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 62 毫秒
1.
Here, we investigated the physiological role of Arabidopsis (Arabidopsis thaliana) AtNUDX6, the gene encoding ADP-ribose (Rib)/NADH pyrophosphohydrolase, using its overexpressor (Pro35S:AtNUDX6) or disruptant (KO-nudx6). The level of NADH in Pro35S:AtNUDX6 and KO-nudx6 plants was decreased and increased, respectively, compared with that of the control plants, while the level of ADP-Rib was not changed in either plant. The activity of pyrophosphohydrolase toward NADH was enhanced and reduced in the Pro35S:AtNUDX6 and KO-nudx6 plants, respectively. The decrease in the activity of NADH pyrophosphohydrolase and the increase in the level of NADH were observed in the rosette and cauline leaves, but not in the roots, of the KO-nudx6 plants. Notably, the expression level of AtNUDX6 and the activity of NADH pyrophosphohydrolase in the control plants, but not in the KO-nudx6 plants, were increased by the treatment with salicylic acid (SA). The expression of SA-induced genes (PR1, WRKY70, NIMIN1, and NIMIN2) depending on NONEXPRESSOR OF PATHOGENESIS-RELATED GENES1 (NPR1), a key component required for pathogen resistance, was significantly suppressed and enhanced in the KO-nudx6 and Pro35S:AtNUDX6 plants, respectively, under the treatment with SA. Induction of thioredoxin h5 (TRX-h5) expression, which catalyzes a SA-induced NPR1 activation, was suppressed and accelerated in the KO-nudx6 and Pro35S:AtNUDX6 plants, respectively. The expression of isochorismate synthase1, required for the regulation of SA synthesis through the NPR1-mediated feedback loop, was decreased and increased in the KO-nudx6 and Pro35S:AtNUDX6 plants, respectively. Judging from seed germination rates, the KO-nudx6 plants had enhanced sensitivity to the toxicity of high-level SA. These results indicated that AtNUDX6 is a modulator of NADH rather than ADP-Rib metabolism and that, through induction of TRX-h5 expression, AtNUDX6 significantly impacts the plant immune response as a positive regulator of NPR1-dependent SA signaling pathways.Nudix (nucleoside diphosphates linked to some moiety X) hydrolases are a phylogenetically widespread enzyme family and are widely distributed among all classes of organisms, such as bacteria, yeast, algae, nematodes, vertebrates, and plants (Bessman et al., 1996; Xu et al., 2004; Kraszewska, 2008). The enzymes catalyze, with varying degrees of substrate specificity, the hydrolysis of a variety of nucleoside diphosphate derivatives: nucleoside diphosphates and triphosphates and their oxidized forms, dinucleoside polyphosphates, nucleotide sugars, NADH, CoA, and the mRNA caps (McLennan, 2006; Kraszewska, 2008; Gunawardana et al., 2009). Since these compounds are often toxic to cells, Nudix hydrolases seem to play protective, regulatory, and signaling roles in metabolism by hydrolytically removing such compounds (Bessman et al., 1996; Xu et al., 2004).We reported the molecular and enzymatic characteristics of Nudix hydrolases (AtNUDX1–AtNUDX27) in Arabidopsis (Arabidopsis thaliana) plants (Ogawa et al., 2005, 2008). Notably, among 27 types of AtNUDXs, cytosolic AtNUDX2, AtNUDX6, AtNUDX7, and AtNUDX10 had pyrophosphohydrolase activity toward both ADP-Rib and NADH in vitro. Recent studies have shown that the actions of NADH and/or ADP-Rib pyrophosphohydrolases are closely related to defense systems in response to biotic and abiotic stresses in higher plants.It has been reported that the expression of AtNUDX7 is induced by avirulent pathogenic attacks. Knockout AtNUDX7 mutants (KO-nudx7) showed enhanced resistance against both virulent and avirulent bacterial strains (Bartsch et al., 2006; Jambunathan and Mahalingam, 2006; Adams-Phillips et al., 2008). In addition, it was revealed that AtNUDX7 functions as a negative regulator on ENHANCED DISEASE SUSCEPTIBILITY1 (EDS1) signaling required for basal resistance to invasive pathogens (Bartsch et al., 2006); EDS1 regulates accumulation of the phenolic defense molecule, salicylic acid (SA), and other as yet unidentified signal intermediates and controls the defense activation and programmed cell death by collaborating with its interaction partner PHYTOALEXIN-DEFICIENT4 in cells surrounding pathogen infection foci. Furthermore, Ge et al. (2007) reported that AtNUDX7 functions to prevent excessive stimulation of the defense response, which is dependent on and independent of NONEXPRESSOR OF PATHOGENESIS-RELATED GENES1 (NPR1), a master regulator of SA-induced defense genes (SAIGs), and SA accumulation.On the other hand, we recently demonstrated the roles of Arabidopsis NADH/ADP-Rib pyrophosphohydrolases (AtNUDX2 and AtNUDX7) in tolerance to oxidative stress using the respective overexpressors (Pro35S:AtNUDX2 and Pro35S:AtNUDX7) or disruptants (KO-nudx7; Ishikawa et al., 2009; Ogawa et al., 2009). Interestingly, overexpression of AtNUDX2 and AtNUDX7 in Arabidopsis plants was responsible for an enhanced tolerance to oxidative stress derived from the treatment with paraquat (an agent producing O2) and salinity. Taken together, these results revealed that both AtNUDX2 and AtNUDX7 function in accelerating nucleotide recycling from ADP-Rib produced by poly(ADP-Rib) metabolism, leading to suppression of the overconsumption of NAD+ and ATP in Arabidopsis cells under stressful conditions. In addition, AtNUDX7 served to balance between NADH and NAD+ by NADH turnover and to regulate the defense mechanisms against DNA damage by modulation of the poly(ADP-ribosyl)ation (PAR) reaction through NADH metabolism in response to oxidative stress (Ishikawa et al., 2009; Ogawa et al., 2009). These findings clearly indicated that the regulation of NADH and/or ADP-Rib metabolism via Nudix hydrolases is involved in the responses to both biotic and abiotic stresses in higher plants.The question that we must consider next is whether the other AtNUDXs (AtNUDX6 and AtNUDX10) with pyrophosphohydrolase activities toward ADP-Rib and NADH are involved in the defense systems against oxidative stress and pathogen attack. The expression of AtNUDX6 has been reported to be induced by pathogenic attacks and treatment with the SA analogs 2,6-dichloroisonicotinic acid and acibenzolar-S-methyl benzo(1,2,3)thiadiazole-7-carbothioic acid S-methyl ester (BTH; Bartsch et al., 2006; Qiu et al., 2008; Knoth et al., 2009). Furthermore, the expression of AtNUDX6 was strongly dependent on EDS1 (Bartsch et al., 2006). However, the functional significance of AtNUDX6 is still unclear, since a loss-of-function mutant of AtNUDX6 has not yet been found.In this paper, to assess the physiological function of AtNUDX6, we identified an Arabidopsis mutant in which T-DNA is inserted into AtNUDX6 and subsequently studied the levels of ADP-Rib and NAD(H), PAR activity, expression of genes related to SA signaling, and SA tolerance in the AtNUDX6 overexpressors and disruptants in comparison with the AtNUDX7 disruptants. The results obtained here indicated that AtNUDX6 positively regulates NPR1-dependent SA signaling via modulation of NADH metabolism in the plant immune response.  相似文献   

2.
