously, we examined the biological role of d-limonene by manipulating oil gland chemistry via the antisense (AS) overexpression of a d-limonene synthase gene from Satsuma mandarin (Citrus unshiu) in orange fruits. Compared with empty vector (EV) controls, fruit peels from AS transformants showed a dramatic reduction in d-limonene accumulation; decreased levels of other monoterpenes, sesquiterpenes, and monoterpene aldehydes; and increased levels of monoterpene alcohols. When challenged with the necrotroph fungus Penicillium digitatum, the causal agent of green mold rot, AS-transformed fruits were highly resistant to fungal infection. Full susceptibility to P. digitatum infection was restored when AS fruits were supplemented with d-limonene but not other monoterpene alcohols, indicating that d-limonene accumulation in the orange peel was required for the successful progress of this plant-pathogen interaction (Rodríguez et al., 2011a, 2011b). Green mold rot is the most important postharvest disease of citrus fruit worldwide, accounting for up to 60% to 80% of total losses during postharvest life of the fruit. P. digitatum is considered to be a specialist pathogen of citrus fruits that efficiently infects the peel through injuries in which ubiquitous fungal spores germinate and rapidly colonize the surrounding areas (Droby et al., 2008). The control of this pathogen relies heavily on the use of synthetic chemicals, but concerns regarding their potential negative effects on human health and also the generation of fungicide-resistant strains have encouraged finding alternatives, such as the generation of citrus trees with fruits that are genetically resistant to the pathogen.In this work, to better understand the mechanism underlying the constitutive resistance to P. digitatum conferred by the reduction of limonene in AS orange fruits, we analyzed the pattern of fruit growth and the morphological and biochemical developmental characteristics and performed a global analysis of gene expression using a 20K citrus microarray. The study is supplemented by examining the possible involvement of key hormone signals and isoprenoid precursors in the fruit peel. We report here that the reduced level of d-limonene in AS fruits is tightly associated with the constitutive activation of defense response signaling cascades. Our results establish, to our knowledge for the first time, a correlation between increased volatile terpene content and the decline of defense responses in a fleshy fruit during maturation, which would facilitate necrotroph fungal infections in citrus fruits.  相似文献   

17.
18.
19.
20.
Organelle movement and positioning play important roles in fundamental cellular activities and adaptive responses to environmental stress in plants. To optimize photosynthetic light utilization, chloroplasts move toward weak blue light (the accumulation response) and escape from strong blue light (the avoidance response). Nuclei also move in response to strong blue light by utilizing the light-induced movement of attached plastids in leaf cells. Blue light receptor phototropins and several factors for chloroplast photorelocation movement have been identified through molecular genetic analysis of Arabidopsis (Arabidopsis thaliana). PLASTID MOVEMENT IMPAIRED1 (PMI1) is a plant-specific C2-domain protein that is required for efficient chloroplast photorelocation movement. There are two PLASTID MOVEMENT IMPAIRED1-RELATED (PMIR) genes, PMIR1 and PMIR2, in the Arabidopsis genome. However, the mechanism in which PMI1 regulates chloroplast and nuclear photorelocation movements and the involvement of PMIR1 and PMIR2 in these organelle movements remained unknown. Here, we analyzed chloroplast and nuclear photorelocation movements in mutant lines of PMI1, PMIR1, and PMIR2. In mesophyll cells, the pmi1 single mutant showed severe defects in both chloroplast and nuclear photorelocation movements resulting from the impaired regulation of chloroplast-actin filaments. In pavement cells, pmi1 mutant plants were partially defective in both plastid and nuclear photorelocation movements, but pmi1pmir1 and pmi1pmir1pmir2 mutant lines lacked the blue light-induced movement responses of plastids and nuclei completely. These results indicated that PMI1 is essential for chloroplast and nuclear photorelocation movements in mesophyll cells and that both PMI1 and PMIR1 are indispensable for photorelocation movements of plastids and thus, nuclei in pavement cells.In plants, organelles move within the cell and become appropriately positioned to accomplish their functions and adapt to the environment (for review, see Wada and Suetsugu, 2004). Light-induced chloroplast movement (chloroplast photorelocation movement) is one of the best characterized organelle