To investigate sepal/petal/lip formation in Oncidium Gower Ramsey, three paleoAPETALA3 genes, O. Gower Ramsey MADS box gene5 (OMADS5; clade 1), OMADS3 (clade 2), and OMADS9 (clade 3), and one PISTILLATA gene, OMADS8, were characterized. The OMADS8 and OMADS3 mRNAs were expressed in all four floral organs as well as in vegetative leaves. The OMADS9 mRNA was only strongly detected in petals and lips. The mRNA for OMADS5 was only strongly detected in sepals and petals and was significantly down-regulated in lip-like petals and lip-like sepals of peloric mutant flowers. This result revealed a possible negative role for OMADS5 in regulating lip formation. Yeast two-hybrid analysis indicated that OMADS5 formed homodimers and heterodimers with OMADS3 and OMADS9. OMADS8 only formed heterodimers with OMADS3, whereas OMADS3 and OMADS9 formed homodimers and heterodimers with each other. We proposed that sepal/petal/lip formation needs the presence of OMADS3/8 and/or OMADS9. The determination of the final organ identity for the sepal/petal/lip likely depended on the presence or absence of OMADS5. The presence of OMADS5 caused short sepal/petal formation. When OMADS5 was absent, cells could proliferate, resulting in the possible formation of large lips and the conversion of the sepal/petal into lips in peloric mutants. Further analysis indicated that only ectopic expression of OMADS8 but not OMADS5/9 caused the conversion of the sepal into an expanded petal-like structure in transgenic Arabidopsis (Arabidopsis thaliana) plants.The ABCDE model predicts the formation of any flower organ by the interaction of five classes of homeotic genes in plants (Yanofsky et al., 1990; Jack et al., 1992; Mandel et al., 1992; Goto and Meyerowitz, 1994; Jofuku et al., 1994; Pelaz et al., 2000, 2001; Theißen and Saedler, 2001; Pinyopich et al., 2003; Ditta et al., 2004; Jack, 2004). The A class genes control sepal formation. The A, B, and E class genes work together to regulate petal formation. The B, C, and E class genes control stamen formation. The C and E class genes work to regulate carpel formation, whereas the D class gene is involved in ovule development. MADS box genes seem to have a central role in flower development, because most ABCDE genes encode MADS box proteins (Coen and Meyerowitz, 1991; Weigel and Meyerowitz, 1994; Purugganan et al., 1995; Rounsley et al., 1995; Theißen and Saedler, 1995; Theißen et al., 2000; Theißen, 2001).The function of B group genes, such as APETALA3 (AP3) and PISTILLATA (PI), has been thought to have a major role in specifying petal and stamen development (Jack et al., 1992; Goto and Meyerowitz, 1994; Krizek and Meyerowitz, 1996; Kramer et al., 1998; Hernandez-Hernandez et al., 2007; Kanno et al., 2007; Whipple et al., 2007; Irish, 2009). In Arabidopsis (Arabidopsis thaliana), mutation in AP3 or PI caused identical phenotypes of second whorl petal conversion into a sepal structure and third flower whorl stamen into a carpel structure (Bowman et al., 1989; Jack et al., 1992; Goto and Meyerowitz, 1994). Similar homeotic conversions for petal and stamen were observed in the mutants of the AP3 and PI orthologs from a number of core eudicots such as Antirrhinum majus, Petunia hybrida, Gerbera hybrida, Solanum lycopersicum, and Nicotiana benthamiana (Sommer et al., 1990; Tröbner et al., 1992; Angenent et al., 1993; van der Krol et al., 1993; Yu et al., 1999; Liu et al., 2004; Vandenbussche et al., 2004; de Martino et al., 2006), from basal eudicot species such as Papaver somniferum and Aquilegia vulgaris (Drea et al., 2007; Kramer et al., 2007), as well as from monocot species such as Zea mays and Oryza sativa (Ambrose et al., 2000; Nagasawa et al., 2003; Prasad and Vijayraghavan, 2003; Yadav et al., 2007; Yao et al., 2008). This indicated that the function of the B class genes AP3 and PI is highly conserved during evolution.It has been thought that B group genes may have arisen from an ancestral gene through multiple gene duplication events (Doyle, 1994; Theißen et al., 1996, 2000; Purugganan, 1997; Kramer et al., 1998; Kramer and Irish, 1999; Lamb and Irish, 2003; Kim et al., 2004; Stellari et al., 2004; Zahn et al., 2005; Hernandez-Hernandez et al., 2007). In the gymnosperms, there was a single putative B class lineage that duplicated to generate the paleoAP3 and PI lineages in angiosperms (Kramer et al., 1998; Theißen et al., 2000; Irish, 2009). The paleoAP3 lineage is composed of AP3 orthologs identified in lower eudicots, magnolid dicots, and monocots (Kramer et al., 1998). Genes in this lineage contain the conserved paleoAP3- and PI-derived motifs in the C-terminal end of the proteins, which have been thought to be characteristics of the B class ancestral gene (Kramer et al., 1998; Tzeng and Yang, 2001; Hsu and Yang, 2002). The PI lineage is composed of PI orthologs that contain a highly conserved PI motif identified in most plant species (Kramer et al., 1998). Subsequently, there was a second duplication at the base of the core eudicots that produced the euAP3 and TM6 lineages, which have been subject to substantial sequence changes in eudicots during evolution (Kramer et al., 1998; Kramer and Irish, 1999). The paleoAP3 motif in the C-terminal end of the proteins was retained in the TM6 lineage and replaced by a conserved euAP3 motif in the euAP3 lineage of most eudicot species (Kramer et al., 1998). In addition, many lineage-specific duplications for paleoAP3 lineage have occurred in plants such as orchids (Hsu and Yang, 2002; Tsai et al., 2004; Kim et al., 2007; Mondragón-Palomino and Theißen, 2008, 2009; Mondragón-Palomino et al., 2009), Ranunculaceae, and Ranunculales (Kramer et al., 2003; Di Stilio et al., 2005; Shan et al., 2006; Kramer, 2009).Unlike the A or C class MADS box proteins, which form homodimers that regulate flower development, the ability of B class proteins to form homodimers has only been reported in gymnosperms and in the paleoAP3 and PI lineages of some monocots. For example, LMADS1 of the lily Lilium longiflorum (Tzeng and Yang, 2001), OMADS3 of the orchid Oncidium Gower Ramsey (Hsu and Yang, 2002), and PeMADS4 of the orchid Phalaenopsis equestris (Tsai et al., 2004) in the paleoAP3 lineage, LRGLOA and LRGLOB of the lily Lilium regale (Winter et al., 2002), TGGLO of the tulip Tulipa gesneriana (Kanno et al., 2003), and PeMADS6 of the orchid P. equestris (Tsai et al., 2005) in the PI lineage, and GGM2 of the gymnosperm Gnetum gnemon (Winter et al., 1999) were able to form homodimers that regulate flower development. Proteins in the euAP3 lineage and in most paleoAP3 lineages were not able to form homodimers and had to interact with PI to form heterodimers in order to regulate petal and stamen development in various plant species (Schwarz-Sommer et al., 1992; Tröbner et al., 1992; Riechmann et al., 1996; Moon et al., 1999; Winter et al., 2002; Kanno et al., 2003; Vandenbussche et al., 2004; Yao et al., 2008). In addition to forming dimers, AP3 and PI were able to interact with other MADS box proteins, such as SEPALLATA1 (SEP1), SEP2, and SEP3, to regulate petal and stamen development (Pelaz et al., 2000; Honma and Goto, 2001; Theißen and Saedler, 2001; Castillejo et al., 2005).Orchids are among the most important plants in the flower market around the world, and research on MADS box genes has been reported for several species of orchids during the past few years (Lu et al., 1993, 2007; Yu and Goh, 2000; Hsu and Yang, 2002; Yu et al., 2002; Hsu et al., 2003; Tsai et al., 2004, 2008; Xu et al., 2006; Guo et al., 2007; Kim et al., 2007; Chang et al., 2009). Unlike the flowers in eudicots, the nearly identical shape of the sepals and petals as well as the production of a unique lip in orchid flowers make them a very special plant species for the study of flower development. Four clades (1–4) of genes in the paleoAP3 lineage have been identified in several orchids (Hsu and Yang, 2002; Tsai et al., 2004; Kim et al., 2007; Mondragón-Palomino and Theißen, 2008, 2009; Mondragón-Palomino et al., 2009). Several works have described the possible interactions among these four clades of paleoAP3 genes and one PI gene that are involved in regulating the differentiation and formation of the sepal/petal/lip of orchids (Tsai et al., 2004; Kim et al., 2007; Mondragón-Palomino and Theißen, 2008, 2009). However, the exact mechanism that involves the orchid B class genes remains unclear and needs to be clarified by more experimental investigations.O. Gower Ramsey is a popular orchid with important economic value in cut flower markets. Only a few studies have been reported on the role of MADS box genes in regulating flower formation in this plant species (Hsu and Yang, 2002; Hsu et al., 2003; Chang et al., 2009). An AP3-like MADS gene that regulates both floral formation and initiation in transgenic Arabidopsis has been reported (Hsu and Yang, 2002). In addition, four AP1/AGAMOUS-LIKE9 (AGL9)-like MADS box genes have been characterized that show novel expression patterns and cause different effects on floral transition and formation in Arabidopsis (Hsu et al., 2003; Chang et al., 2009). Compared with other orchids, the production of a large and well-expanded lip and five small identical sepals/petals makes O. Gower Ramsey a special case for the study of the diverse functions of B class MADS box genes during evolution. Therefore, the isolation of more B class MADS box genes and further study of their roles in the regulation of perianth (sepal/petal/lip) formation during O. Gower Ramsey flower development are necessary. In addition to the clade 2 paleoAP3 gene OMADS3, which was previously characterized in our laboratory (Hsu and Yang, 2002), three more B class MADS box genes, OMADS5, OMADS8, and OMADS9, were characterized from O. Gower Ramsey in this study. Based on the different expression patterns and the protein interactions among these four orchid B class genes, we propose that the presence of OMADS3/8 and/or OMADS9 is required for sepal/petal/lip formation. Further sepal and petal formation at least requires the additional presence of OMADS5, whereas large lip formation was seen when OMADS5 expression was absent. Our results provide a new finding and information pertaining to the roles for orchid B class MADS box genes in the regulation of sepal/petal/lip formation.  相似文献   

3.
4.
5.
The viral genome-linked protein, VPg, of potyviruses is a multifunctional protein involved in viral genome translation and replication. Previous studies have shown that both eukaryotic translation initiation factor 4E (eIF4E) and eIF4G or their respective isoforms from the eIF4F complex, which modulates the initiation of protein translation, selectively interact with VPg and are required for potyvirus infection. Here, we report the identification of two DEAD-box RNA helicase-like proteins, PpDDXL and AtRH8 from peach (Prunus persica) and Arabidopsis (Arabidopsis thaliana), respectively, both interacting with VPg. We show that AtRH8 is dispensable for plant growth and development but necessary for potyvirus infection. In potyvirus-infected Nicotiana benthamiana leaf tissues, AtRH8 colocalizes with the chloroplast-bound virus accumulation vesicles, suggesting a possible role of AtRH8 in viral genome translation and replication. Deletion analyses of AtRH8 have identified the VPg-binding region. Comparison of this region and the corresponding region of PpDDXL suggests that they are highly conserved and share the same secondary structure. Moreover, overexpression of the VPg-binding region from either AtRH8 or PpDDXL suppresses potyvirus accumulation in infected N. benthamiana leaf tissues. Taken together, these data demonstrate that AtRH8, interacting with VPg, is a host factor required for the potyvirus infection process and that both AtRH8 and PpDDXL may be manipulated for the development of genetic resistance against potyvirus infections.Plant viruses are obligate intracellular parasites that infect many agriculturally important crops and cause severe losses each year. One of the common characteristics of plant viruses is their relatively small genome that encodes a limited number of viral proteins, making them dependent on host factors to fulfill their infection cycles (Maule et al., 2002; Whitham and Wang, 2004; Nelson and Citovsky, 2005; Decroocq et al., 2006). In order to establish a successful infection, the invading virus must recruit an array of host proteins (host factors) to translate and replicate its genome and to move locally from cell to cell via the plasmodesmata and systemically via the vascular system. It has been suggested that down-regulation or mutation of some of the required host factors may result in recessively inherited resistance to viruses (Kang et al., 2005b).Potyviruses, belonging to the genus Potyvirus in the family Potyviradae, constitute the largest group of plant viruses (Rajamäki et al., 2004). Potyviruses have a single positive-strand RNA genome approximately 10 kb in length, with a viral genome-linked protein (VPg) covalently attached to the 5′ end and a poly(A) tail at the 3′ end (Urcuqui-Inchima et al., 2001; Rajamäki et al., 2004). The viral genome contains a single open reading frame (ORF) that translates into a polypeptide with a molecular mass of approximately 350 kD, which is cleaved into 10 mature proteins by viral proteases (Urcuqui-Inchima et al., 2001). Recently, a novel viral protein resulting from a frameshift in the P3 cistron has been reported (Chung et al., 2008). Of the 11 viral proteins, VPg is a multifunctional protein and the only other viral protein present in the viral particles (virions) besides the coat protein and the cylindrical inclusion protein (CI; Oruetxebarria et al., 2001; Puustinen et al., 2002; Gabrenaite-Verkhovskaya et al., 2008). The nonstructural protein is linked to the viral RNA by a phosphodiester bond between the 5′ terminal uridine residue of the RNA and the O4-hydroxyl group of amino acid Tyr (Murphy et al., 1996; Oruetxebarria et al., 2001; Puustinen et al., 2002). Mutation of the Tyr residue that links VPg to the viral RNA abolishes virus infectivity completely (Murphy et al., 1996). In infected cells, VPg and its precursor NIa are present in the nucleus and in the membrane-associated virus replication vesicles in the cytoplasm (Carrington et al., 1993; Rajamäki and Valkonen, 2003; Cotton et al., 2009). As a component of the replication complex, VPg may serve as a primer for viral RNA replication (Puustinen and Mäkinen, 2004) and as an analog of the m7G cap of mRNAs for the viral genome to recruit the translation complex for translation (Michon et al., 2006; Beauchemin et al., 2007; Khan et al., 2008). Furthermore, VPg has been suggested to be an avirulence factor for recessive resistance genes in diverse plant species (Moury et al., 2004; Kang et al., 2005b; Bruun-Rasmussen et al., 2007). Thus, VPg plays a pivotal role in the virus infection process. The molecular identification of VPg-interacting host proteins and the subsequent functional characterization of such interactions may advance knowledge of the intricate virus replication mechanisms and help develop novel antiviral strategies.Previous studies have shown that VPg and its precursor NIa interact with several host proteins, including three essential components of the host protein translation apparatus (Thivierge et al., 2008). The first protein is the cellular translation initiation factor eIF4E or its isoform eIF(iso)4E, identified through a yeast two-hybrid screen using VPg as a bait (Wittmann et al., 1997; Schaad et al., 2000). The protein complex of VPg and eIF4E is an essential component for virus infectivity (Robaglia and Caranta, 2006). Mutations and knockout of eIF4E or eIF(iso)4E confer resistance to infection (Lellis et al., 2002; Ruffel et al., 2002; Nicaise et al., 2003; Gao et al., 2004; Kang et al., 2005a; Ruffel et al., 2005; Decroocq et al., 2006; Bruun-Rasmussen et al., 2007). It is well known that potyviruses recruit selectively one of the eIF4E isoforms, depending on specific virus-host combinations (German-Retana et al., 2008). For instance, in Arabidopsis (Arabidopsis thaliana), eIF(iso)4E is required for infection by Turnip mosaic virus (TuMV), Plum pox virus (PPV), and Lettuce mosaic virus, while eIF4E is indispensable for infection by Clover yellow vein virus (Duprat et al., 2002; Lellis et al., 2002; Sato et al., 2005; Decroocq et al., 2006). The second cellular protein interacting with VPg is another translation initiation factor, eIF4G. Analysis of Arabidopsis knockout mutants for eIF4G or its isomers eIF(iso)4G1 and eIF(iso)4G2 has yielded results supporting the idea that the recruitment of eIF4G for potyvirus infection is also isoform dependent (Nicaise et al., 2007). Recently, poly(A)-binding protein (PABP), the translation initiation factor that bridges the 5′ and 3′ termini of the mRNA into proximity, has been proposed to be essential for efficient multiplication of TuMV (Dufresne et al., 2008). PABP was previously documented to interact with NIa, a VPg precursor containing both VPg and the proteinase NIa-Pro (Léonard et al., 2004). As the translation factors eIF(iso)4E and PABP have been found to be internalized in virus-induced vesicles, it has been suggested that the interactions between VPg and these translation factors are crucial for viral RNA translation and/or replication (Beauchemin and Laliberté, 2007; Beauchemin et al., 2007; Cotton et al., 2009). Besides these three translation factors, a Cys-rich plant protein, potyvirus VPg-interaction protein, was also found to associate with VPg (Dunoyer et al., 2004). This plant-specific VPg-interacting host protein contains a PHD finger domain and acts as an ancillary factor to support potyvirus infection and movement (Dunoyer et al., 2004).In this study, we describe the identification of an Arabidopsis DEAD-box RNA helicase (DDX), AtRH8, and a peach (Prunus persica) DDX-like protein, PpDDXL, both interacting with the potyviral VPg protein. Using the atrh8 mutant, we demonstrate that AtRH8 is not required for plant growth and development in Arabidopsis but is necessary for infection by two plant potyviruses, PPV and TuMV. Furthermore, we present evidence that AtRH8 colocalizes with the virus accumulation complex in potyvirus-infected leaf tissues, which reveals a possible role of AtRH8 in virus infection. Finally, we have identified the VPg-binding region (VPg-BR) of AtRH8 and PpDDX and show that overexpression of the VPg-BR either from AtRH8 or PpDDXL suppresses virus accumulation.  相似文献   

6.
7.
Fructose (Fru) is a major storage form of sugars found in vacuoles, yet the molecular regulation of vacuolar Fru transport is poorly studied. Although SWEET17 (for SUGARS WILL EVENTUALLY BE EXPORTED TRANSPORTERS17) has been characterized as a vacuolar Fru exporter in leaves, its expression in leaves is low. Here, RNA analysis and SWEET17-β-glucuronidase/-GREEN FLUORESCENT PROTEIN fusions expressed in Arabidopsis (Arabidopsis thaliana) reveal that SWEET17 is highly expressed in the cortex of roots and localizes to the tonoplast of root cells. Expression of SWEET17 in roots was inducible by Fru and darkness, treatments that activate accumulation and release of vacuolar Fru, respectively. Mutation and ectopic expression of SWEET17 led to increased and decreased root growth in the presence of Fru, respectively. Overexpression of SWEET17 specifically reduced the Fru content in leaves by 80% during cold stress. These results intimate that SWEET17 functions as a Fru-specific uniporter on the root tonoplast. Vacuoles overexpressing SWEET17 showed increased [14C]Fru uptake compared with the wild type. SWEET17-mediated Fru uptake was insensitive to ATP or treatment with NH4Cl or carbonyl cyanide m-chlorophenyl hydrazone, indicating that SWEET17 functions as an energy-independent facilitative carrier. The Arabidopsis genome contains a close paralog of SWEET17 in clade IV, SWEET16. The predominant expression of SWEET16 in root vacuoles and reduced root growth of mutants under Fru excess indicate that SWEET16 also functions as a vacuolar transporter in roots. We propose that in addition to a role in leaves, SWEET17 plays a key role in facilitating bidirectional Fru transport across the tonoplast of roots in response to metabolic demand to maintain cytosolic Fru homeostasis.Sugars are main energy sources for generating ATP, major precursors to various storage carbohydrates as well as key signaling molecules important for normal growth in higher plants (Rolland et al., 2006). Depending on the metabolic demand, sugars are translocated over long distances or stored locally. SWEET (for SUGARS WILL EVENTUALLY BE EXPORTED TRANSPORTERS) and SUC/SUT (for Sucrose transporter/Sugar transporter)-type transporters are responsible for transfer of Suc from the phloem parenchyma into the sieve element companion cell complex for long-distance translocation (Riesmeier et al., 1992; Sauer, 2007; Kühn and Grof, 2010; Chen et al., 2012). Suc or hexoses derived from Suc hydrolysis in the cell wall are then taken up into sink cells by SUT (Braun and Slewinski, 2009) or monosaccharide transporters, such as sugar transporter1 (Sauer et al., 1990; Pego and Smeekens, 2000; Sherson et al., 2003). Alternatively, sugars are thought to move between cells via plasmodesmata (Voitsekhovskaja et al., 2006; Ayre, 2011). Major sugar storage pools within plant cells are soluble sugars stored in the vacuole, starch in plastids, and lipids in oil bodies.Vacuoles, which can account for approximately 90% of the cell volume (Winter et al., 1993), play central roles in temporary and long-term storage of soluble sugars (Martinoia et al., 2007; Etxeberria et al., 2012). Some agriculturally important crops such as sugar beet (Beta vulgaris; Leigh, 1984; Getz and Klein, 1995), citrus (Citrus spp.; Echeverria and Valich, 1988), sugarcane (Saccharum officinarum; Thom et al., 1982), and carrot (Daucus carota; Keller, 1988) can store considerable amounts (>10% of plant dry weight) of