movements in plants (Suetsugu and Wada, 2012). Under weak light conditions, chloroplasts move toward light to capture light efficiently (the accumulation response; Zurzycki, 1955). Under strong light conditions, chloroplasts escape from light to avoid photodamage (the avoidance response; Kasahara et al., 2002; Sztatelman et al., 2010; Davis and Hangarter, 2012; Cazzaniga et al., 2013). In most green plant species, these responses are induced primarily by the blue light receptor phototropin (phot) in response to a range of wavelengths from UVA to blue light (approximately 320–500 nm; for review, see Suetsugu and Wada, 2012; Wada and Suetsugu, 2013; Kong and Wada, 2014). Phot-mediated chloroplast movement has been shown in land plants, such as Arabidopsis (Arabidopsis thaliana; Jarillo et al., 2001; Kagawa et al., 2001; Sakai et al., 2001), the fern Adiantum capillus-veneris (Kagawa et al., 2004), the moss Physcomitrella patens (Kasahara et al., 2004), and the liverwort Marchantia polymorpha (Komatsu et al., 2014). Two phots in Arabidopsis, phot1 and phot2, redundantly mediate the accumulation response (Sakai et al., 2001), whereas phot2 primarily regulates the avoidance response (Jarillo et al., 2001; Kagawa et al., 2001; Luesse et al., 2010). M. polymorpha has only one phot that mediates both the accumulation and avoidance responses (Komatsu et al., 2014), although two or more phots mediate chloroplast photorelocation movement in A. capillus-veneris (Kagawa et al., 2004) and P. patens (Kasahara et al., 2004). Thus, duplication and functional diversification of PHOT genes have occurred during land plant evolution, and plants have gained a sophisticated light sensing system for chloroplast photorelocation movement.In general, movements of plant organelles, including chloroplasts, are dependent on actin filaments (for review, see Wada and Suetsugu, 2004). Most organelles common in eukaryotes, such as mitochondria, peroxisomes, and Golgi bodies, use the myosin motor for their movements, but there is no clear evidence that chloroplast movement is myosin dependent (for review, see Suetsugu et al., 2010a). Land plants have innovated a novel actin-based motility system that is specialized for chloroplast movement as well as a photoreceptor system (for review, see Suetsugu et al., 2010a; Wada and Suetsugu, 2013; Kong and Wada, 2014). Chloroplast-actin (cp-actin) filaments, which were first found in Arabidopsis, are short actin filaments specifically localized around the chloroplast periphery at the interface between the chloroplast and the plasma membrane (Kadota et al., 2009). Strong blue light induces the rapid disappearance of cp-actin filaments and then, their subsequent reappearance preferentially at the front region of the moving chloroplasts. This asymmetric distribution of cp-actin filaments is essential for directional chloroplast movement (Kadota et al., 2009; Kong et al., 2013a). The greater the difference in the amount of cp-actin filaments between the front and rear regions of chloroplasts becomes, the faster the chloroplasts move, in which the magnitude of the difference is determined by fluence rate (Kagawa and Wada, 2004; Kadota et al., 2009; Kong et al., 2013a). Strong blue light-induced disappearance of cp-actin filaments is regulated in a phot2-dependent manner before the intensive polymerization of cp-actin filaments at the front region occurs (Kadota et al., 2009; Ichikawa et al., 2011; Kong et al., 2013a). This phot2-dependent response contributes to the greater difference in the amount of cp-actin filaments between the front and rear regions of chloroplasts. Similar behavior of cp-actin filaments has also been observed in A. capillus-veneris (Tsuboi and Wada, 2012) and P. patens (Yamashita et al., 2011).Like chloroplasts, nuclei also show light-mediated movement and positioning (nuclear photorelocation movement) in land plants (for review, see Higa et al., 2014b). In gametophytic cells of A. capillus-veneris, weak light induced the accumulation responses of both chloroplasts and nuclei, whereas strong light induced avoidance responses (Kagawa and Wada, 1993, 1995; Tsuboi et al., 2007). However, in mesophyll cells of Arabidopsis, strong blue light induced both chloroplast and nuclear avoidance responses, but weak blue light induced only the chloroplast accumulation response (Iwabuchi et al., 2007, 2010; Higa et al., 2014a). In Arabidopsis pavement cells, small numbers of tiny plastids were found and showed autofluorescence under the confocal laser-scanning microscopy (Iwabuchi et al., 2010; Higa et al., 2014a). Hereafter, the plastid in the pavement cells is called the pavement cell plastid. Strong blue light-induced avoidance responses of pavement cell plastids and nuclei were induced in a phot2-dependent