Suc, Glc, or Fru in vacuoles of the storage parenchyma. Due to a high capacity of vacuoles for storing sugars, vacuolar sugars can serve as an important carbohydrate source during energy starvation, e.g. after starch has been exhausted (Echeverria and Valich, 1988), as well as for the production of other compounds (e.g. osmoprotectants). Sugars are known to regulate photosynthesis; therefore, the release of sugars from vacuoles could be important for modulating photosynthesis (Kaiser and Heber, 1984). Moreover, vacuole-derived sugars are commercially used to produce biofuels, such as ethanol, from sugarcane. Knowledge of the key transporters involved in sugar exchange between the vacuole and cytoplasm is thus relevant in the context of bioenergy (Grennan and Gragg, 2009).To facilitate the exchange of sugars across the tonoplast, plant vacuoles are equipped with a multitude of transporters (Neuhaus, 2007; Etxeberria et al., 2012; Martinoia et al., 2012) comprising both facilitated diffusion and active transport systems of vacuolar sugars (Martinoia et al., 2000). Typically, Suc is actively imported into vacuoles by tonoplast monosaccharide transporter (AtTMT1/AtTMT2; Schulz et al., 2011) and exported by the SUT4 family (AtSUC4, OsSUT2; Eom et al., 2011; Payyavula et al., 2011; Schulz et al., 2011). Two H+-dependent sugar antiporters, the vacuolar Glc transporter (AtVGT1; Aluri and Büttner, 2007) and AtTMT1 (Wormit et al., 2006), mediate Glc uptake across the tonoplast to promote carbohydrate accumulation in Arabidopsis (Arabidopsis thaliana). The Early Responsive to Dehydration-Like6 protein has been shown to export vacuolar Glc into the cytosol (Poschet et al., 2011), likely via an energy-independent diffusion mechanism (Yamada et al., 2010). Defects in these vacuolar sugar transporters alter carbohydrate partitioning and allocation and inhibit plant growth and seed yield (Aluri and Büttner, 2007; Wingenter et al., 2010; Eom et al., 2011; Poschet et al., 2011).In contrast to numerous studies on vacuolar transport of Suc and Glc, limited efforts have been devoted to the molecular mechanism of vacuolar Fru transport even though Fru is predominantly located in vacuoles (Martinoia et al., 1987; Voitsekhovskaja et al., 2006; Tohge et al., 2011). Vacuolar Fru is important for the regulation of turgor pressure (Pontis, 1989), antioxidative defense (Bogdanović et al., 2008), and signal transduction during early seedling development (Cho and Yoo, 2011; Li et al., 2011). Thus, control of Fru transport across the tonoplast is thought to be important for plant growth and development. One vacuolar Glc transporter from the Arabidopsis monosaccharide transporter family, VGT1, has been reported to mediate low-affinity Fru uptake when expressed in yeast (Saccharomyces cerevisiae) vacuoles (Aluri and Büttner, 2007). Yet, the high vacuolar uptake activity to Fru intimates the existence of additional high-capacity Fru-specific vacuolar transporters (Thom et al., 1982). Recently, quantitative mapping of a quantitative trait locus for Fru content of leaves led to the identification of the Fru-specific vacuolar transporter SWEET17 (Chardon et al., 2013).SWEET17 belongs to the recently identified SWEET (PFAM:PF03083) super family, which contains 17 members in Arabidopsis and 21 in rice (Oryza sativa; Chen et al., 2010; Frommer et al., 2013; Xuan et al., 2013). Based on homology with 27% to 80% amino acid identity, plant SWEET proteins were grouped into four subclades (Chen et al., 2010). Analysis of GFP fusions indicated that most SWEET transporters are plasma membrane localized. Transport assays using radiotracers in Xenopus laevis oocytes and sugar nanosensors in mammalian cells showed that they function as largely pH-independent low-affinity uniporters with both uptake and efflux activity (Chen et al., 2010, 2012). In particular, clade I and II SWEETs transport monosaccharides and clade III SWEETs transport disaccharides, mainly Suc (Chen et al., 2010, 2012). Mutant phenotypes and developmental expression of several SWEET transporters support important roles in sugar translocation between organs. The clade III SWEETs, in particular SWEET11 and 12, mediate the key step of Suc efflux from phloem parenchyma cells for phloem translocation (Chen et al., 2012). Moreover, SWEETs are coopted by pathogens, likely to provide energy resources and carbon at the site of infection (Chen et al., 2010). Mutations of SWEET8/Ruptured pollen grain1 in Arabidopsis, and RNA inhibition of OsSWEET11 (also called Os8N3 or Xa13) in rice, and petunia (Petunia hybrida) NEC1 resulted in male sterility (Ge et al., 2001; Yang et al., 2006; Guan et al., 2008), possibly caused by inhibiting the Glc supply to developing pollen (Guan et al., 2008). Interestingly, two members, SWEET16 and SWEET17, of the family localize to the tonoplast (Chardon et al., 2013; Klemens et al., 2013). Allelic variation or mutations that affect SWEET17 expression caused Fru accumulation in Arabidopsis leaves, indicating that it plays a key role in exporting Fru from leaf vacuoles (Chardon et al., 2013). A more recent study demonstrated that SWEET16 also functions as a vacuolar sugar transporter (Klemens et al., 2013). Surprisingly, however, SWEET17 expression in mature leaves was comparatively low (Chardon et al., 2013), which leads us to ask whether SWEET17 could mainly function in other tissues under specific developmental or environmental conditions. Although Arabidopsis SWEET17 has been shown to transport Fru in a heterologous system where it accumulated in part at the plasma membrane (Chardon et al., 2013), the biochemical properties of SWEET17 were still elusive. SWEET16 and SWEET17 from Arabidopsis belong to the clade IV SWEETs. Whether clade IV proteins both transport vacuolar sugars in planta deserves further studies.Here, we used GUS/GFP fusions to reveal the root-dominant expression and vacuolar localization of the SWEET17 protein in vivo and its regulation by Fru levels. Phenotypes of mutants and overexpressors were consistent with a role of SWEET17 in bidirectional Fru transport across root vacuoles. The uniport feature of SWEET17 transport was further confirmed using isolated mesophyll vacuoles. Similarly, SWEET16 is also shown to function in vacuolar sugar transport in roots. Our work, performed in parallel to the two other studies (Chardon et al., 2013; Klemens et al., 2013), provides direct evidence for Fru uniport by SWEET17 and presents functional analyses to uncover important roles of these vacuolar transporters in maintaining intracellular Fru homeostasis in roots.  相似文献   

8.
9.
10.
11.
The endoplasmic reticulum (ER) consists of dynamically changing tubules and cisternae. In animals and yeast, homotypic ER membrane fusion is mediated by fusogens (atlastin and Sey1p, respectively) that are membrane-associated dynamin-like GTPases. In Arabidopsis (Arabidopsis thaliana), another dynamin-like GTPase, ROOT HAIR DEFECTIVE3 (RHD3), has been proposed as an ER membrane fusogen, but direct evidence is lacking. Here, we show that RHD3 has an ER membrane fusion activity that is enhanced by phosphorylation of its C terminus. The ER network was RHD3-dependently reconstituted from the cytosol and microsome fraction of tobacco (Nicotiana tabacum) cultured cells by exogenously adding GTP, ATP, and F-actin. We next established an in vitro assay system of ER tubule formation with Arabidopsis ER vesicles, in which addition of GTP caused ER sac formation from the ER vesicles. Subsequent application of a shearing force to this system triggered the formation of tubules from the ER sacs in an RHD-dependent manner. Unexpectedly, in the absence of a shearing force, Ser/Thr kinase treatment triggered RHD3-dependent tubule formation. Mass spectrometry showed that RHD3 was phosphorylated at multiple Ser and Thr residues in the C terminus. An antibody against the RHD3 C-terminal peptide abolished kinase-triggered tubule formation. When the Ser cluster was deleted or when the Ser residues were replaced with Ala residues, kinase treatment had no effect on tubule formation. Kinase treatment induced the oligomerization of RHD3. Neither phosphorylation-dependent modulation of membrane fusion nor oligomerization has been reported for atlastin or Sey1p. Taken together, we propose that phosphorylation-stimulated oligomerization of RHD3 enhances ER membrane fusion to form the ER network.In eukaryotic cells, the endoplasmic reticulum (ER) is the organelle with the largest membrane area. The ER consists of an elaborate network of interconnected membrane tubules and cisternae that is continually moving and being remodeled (Friedman and Voeltz, 2011). In plant cells, ER movement and remodeling is primarily driven by the actin-myosin XI cytoskeleton (Sparkes et al., 2009; Ueda et al., 2010; Yokota et al., 2011; Griffing et al., 2014) and secondarily by the microtubule cytoskeleton (Hamada et al., 2014). Several factors involved in creating the ER architecture have been also identified (Anwar et al., 2012; Chen et al., 2012; Goyal and Blackstone, 2013; Sackmann, 2014; Stefano et al., 2014a; Westrate et al., 2015). Among them, ER membrane-bound GTPases, animal atlastins and yeast Sey1p (Synthetic Enhancement of Yop1), function as ER fusogens to form the interconnected tubular network (Hu et al., 2009; Orso et al., 2009; Anwar et al., 2012). Atlastin molecules on the two opposed membranes have been proposed to transiently dimerize to attract the two membranes to each other (Bian et al., 2011; Byrnes and Sondermann, 2011; Morin-Leisk et al., 2011; Moss et al., 2011; Lin et al., 2012; Byrnes et al., 2013). Closely attracted lipid bilayers are supposed to be destabilized by an amphipathic helical domain at the atlastin C terminus to facilitate membrane fusion (Bian et al., 2011; Liu et al., 2012; Faust et al., 2015). Knockdown of atlastins leads to fragmentation of the ER and unbranched ER tubules, while overexpression of atlastins enhances ER membrane fusion, which enlarges the ER profiles (Hu et al., 2009; Orso et al., 2009).An Arabidopsis (Arabidopsis thaliana) protein, ROOT HAIR DEFECTIVE3 (RHD3), has been proposed as a fusogen because (1) when it is disrupted, the ER network is modified into large cable-like strands of poorly branched membranes (Zheng et al., 2004; Chen et al., 2011; Stefano et al., 2012), (2) it shares sequence similarity with the above-mentioned fusogen Sey1p (Hu et al., 2009), and (3) it has structural similarity to atlastin and Sey1p, with a functional GTPase domain at the N-terminal cytosolic domain (Stefano et al., 2012) followed by two transmembrane domains and a cytosolic tail. RHD3 has a longer cytosolic C-terminal tail than do atlastin and Sey1p (Stefano and Brandizzi, 2014). It contains not only an amphipathic region but also a Ser/Thr-rich C terminus.Arabidopsis has two RHD3 isoforms called RHD3-Like 1 and RHD3-Like 2. Fluorescently tagged RHD3 and RHD3-Like 2 localize to the ER (Chen et al., 2011; Stefano et al., 2012; Lee et al., 2013). RHD3 and the two RHD3-Like proteins likely have redundant roles in ER membrane fusion (Zhang et al., 2013). Overexpression of either RHD3 or RHD3-Like 2 with a defective GTPase domain phenocopies the aberrant ER morphology in rhd3-deficient mutants (Chen et al., 2011; Lee et al., 2013).In this study, we show that the Ser/Thr-rich C terminus enhances ER membrane fusion following phosphorylation of its C terminus. We propose a model in which phosphorylation and oligomerization of RHD3 is required for efficient ER membrane fusion. Our findings clarify the mechanisms that regulate RHD3 and consequently the homeostasis of membrane fusion in the ER.  相似文献   

12.
Initial pollen-pistil interactions in the Brassicaceae are regulated by rapid communication between pollen grains and stigmatic papillae and are fundamentally important, as they are the first step toward successful fertilization. The goal of this study was to examine the requirement of exocyst subunits, which function in docking secretory vesicles to sites of polarized secretion, in the context of pollen-pistil interactions. One of the exocyst subunit genes, EXO70A1, was previously identified as an essential factor in the stigma for the acceptance of compatible pollen in Arabidopsis (Arabidopsis thaliana) and Brassica napus. We hypothesized that EXO70A1, along with other exocyst subunits, functions in the Brassicaceae dry stigma to deliver cargo-bearing secretory vesicles to the stigmatic papillar plasma membrane, under the pollen attachment site, for pollen hydration and pollen tube entry. Here, we investigated the functions of exocyst complex genes encoding the remaining seven subunits, SECRETORY3 (SEC3), SEC5, SEC6, SEC8, SEC10, SEC15, and EXO84, in Arabidopsis stigmas following compatible pollinations. Stigma-specific RNA-silencing constructs were used to suppress the expression of each exocyst subunit individually. The early postpollination stages of pollen grain adhesion, pollen hydration, pollen tube penetration, seed set, and overall fertility were analyzed in the transgenic lines to evaluate the requirement of each exocyst subunit. Our findings provide comprehensive evidence that all eight exocyst subunits are necessary in the stigma for the acceptance of compatible pollen. Thus, this work implicates a fully functional exocyst complex as a component of the compatible pollen response pathway to promote pollen acceptance.In flowering plants, sexual reproduction occurs as a result of constant communication between the male gametophyte and the female reproductive organ, from the initial acceptance of compatible pollen to final step of successful fertilization (for review, see Beale and Johnson, 2013; Dresselhaus and Franklin-Tong, 2013; Higashiyama and Takeuchi, 2015). In the Brassicaceae, the stigmas that present a receptive surface for pollen are categorized as dry and covered with unicellular papillae (Heslop-Harrison and Shivanna, 1977). Communication is initiated rapidly following contact of a pollen grain with a stigmatic papilla, as the role of the papillae is to regulate the early cellular responses leading to compatible pollen germination. The basal compatible pollen recognition response also presents a barrier to foreign pollen or is inhibited with self-incompatible pollen (for review, see Dickinson, 1995; Hiscock and Allen, 2008; Chapman and Goring, 2010; Indriolo et al., 2014b).The initial adhesive interaction between the pollen grain and the papilla cell in the Brassicaceae is mediated by the exine of the pollen grain and the surface of the stigmatic papilla (Preuss et al., 1993; Zinkl et al., 1999). A stronger connection results between the adhered pollen grain and the stigmatic papilla with the formation of a lipid-protein interface (foot) derived from the pollen coat and the stigmatic papillar surface (Mattson et al., 1974; Stead et al., 1980; Gaude and Dumas, 1986; Elleman and Dickinson, 1990; Elleman et al., 1992; Preuss et al., 1993; Mayfield et al., 2001). It is at this point that a Brassicaceae-specific recognition of compatible pollen is proposed to occur (Hülskamp et al., 1995; Pruitt, 1999), though the nature of this recognition system is not clearly defined. Two stigma-specific Brassica oleracea glycoproteins, the S-Locus Glycoprotein and S-Locus Related1 (SLR1) protein, play a role in compatible pollen adhesion (Luu et al., 1997, 1999), potentially through interactions with the pollen coat proteins, PCP-A1 and SLR1-BP, respectively (Doughty et al., 1998; Takayama et al., 2000). The simultaneous recognition of self-incompatible pollen would also take place at this stage (for review, see Dresselhaus and Franklin-Tong, 2013; Indriolo et al., 2014b; Sawada et al., 2014). Thus, this interface not only provides a strengthened bond between the pollen grain and stigmatic papilla, but likely facilitates the interaction of signaling proteins from both partners to promote specific cellular responses in the stigmatic papilla toward the pollen grain.One response regulated by these interactions is the release of water from the stigmatic papilla to the adhered compatible pollen grain to enable the pollen grain to rehydrate, germinate, and produce a pollen tube (Zuberi and Dickinson, 1985; Preuss et al., 1993). Upon hydration, the pollen tube emerges at the site of pollen-papilla contact and penetrates the stigma surface between the plasma membrane and the overlaying cell wall (Elleman et al., 1992; Kandasamy et al., 1994). Pollen tube entry into the stigmatic surface represents a second barrier, selecting compatible pollen tubes. Subsequently, the compatible pollen tubes traverse down to the base of the stigma, enter the transmitting tract, and grow intracellularly toward ovules for fertilization. Pollen-pistil interactions at these later stages are also highly regulated (for review, see Beale and Johnson, 2013; Dresselhaus and Franklin-Tong, 2013; Higashiyama and Takeuchi, 2015).EXO70A1, a subunit of the exocyst, was identified as a factor involved in early pollen-stigma interactions, where it is required in the stigma for the acceptance of compatible pollen and inhibited by the self-incompatibility response (Samuel et al., 2009). Stigmas from the Arabidopsis (Arabidopsis thaliana) exo70A1 mutant display constitutive rejection of wild-type-compatible pollen (Samuel et al., 2009; Safavian et al., 2014). This stigmatic defect was rescued by the stigma-specific expression of an Red Fluorescent Protein (RFP):EXO70A1 transgene (Samuel et al., 2009) or partially rescued by providing a high relative humidity environment (Safavian et al., 2014). In addition, the stigma-specific expression of an EXO70A1 RNA interference construct in Brassica napus ‘Westar’ resulted in impaired compatible pollen acceptance and a corresponding reduction in seed production compared with compatible pollinations with wild-type B. napus ‘Westar’ pistils (Samuel et al., 2009). From these studies, EXO70A1 was found to be a critical component in stigmatic papillae to promote compatible pollen hydration and pollen tube entry through the stigma surface. One of the functions of the exocyst is to mediate polar secretion (for review, see Heider and Munson, 2012; Zárský et al., 2013; Synek et al., 2014). Consistent with this, previous studies have observed vesicle-like structures in proximity to the stigmatic papillar plasma membrane in response to compatible pollen in both Brassica spp. and Arabidopsis species (Elleman and Dickinson, 1990, 1996; Dickinson, 1995; Safavian and Goring, 2013; Indriolo et al., 2014a). The secretory activity is predicted to promote pollen hydration and pollen tube entry. As well, consistent with the proposed inhibition of EXO70A1 by the self-incompatibility pathway (Samuel et al., 2009), a complete absence or a significant reduction of vesicle-like structures at the stigmatic papillar plasma membrane was observed in the exo70A1 mutant and with self-incompatible pollen (Safavian and Goring, 2013; Indriolo et al., 2014a).The exocyst is a well-defined complex in yeast (Saccharomyces cerevisiae) and animal systems, consisting of eight subunits, SEC3, SEC5, SEC6, SEC8, SEC10, SEC15, EXO70, and EXO84 (TerBush et al., 1996; Guo et al., 1999). Exocyst subunit mutants were first identified in yeast as secretory mutants displaying a cytosolic accumulation of secretory vesicles (Novick et al., 1980). Subsequent work defined roles for the exocyst in vesicle docking at target membranes in processes such as regulated secretion, polarized exocytosis, and cytokinesis to facilitate membrane fusion by Soluble NSF Attachment protein Receptor (SNARE) complexes (for review, see Heider and Munson, 2012; Liu and Guo, 2012). In plants, genes encoding all eight exocyst subunits have been identified, and many of these genes exist as multiple copies. For example, the Arabidopsis genome contains single copy genes for SEC6 and SEC8, two copies each for SECRETORY3 (SEC3), SEC5, SEC10, and SEC15, three EXO84 genes, and 23 EXO70 genes (Chong et al., 2010; Cvrčková et al., 2012; Vukašinović et al., 2014). Ultrastructural studies using electron tomography uncovered the existence of a structure resembling the exocyst in Arabidopsis (Otegui and Staehelin, 2004; Seguí-Simarro et al., 2004). Localization studies of specific Arabidopsis exocyst subunits also supported conserved roles in polarized exocytosis and cytokinesis in plants. Localization studies have shown EXO70, SEC6, and SEC8 at the growing tip of pollen tubes (Hála et al., 2008), EXO70A1 at the stigmatic papillar plasma membrane (Samuel et al., 2009), SEC3a, SEC6, SEC8, SEC15b, EXO70A1, and EXO84b at the root epidermal cell plasma membrane and developing cell plate (Fendrych et al., 2010, 2013; Wu et al., 2013; Zhang et al., 2013; Rybak et al., 2014), and SEC3a at the plasma membrane in the embryo and root hair (Zhang et al., 2013). Similar to the yeast exocyst mutants, vesicle accumulation has also been observed in the exo70A1 and exo84b mutants (Fendrych et al., 2010; Safavian and Goring, 2013). Taken together, these findings strongly support that plant exocyst subunits function in vivo in vesicle docking at sites of polarized secretion and cytokinesis (for review, see Zárský et al., 2013). In support of this, a recent study investigating Transport Protein Particle (TRAPP)II and exocyst complexes during cytokinesis in Arabidopsis has identified all eight exocyst components in immunoprecipitated complexes (SEC3a/SEC3b, SEC5a, SEC6, SEC8, SEC10, SEC15b, EXO70A1, EXO70H2, and EXO84b; Rybak et al., 2014).Several plant exocyst subunit genes have been implicated in biological processes that rely on regulated vesicle trafficking, where corresponding mutants have displayed a range of growth defects. At the cellular level, these phenotypes have been associated with decreased cell elongation and polar growth (Cole et al., 2005, 2014; Wen et al., 2005; Synek et al., 2006), defects in cytokinesis and cell plate formation (Fendrych et al., 2010; Wu et al., 2013; Rybak et al., 2014), and disrupted Pin-Formed (PIN) auxin efflux carrier recycling and polar auxin transport (Drdová et al., 2013). Several Arabidopsis subunit mutants display strong growth defects such as the sec3a mutant with an embryo-lethal phenotype (Zhang et al., 2013), sec6, sec8, and exo84b mutants with severely dwarfed phenotypes and defects in root growth (Fendrych et al., 2010; Wu et al., 2013; Cole et al., 2014), and exo70A1 with a milder dwarf phenotype (Synek et al., 2006). The Arabidopsis exo70A1 mutant has also been reported to have defects in root hair elongation, hypocotyl elongation, compatible pollen acceptance, seed coat deposition, and tracheary element differentiation (Synek et al., 2006; Samuel et al., 2009; Kulich et al., 2010; Li et al., 2013). Essential roles for other exocyst subunits include Arabidopsis SEC5a/SEC5b, SEC6, SEC8, and SEC15a/SEC15b in male gametophyte development and pollen tube growth (Cole et al., 2005; Hála et al., 2008; Wu et al., 2013), SEC8 in seed coat deposition (Kulich et al., 2010), SEC5a, SEC8, EXO70A1, and EXO84b in root meristem size and root cell elongation (Cole et al., 2014), and a maize (Zea mays) SEC3 homolog in root hair elongation (Wen et al., 2005). Finally, the Arabidopsis EXO70B1, EXO70B2, and EXO70H1 subunits have been implicated in plant defense responses (Pecenková et al., 2011; Stegmann et al., 2012; Kulich et al., 2013; Stegmann et al., 2013).Even with these detailed studies on the functions of exocyst subunits in plants, a systematic demonstration of the requirement of all eight exocyst subunits in a specific plant biological process is currently lacking. EXO70A1 was previously identified as an essential factor in the stigma for compatible pollen-pistil interactions in Arabidopsis and B. napus (Samuel et al., 2009), and we hypothesized that this protein functions as part of the exocyst complex to tether post-Golgi secretory vesicles to stigmatic papillar plasma membrane (Safavian and Goring, 2013). To provide support for the proposed biological role of the exocyst in the stigma for compatible pollen acceptance, we investigated the roles of the remaining seven subunits, SEC3, SEC5, SEC6, SEC8, SEC10, SEC15, and EXO84, in Arabidopsis stigmatic papillae. Given that some Arabidopsis exocyst subunits were previously determined to be essential at earlier growth stages, stigma-specific RNA-silencing constructs were used for each exocyst subunit, and the early postpollination stages were analyzed for these transgenic lines. Our collective data demonstrates that all eight exocyst subunits are required in the stigma for the early stages of compatible pollen-pistil interactions.  相似文献   

13.