manner, but the accumulation response was not detected for either organelle (Iwabuchi et al., 2007, 2010; Higa et al., 2014a). In both Arabidopsis and A. capillus-veneris, phots mediate nuclear photorelocation movement, and phot2 mediates the nuclear avoidance response (Iwabuchi et al., 2007, 2010; Tsuboi et al., 2007). The nuclear avoidance response is dependent on actin filaments in both mesophyll and pavement cells of Arabidopsis (Iwabuchi et al., 2010). Recently, it was shown that the nuclear avoidance response relies on cp-actin-dependent movement of pavement cell plastids, where nuclei are associated with pavement cell plastids of Arabidopsis (Higa et al., 2014a). In mesophyll cells, nuclear avoidance response is likely dependent on cp-actin filament-mediated chloroplast movement, because the mutants deficient in chloroplast movement were also defective in nuclear avoidance response (Higa et al., 2014a). Thus, phots mediate both chloroplast (and pavement cell plastid) and nuclear photorelocation movement by regulating cp-actin filaments.Molecular genetic analyses of Arabidopsis mutants deficient in chloroplast photorelocation movement have identified many molecular factors involved in signal transduction and/or motility systems as well as those involved in the photoreceptor system for chloroplast photorelocation movement (and thus, nuclear photorelocation movement; for review, see Suetsugu and Wada, 2012; Wada and Suetsugu, 2013; Kong and Wada, 2014). CHLOROPLAST UNUSUAL POSITIONING1 (CHUP1; Oikawa et al., 2003) and KINESIN-LIKE PROTEIN FOR ACTIN-BASED CHLOROPLAST MOVEMENT (KAC; Suetsugu et al., 2010b) are key factors for generating and/or maintaining cp-actin filaments. Both proteins are highly conserved in land plants and essential for the movement and attachment of chloroplasts to the plasma membrane in Arabidopsis (Oikawa et al., 2003, 2008; Suetsugu et al., 2010b), A. capillus-veneris (Suetsugu et al., 2012), and P. patens (Suetsugu et al., 2012; Usami et al., 2012). CHUP1 is localized on the chloroplast outer membrane and binds to globular and filamentous actins and profilin in vitro (Oikawa et al., 2003, 2008; Schmidt von Braun and Schleiff, 2008). Although KAC is a kinesin-like protein, it lacks microtubule-dependent motor activity but has filamentous actin binding activity (Suetsugu et al., 2010b). An actin-bundling protein THRUMIN1 (THRUM1) is required for efficient chloroplast photorelocation movement (Whippo et al., 2011) and interacts with cp-actin filaments (Kong et al., 2013a). chup1 and kac mutant plants were shown to lack detectable cp-actin filaments (Kadota et al., 2009; Suetsugu et al., 2010b; Ichikawa et al., 2011; Kong et al., 2013a). Similarly, cp-actin filaments were rarely detected in thrum1 mutant plants (Kong et al., 2013a), indicating that THRUM1 also plays an important role in maintaining cp-actin filaments.Other proteins J-DOMAIN PROTEIN REQUIRED FOR CHLOROPLAST ACCUMULATION RESPONSE1 (JAC1; Suetsugu et al., 2005), WEAK CHLOROPLAST MOVEMENT UNDER BLUE LIGHT1 (WEB1; Kodama et al., 2010), and PLASTID MOVEMENT IMPAIRED2 (PMI2; Luesse et al., 2006; Kodama et al., 2010) are involved in the light regulation of cp-actin filaments and chloroplast photorelocation movement. JAC1 is an auxilin-like J-domain protein that mediates the chloroplast accumulation response through its J-domain function (Suetsugu et al., 2005; Takano et al., 2010). WEB1 and PMI2 are coiled-coil proteins that interact with each other (Kodama et al., 2010). Although web1 and pmi2 were partially defective in the avoidance response, the jac1 mutation completely suppressed the phenotype of web1 and pmi2, suggesting that the WEB1/PMI2 complex suppresses JAC1 function (i.e. the accumulation response) under strong light conditions (Kodama et al., 2010). Both web1 and pmi2 showed impaired disappearance of cp-actin filaments in response to strong blue light (Kodama et al., 2010). However, the exact molecular functions of these proteins are unknown.In this study, we characterized mutant plants deficient in the PMI1 gene and two homologous genes PLASTID MOVEMENT IMPAIRED1-RELATED1 (PMIR1) and PMIR2. PMI1 was identified through molecular genetic analyses of pmi1 mutants that showed severe defects in chloroplast accumulation and avoidance responses (DeBlasio et al., 2005). PMI1 is a plant-specific C2-domain protein (DeBlasio et al., 2005; Zhang and Aravind, 2010), but its roles and those of PMIRs in cp-actin-mediated chloroplast and nuclear photorelocation movements remained unclear. Thus, we analyzed chloroplast and nuclear photorelocation movements in the single, double, and triple mutants of pmi1, pmir1, and pmir2.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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