Sugars, such as sucrose and glucose, have been implicated in the regulation of diverse developmental events in plants and other organisms. We isolated an Arabidopsis (Arabidopsis thaliana) mutant, sugar-insensitive3 (sis3), that is resistant to the inhibitory effects of high concentrations of exogenous glucose and sucrose on early seedling development. In contrast to wild-type plants, sis3 mutants develop green, expanded cotyledons and true leaves when sown on medium containing high concentrations (e.g. 270 mm) of sucrose. Unlike some other sugar response mutants, sis3 exhibits wild-type responses to the inhibitory effects of abscisic acid and paclobutrazol, a gibberellic acid biosynthesis inhibitor, on seed germination. Map-based cloning revealed that SIS3 encodes a RING finger protein. Complementation of the sis3-2 mutant with a genomic SIS3 clone restored sugar sensitivity of sis3-2, confirming the identity of the SIS3 gene. Biochemical analyses demonstrated that SIS3 is functional in an in vitro ubiquitination assay and that the RING motif is sufficient for its activity. Our results indicate that SIS3 encodes a ubiquitin E3 ligase that is a positive regulator of sugar signaling during early seedling development.Almost all living organisms rely on the products of plant photosynthesis for sustenance, either directly or indirectly. Carbohydrates, the major photosynthates, provide both energy and carbon skeletons for fungi, plants, and animals. In addition, sugars, such as Suc and Glc, function as signaling molecules to regulate plant growth, development, gene expression, and metabolic processes. Sugar response pathways are integrated with other signaling pathways, such as those for light, phytohormones, stress, and nitrogen (Dijkwel et al., 1997; Zhou et al., 1998; Roitsch, 1999; Arenas-Huertero et al., 2000; Huijser et al., 2000; Laby et al., 2000; Coruzzi and Zhou, 2001; Rook et al., 2001; Rolland et al., 2006).Several components of plant sugar response pathways have been identified based on the conservation of sugar-sensing mechanisms among eukaryotic cells (Rolland et al., 2001, 2006) or by mutant screens. Yeast HEXOKINASE2 functions in the Glc-mediated catabolite repression pathway (Entian, 1980). In Arabidopsis (Arabidopsis thaliana), mutations in HEXOKINASE1 (HXK1) cause a Glc-insensitive phenotype, and HXK1 demonstrates dual functions in Glc sensing and metabolism (Moore et al., 2003; Cho et al., 2006). Recent studies revealed the involvement of G-protein-coupled receptor systems in sugar response in yeast and Arabidopsis (Chen et al., 2003; Lemaire et al., 2004). Arabidopsis regulator of G-protein signaling1 (rgs1) mutant seedlings are insensitive to 6% Glc (Chen and Jones, 2004), whereas G-protein α-subunit (gpa1) null mutant seedlings are hypersensitive to Glc (Chen et al., 2003). The SNF1/AMPK/SnRK1 protein kinases are postulated to be global regulators of energy control (Polge and Thomas, 2007). Studies conducted on two members of the Arabidopsis SnRK1 (for SNF1-Related Protein Kinases1) family, AKIN10 and AKIN11, have revealed their pivotal roles in stress and sugar signaling (Baena-González et al., 2007). A genetic screen for reduced seedling growth on 175 mm Suc identified the pleiotropic regulatory locus1 (prl1) mutant, which encodes a nuclear WD protein. Further analyses revealed that PRL1 functions in Glc and phytohormone responses (Németh et al., 1998). Interestingly, PRL1 negatively regulates the Arabidopsis SnRK1s AKIN10 and AKIN11 in vitro (Bhalerao et al., 1999).Isolation of additional mutants defective in sugar response has revealed cross talk between sugar and phytohormone response pathways. For example, abscisic acid (ABA) biosynthesis and signaling mutants have been isolated by several genetic screens for seedlings with reduced responses to the inhibitory effects of high levels of Suc or Glc on seedling development. These mutants include abscisic acid-deficient1 (aba1), aba2, aba3, salt-tolerant1/nine-cis-epoxycarotenoid dioxygenase3, abscisic acid-insensitive3 (abi3), and abi4 (Arenas-Huertero et al., 2000; Huijser et al., 2000; Laby et al., 2000; Rook et al., 2001; Cheng et al., 2002; Rolland et al., 2002; Huang et al., 2008), indicating interplay between ABA- and sugar-mediated signaling. Ethylene also exhibits interactions with sugars in controlling seedling development. Both the ethylene overproduction mutant eto1 and the constitutive ethylene response mutant ctr1 exhibit Glc (Zhou et al., 1998) and Suc (Gibson et al., 2001) insensitivity, whereas the ethylene-insensitive mutants etr1, ein2, and ein4 show sugar hypersensitivity (Zhou et al., 1998; Gibson et al., 2001; Cheng et al., 2002).Further characterization of sugar response factors has suggested that ubiquitin-mediated protein degradation may play a role in sugar response. In particular, the PRL1-binding domains of SnRK1s have been shown to recruit SKP1/ASK1, a conserved SCF ubiquitin ligase subunit, as well as the α4/PAD1 proteasomal subunit, indicating a role for SnRK1s in mediating proteasomal binding of SCF ubiquitin ligases (Farrás et al., 2001). In addition, recent studies indicate that PRL1 is part of a CUL4-based E3 ligase and that AKIN10 exhibits decreased rates of degradation in prl1 than in wild-type extracts (Lee et al., 2008). The ubiquitin/26S proteasome pathway plays important roles in many cellular processes and signal transduction pathways in yeast, animals, and plants (Hochstrasser, 1996; Hershko and Ciechanover, 1998; Smalle and Vierstra, 2004). The key task of the pathway is to selectively ubiquitinate substrate proteins and target them for degradation by the 26S proteasome. In short, the multistep ubiquitination process starts with the formation of a thiol-ester linkage between ubiquitin and a ubiquitin-activating enzyme (E1). The activated ubiquitin is then transferred to a ubiquitin-conjugating enzyme (E2), and a ubiquitin protein ligase (E3) then mediates the covalent attachment of ubiquitin to the substrate protein. The specificity of the pathway is largely realized by the E3s, which recognize the substrates that should be ubiquitinated. In Arabidopsis, more than 1,300 genes encode putative E3 subunits and the E3 ligases can be grouped into defined families based upon the presence of HECT (for Homology to E6-AP C Terminus), RING (for Really Interesting New Gene), or U-box domains (Smalle and Vierstra, 2004). The RING-type E3s can be subdivided into single-subunit E3s, which contain the substrate recognition and RING finger domains on the same protein, and multisubunit E3s, which include the SCF (for Skp1-Cullin-F-box), CUL3-BTB (for Broad-complex, Tramtrack, Bric-a-Brac), and APC (for Anaphase-Promoting Complex) complexes (Weissman, 2001; Moon et al., 2004).The Cys-rich RING finger was first described in the early 1990s (Freemont et al., 1991). It is defined as a linear series of conserved Cys and His residues (C3HC/HC3) that bind two zinc atoms in a cross-brace arrangement. RING fingers can be divided into two types, C3HC4 (RING-HC) and C3H2C3 (RING-H2), depending on the presence of either a Cys or a His residue in the fifth position of the motif (Lovering et al., 1993; Freemont, 2000). A recent study of the RING finger ubiquitin ligase family encoded by the Arabidopsis genome resulted in the identification of 469 predicted proteins containing one or more RING domains (Stone et al., 2005). However, the in vivo biological functions of all but a few of the RING proteins remain unknown. Recent studies have implicated several Arabidopsis RING proteins in a variety biological processes, including COP1 and CIP8 (photomorphogenesis; Hardtke et al., 2002; Seo et al., 2004), SINAT5 (auxin signaling; Xie et al., 2002), ATL2 (defense signaling; Serrano and Guzman, 2004), BRH1 (brassinosteroid response; Molnár et al., 2002), RIE1 (seed development; Xu and Li, 2003), NLA (nitrogen limitation adaptation; Peng et al., 2007), HOS1 (cold response; Dong et al., 2006), AIP2 (ABA signaling; Zhang et al., 2005), KEG (ABA signaling; Stone et al., 2006), and SDIR1 (ABA signaling; Zhang et al., 2007).Here, we report the isolation, identification, and characterization of an Arabidopsis mutant, sugar-insensitive3 (sis3), which is resistant to the early seedling developmental arrest caused by high exogenous sugar levels. The responsible locus, SIS3, was identified through a map-based cloning approach and confirmed with additional T-DNA insertional mutants and complementation tests. The SIS3 gene encodes a protein with a RING-H2 domain and three putative transmembrane domains. Glutathione S-transferase (GST)-SIS3 recombinant proteins exhibit in vitro ubiquitin E3 ligase activity. Together, these results indicate that a ubiquitination pathway involving the SIS3 RING protein is required to mediate the sugar response during early seedling development.  相似文献   

14.
In plants and algae, the serine/threonine kinase STN7/STT7, orthologous protein kinases in Chlamydomonas reinhardtii and Arabidopsis (Arabidopsis thaliana), respectively, is an important regulator in acclimation to changing light environments. In this work, we assessed STT7-dependent protein phosphorylation under high light in C. reinhardtii, known to fully induce the expression of LIGHT-HARVESTING COMPLEX STRESS-RELATED PROTEIN3 (LHCSR3) and a nonphotochemical quenching mechanism, in relationship to anoxia where the activity of cyclic electron flow is stimulated. Our quantitative proteomics data revealed numerous unique STT7 protein substrates and STT7-dependent protein phosphorylation variations that were reliant on the environmental condition. These results indicate that STT7-dependent phosphorylation is modulated by the environment and point to an intricate chloroplast phosphorylation network responding in a highly sensitive and dynamic manner to environmental cues and alterations in kinase function. Functionally, the absence of the STT7 kinase triggered changes in protein expression and photoinhibition of photosystem I (PSI) and resulted in the remodeling of photosynthetic complexes. This remodeling initiated a pronounced association of LHCSR3 with PSI-LIGHT HARVESTING COMPLEX I (LHCI)-ferredoxin-NADPH oxidoreductase supercomplexes. Lack of STT7 kinase strongly diminished PSII-LHCII supercomplexes, while PSII core complex phosphorylation and accumulation were significantly enhanced. In conclusion, our study provides strong evidence that the regulation of protein phosphorylation is critical for driving successful acclimation to high light and anoxic growth environments and gives new insights into acclimation strategies to these environmental conditions.Oxygenic photosynthesis converts solar energy into chemical energy. This energy is utilized for carbon dioxide assimilation, allowing the formation of complex organic material. Plant photosynthesis is performed by a series of reactions in and at the thylakoid membrane, resulting in light-dependent water oxidation, NADP reduction, and ATP formation (Whatley et al., 1963). These light reactions are catalyzed by two photosystems (PSI and PSII). A third multiprotein complex, also embedded in the thylakoid membrane, is the cytochrome b6f (cyt b6f) complex that links photosynthetic electron transfer processes between the two photosystems and functions in proton translocation. The ATP synthase takes advantage of the proton-motive force that is generated by the light reactions (Mitchell, 1961) to produce ATP. ATP and NADPH, generated through linear electron flow from PSII to PSI, drive the Calvin-Benson-Bassham cycle (Bassham et al., 1950) to fix CO2. Alternatively, cyclic electron flow (CEF) between PSI and the cyt b6f complex solely produces ATP (Arnon, 1959).Under normal growth conditions, CEF provides additionally required ATP for CO2 fixation (Lucker and Kramer, 2013), counteracts overreduction of the PSI acceptor side under stressful environmental cues, and readjusts the ATP poise, leading to increased lumen acidification important for photoprotection (Alric, 2010; Peltier et al., 2010; Leister and Shikanai, 2013; Shikanai, 2014). In microalgae and vascular plants, CEF relies on the NAD(P)H dehydrogenase-dependent and/or PROTON GRADIENT REGULATION5 (PGR5)-related pathways (Munekage et al., 2002, 2004; Petroutsos et al., 2009; Tolleter et al., 2011; Johnson et al., 2014). For both pathways, supercomplexes consisting of PSI-LIGHT HARVESTING COMPLEX I (LHCI) and components of the respective electron transfer routes have been identified. In Arabidopsis (Arabidopsis thaliana), a unique NAD(P)H dehydrogenase-PSI supercomplex with a molecular mass of more than 1,000 kD was discovered (Peng et al., 2008). From Chlamydomonas reinhardtii, Iwai et al. (2010) isolated a protein supercomplex composed of PSI-LHCI, LHCII, the cyt b6/f complex, ferredoxin-NADPH oxidoreductase (FNR), and PROTON GRADIENT REGULATION-LIKE1 (PGRL1).PGRL1 and PGR5 interact physically in Arabidopsis and associate with PSI to allow the operation of CEF (DalCorso et al., 2008). Functional data suggest that PGRL1 might operate as a ferredoxin-plastoquinone reductase (Hertle et al., 2013). The PGRL1-containing CEF supercomplex isolated from C. reinhardtii is capable of CEF under in vitro conditions in the presence of exogenously added soluble plastocyanin and ferredoxin (Iwai et al., 2010). Terashima et al. (2012) isolated a CEF supercomplex of similar composition from anaerobic growth conditions that was active in vitro and contained proteins such as the chloroplast-localized Ca2+ sensor CAS and ANAEROBIC RESPONSE1 (ANR1), which were also shown to be functionally important for efficient CEF in the alga. Notably, it was suggested that the onset of CEF in C. reinhardtii is redox controlled (Takahashi et al., 2013).It has been demonstrated that efficient CEF is crucial for successful acclimation to excess light (Munekage et al., 2004; Dang et al., 2014; Johnson et al., 2014; Kukuczka et al., 2014). The most rapid response to excess light, however, relies on a mechanism called nonphotochemical quenching (NPQ). The fastest constituent of NPQ is energy-dependent (qE) quenching, which operates at a time scale of seconds to minutes and regulates the thermal dissipation of excess absorbed light energy, thereby providing effective photoprotection. In vascular plants, the PSII protein PSII SUBUNIT S is essential for qE (Li et al., 2000), whereas qE induction in the green alga C. reinhardtii is mediated by LIGHT-HARVESTING COMPLEX STRESS-RELATED PROTEIN3 (LHCSR3), an ancient light-harvesting protein that is missing in vascular plants (Peers et al., 2009). CEF and qE are complementary for acclimation to excess light, as double mutants deficient in both mechanisms possess additive phenotypes and are highly sensitive to light (Kukuczka et al., 2014). Another constituent of NPQ is the quenching by state transitions. State transitions are important to balance the excitation energy between PSI and PSII (Bonaventura and Myers, 1969; Murata, 1969). Under light conditions where PSII is preferentially excited, both PSII core and LHCII proteins become phosphorylated (Lemeille and Rochaix, 2010). As a consequence, phosphorylated LHCII proteins detach from PSII and partly connect to PSI (state 2). Under conditions where PSI excitation is predominant, this process is reversed. LHCII proteins are dephosphorylated and associate with PSII (state 1). The extent of state transition between vascular plants such as Arabidopsis and C. reinhardtii differs significantly. The proportion of mobile LHCII antenna is about 80% in the alga, whereas in Arabidopsis, only 15% to 20% of LHCII is transferred to PSI under state 2 conditions (Lemeille and Rochaix, 2010). However, the large increase in PSI antenna size in C. reinhardtii has recently been challenged (Nagy et al., 2014; Ünlü et al., 2014): while 70% to 80% of mobile LHCII detached from PSII in response to transition to state 2 conditions, only a fraction of about 20% functionally attached to PSI.Phosphorylation of LHC proteins requires the function of the STT7 kinase or its ortholog STN7 in C. reinhardtii or Arabidopsis, respectively. In the absence of the STT7/STN7 kinase, the initiation of state transitions is blocked (Depège et al., 2003; Bellafiore et al., 2005). The mobile LHCII fraction of C. reinhardtii includes the two monomeric minor LHCII antenna proteins, CP26 and CP29 (encoded by lhcb5 and lhcb4 genes), and the major chlorophyll a/b binding protein of LHCII, LHCBM5 (Takahashi et al., 2006), but also the LHCSR3 protein was suggested to migrate during state transitions (Allorent et al., 2013). Takahashi et al. (2014) suggested that only CP29 and LHCBM5 directly associate with PSI to form the PSI-LHCI-LHCII supercomplex, while the binding of CP26 could occur indirectly or via the other two proteins. However, it is not yet known whether STT7 directly phosphorylates the LHCII proteins or if this takes place as part of a kinase cascade (Rochaix, 2007). Nevertheless, the direct interaction between STT7 and the LHCII proteins is quite likely, since none of the other chloroplast kinases was found to be specifically required for LHCII phosphorylation (Rochaix, 2014). The activity of the STT7 kinase is mainly determined by the redox status of the plastoquinone pool (Vener et al., 1997; Zito et al., 1999). The identification of a PROTEIN PHOSPHATASE 2C (PP2C)-type phosphatase responsible for the dephosphorylation of the LHCII proteins in Arabidopsis has been described by two studies in parallel pointing to the fact that this enzyme, called PROTEIN PHOSPHATASE1/THYLAKOID-ASSOCIATED PHOSPHATASE38, acts directly on phosphorylated LHCII proteins, in particular when they are associated with the PSI-LHCI supercomplex (Pribil et al., 2010; Shapiguzov et al., 2010). Moreover, it is not known whether these phosphatases are constitutively active or if they are regulated by other means, for example through the redox state of the plastoquinone pool. Nonetheless, both enzymes are conserved in land plants and exhibit orthologous proteins in C. reinhardtii (Rochaix et al., 2012).Another kinase related to STN7/STT7 is encoded in the Arabidopsis and C. reinhardtii genomes and named STN8 and STATE TRANSITION-LIKE1 (STL1), respectively. STN8 is involved in PSII core subunit phosphorylation and influences the repair of PSII after photodamage (Bonardi et al., 2005; Vainonen et al., 2005). Remarkably, the disassembly of the PSII holocomplex is inhibited in STN7/STN8 double mutants (Tikkanen et al., 2008; Fristedt et al., 2009; Dietzel et al., 2011; Nath et al., 2013), suggesting that the phosphorylation of core subunits is required for PSII disassembly. It was further suggested that STN8 controls the transition between linear electron flow and CEF by the phosphorylation of PGRL1 in Arabidopsis (Reiland et al., 2011). As described for STN7, the activity of STN8 is probably regulated via the redox state of the plastoquinone pool (Bennett, 1991; Fristedt et al., 2009). Notably, the action of STN8 is counteracted by a chloroplast PP2C phosphatase (Samol et al., 2012), allowing for the fast reversibility of STN8-mediated acclimation responses. Thus, it appears that an intricate regulatory network of chloroplast protein kinases and phosphatases evolved in vascular plants and algae that drives the acclimation response to various environmental cues, including excess and changing light settings (Rochaix et al., 2012). As STN7/STT7 and STN8/STL1 kinase activities appear to be controlled by the redox poise of the plastoquinone pool, the plastoquinone pool would be a central player in these acclimation responses. On the other hand, the kinases themselves are subjected to phosphorylation (Reiland et al., 2009, 2011; Lemeille et al., 2010; Wang et al., 2013). However, the functional consequences of this phosphorylation are unknown.Recent comparative analyses revealed the presence of at least 15 distinct chloroplast protein kinases, suggesting an intricate kinase phosphorylation network in the chloroplast (Bayer et al., 2012). Generally, the phosphorylation of proteins is one of the most abundant posttranslational modifications. In complex eukaryotic systems, protein phosphorylation occurs most frequently on Ser followed by Thr residues, whereas protein phosphorylation of Tyr residues (1,800:200:1) is comparatively rare (Hunter, 1998; Mann et al., 2002). Protein phosphorylation is a general phenomenon in vivo; it is assumed that about one-third of all proteins are phosphorylated at a given time (Cohen, 2000; Ahn and Resing, 2001; Venter et al., 2001; Manning et al., 2002; Knight et al., 2003). A recent large-scale quantitative evaluation of human proteomic data strengthened the importance of protein phosphorylation for cellular function and human biology (Wilhelm et al., 2014). The C. reinhardtii and Arabidopsis genomes encode large kinase families (Arabidopsis Genome Initiative, 2000; Kerk et al., 2002; Merchant et al., 2007), supporting the view that protein phosphorylation also plays an important role in a plant’s life cycle. It is thus evident that the understanding of protein phosphorylation, including the specificity of residues phosphorylated or dephosphorylated in response to cellular as well as environmental factors, is one key to understanding the complex functional biological networks at the whole-system level. Likewise, it is crucial to design experimental setups allowing the linkage between phosphorylation events and particular physiological consequences to be elucidated.In this regard, we designed experiments to investigate STT7 kinase-dependent phosphorylation dynamics in C. reinhardtii in response to high light and anoxia, employing quantitative proteomics in conjunction with in-depth physiological characterization. These conditions are particularly interesting, as high light conditions are known to fully induce LHCSR3 protein expression and qE, while anoxia promotes CEF activity. Recently, it was demonstrated that qE and CEF are complementary and crucial in acclimation to these environmental cues (Kukuczka et al., 2014). Notably, LHCSR3 phosphorylation was suggested to depend on STT7 function (Bonente et al., 2011), while CEF supercomplex formation was found to be independent of STT7 kinase function (Takahashi et al., 2013), indicating that STT7 function might impact the acclimation to high light and anoxia in different ways. However, our quantitative proteomics and physiological data reveal that STT7-dependent variations in protein phosphorylation profiles have similar dramatic phenotypic consequences in both conditions, strongly suggesting that the regulation of protein phosphorylation is critical for driving successful acclimation to high light and anoxic growth environments.  相似文献   

15.
Ca2+-dependent protein kinases (CPKs) form a large family of 34 genes in Arabidopsis (Arabidopsis thaliana). Based on their dependence on Ca2+, CPKs can be sorted into three types: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially calcium-insensitive CPKs. Here, we report on the third type of CPK, CPK13, which is expressed in guard cells but whose role is still unknown. We confirm the expression of CPK13 in Arabidopsis guard cells, and we show that its overexpression inhibits light-induced stomatal opening. We combine several approaches to identify a guard cell-expressed target. We provide evidence that CPK13 (1) specifically phosphorylates peptide arrays featuring Arabidopsis K+ Channel KAT2 and KAT1 polypeptides, (2) inhibits KAT2 and/or KAT1 when expressed in Xenopus laevis oocytes, and (3) closely interacts in plant cells with KAT2 channels (Förster resonance energy transfer-fluorescence lifetime imaging microscopy). We propose that CPK13 reduces stomatal aperture through its inhibition of the guard cell-expressed KAT2 and KAT1 channels.Stomata are microscopic organs at the leaf surface, each made of two so-called guard cells forming a pore. Opening or closing these pores is the way through which plants control their gas exchanges with the atmosphere (i.e. carbon dioxide uptake to feed the photosynthetic process and transpirational loss of water vapor). Stomatal movements result from osmotically driven fluxes of water, which follow massive exchanges of solutes, including K+ ions, between the guard cells and the surrounding tissues (Hetherington, 2001; Nilson and Assmann, 2007).Both Ca2+-dependent and Ca2+-independent signaling pathways are known to control stomatal movements (MacRobbie, 1993, 1998; Blatt, 2000; Webb et al., 2001; Mustilli et al., 2002; Israelsson et al., 2006; Marten et al., 2007; Laanemets et al., 2013). In particular, Ca2+ signals have been reported to promote stomatal closure through the inhibition of inward K+ channels and the activation of anion channels (Blatt, 1991, 1992, 2000; Thiel et al., 1992; Grabov and Blatt, 1999; Schroeder et al., 2001; Hetherington and Brownlee, 2004; Mori et al., 2006; Marten et al., 2007; Geiger et al., 2010; Brandt et al., 2012; Scherzer et al., 2012). However, little is known about the molecular identity of the links between Ca2+ events and Shaker K+ channel activity. Several kinases and phosphatases are believed to be involved in both the Ca2+-dependent and Ca2+-independent signaling pathways. Plants express two large kinase families whose activity is related to Ca2+ signaling. Firstly, CBL-interacting protein kinases (CIPKs; 25 genes in Arabidopsis [Arabidopsis thaliana]) are indirectly controlled by their interaction with a set of calcium sensors, the calcineurin B-like proteins (CBLs; 10 genes in Arabidopsis). This complex forms a fascinating network of potential Ca2+ signaling decoders (Luan, 2009; Weinl and Kudla, 2009), which have been addressed in numerous reports (Xu et al., 2006; Hu et al., 2009; Batistic et al., 2010; Held et al., 2011; Chen et al., 2013). In particular, some CBL-CIPK pairs have been shown to regulate Shaker channels such as Arabidopsis K+ Transporter1 (AKT1; Xu et al., 2006; Lan et al., 2011) or AKT2 (Held et al., 2011). Second, Ca2+-dependent protein kinases (CPKs) form an even larger family (34 genes in Arabidopsis) of proteins combining a kinase domain with the ability to bind Ca2+, thanks to the so-called EF hands (Harmon et al., 2000; Harper et al., 2004). CPKs, which, interestingly, are not found in animal cells, exhibit different calcium dependencies (Boudsocq et al., 2012). With respect to this, three types of CPKs can be considered: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially Ca2+-insensitive CPKs (however, structurally close to kinases of groups 1 and 2).Pioneering work by Luan et al. (1993) demonstrated in Vicia faba guard cells that inward K+ channels were regulated by some Ca2+-dependent kinases. Then, such a Ca2+-dependent kinase was purified from guard cell protoplasts of V. faba and shown to actually phosphorylate the in vitro-translated KAT1 protein, a Shaker channel subunit natively expressed in Arabidopsis guard cells (Li et al., 1998). KAT1 regulation by CPK was shown by the inhibition of KAT1 currents after the coexpression of KAT1 and CDPK from soybean (Glycine max) in oocytes (Berkowitz et al., 2000). Since then, several cpk mutant lines of Arabidopsis have been shown to be impaired in stomatal movements, for example cpk10 (Ca2+ insensitive), cpk4/cpk11 (Ca2+ dependent), and cpk3/cpk6/cpk23 (Ca2+ dependent; Mori et al., 2006; Geiger et al., 2010; Munemasa et al., 2011; Hubbard et al., 2012).Of the nine genes encoding voltage-dependent K+ channels (Shaker) in Arabidopsis (Véry and Sentenac, 2002, 2003; Lebaudy et al., 2007; Hedrich, 2012), six are expressed in guard cells and play a role in stomatal movements: the Gated Outwardly-Rectifying K+ (GORK) gene, encoding an outward K+ channel subunit, and the AKT1, AKT2, Arabidopsis K+ Rectifying Channel1 (AtKC1), KAT1, and KAT2 genes, encoding inward K+ channel subunits (Pilot et al., 2001; Szyroki et al., 2001; Hosy et al., 2003; Pandey et al., 2007; Lebaudy et al., 2008a). Shaker channels result from the assembly of four subunits, and it has been shown that inward subunits tend to heterotetramerize, thus potentially widening the functional and regulatory scope of inward K+ conductance in guard cells (Xicluna et al., 2007; Jeanguenin et al., 2008; Lebaudy et al., 2008a, 2010). Inhibition of inward K+ channels has been shown to reduce stomatal opening (Liu et al., 2000; Kwak et al., 2001). This has grounded a strategy for disrupting inward K+ channel conductance in guard cells by expressing a nonfunctional KAT2 subunit (dominant negative mutation) in a kat2 knockout Arabidopsis line. The resulting Arabidopsis lines, named kincless, have no functional inward K+ channels and exhibit delayed stomatal opening (Lebaudy et al., 2008b) with, in the long term, a biomass reduction compared with the Arabidopsis wild-type line.Among the CPKs presumably expressed in Arabidopsis guard cells (Leonhardt et al., 2004), we looked for CPK13, which belongs to the atypical Ca2+-insensitive type of CPKs (Kanchiswamy et al., 2010; Boudsocq et al., 2012; Liese and Romeis, 2013) and whose role remains unknown in stomatal movements. Here, we confirm first that CPK13 kinase activity is independent of Ca2+ and show that CPK13 expression is predominant in Arabidopsis guard cells using CPK13-GUS lines. We then report that overexpression of CPK13 in Arabidopsis induces a dramatic default in stomatal aperture. Based on the previously reported kincless phenotype (Lebaudy et al., 2008b), we propose that CPK13 could reduce the activity of inward K+ channels in guard cells, particularly that of KAT2. We confirm this hypothesis by voltage-clamp experiments and show an inhibition of KAT2 and KAT1 activity by CPK13 (but not that of AKT2). In addition, we present peptide array phosphorylation assays showing that CPK13 targets, with some specificity, several KAT2 and KAT1 polypeptides. Finally, we demonstrate that KAT2 and CPK13 interact in planta using Förster resonance energy transfer (FRET)-fluorescence lifetime imaging microscopy (FLIM).  相似文献   

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

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