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.
3.
The salt stress-induced SALT-OVERLY-SENSITIVE (SOS) pathway in Arabidopsis (Arabidopsis thaliana) involves the perception of a calcium signal by the SOS3 and SOS3-like CALCIUM-BINDING PROTEIN8 (SCaBP8) calcium sensors, which then interact with and activate the SOS2 protein kinase, forming a complex at the plasma membrane that activates the SOS1 Na+/H+ exchanger. It has recently been reported that phosphorylation of SCaBP proteins by SOS2-like protein kinases (PKSs) stabilizes the interaction between the two proteins as part of a regulatory mechanism that was thought to be common to all SCaBP and PKS proteins. Here, we report the calcium-independent activation of PKS24 by SCaBP1 and show that activation is dependent on interaction of PKS24 with the C-terminal tail of SCaBP1. However, unlike what has been found for other PKS-SCaBP pairs, multiple amino acids in SCaBP1 are phosphorylated by PKS24, and this phosphorylation is dependent on the interaction of the proteins through the PKS24 FISL motif and on the efficient activation of PKS24 by the C-terminal tail of SCaBP1. In addition, we show that Thr-211 and Thr-212, which are not common phosphorylation sites in the conserved PFPF motif found in most SCaBP proteins, are important for this activation. Finally, we also found that SCaBP1-regulated PKS24 kinase activity is important for inactivating the Arabidopsis plasma membrane proton-translocating adenosine triphosphatase. Together, these results suggest the existence of a novel SCaBP-PKS regulatory mechanism in plants.Calcium is a ubiquitous second messenger that plays an important role in the regulation of plant growth and development. Many different types of calcium-binding proteins have been identified in plants (Harper et al., 2004), including the SALT-OVERLY-SENSITIVE3 (SOS3)-LIKE CALCIUM BINDING PROTEINS (SCaBPs; Liu and Zhu, 1998; Gong et al., 2004). Because the calcium-binding domain of these proteins shares sequence similarity with the yeast calcineurin B subunit, they have also been called CALCINEURIN B-LIKE PROTEINS (CBLs; Kudla et al., 1999; Luan et al., 2002). The founding member of this gene family, SOS3, was identified in a genetic screen from a salt-sensitive Arabidopsis (Arabidopsis thaliana) mutant (Liu and Zhu, 1998). SCaBP/CBL proteins interact with the SOS2-LIKE PROTEIN KINASES (PKSs)/CBL-INTERACTING PROTEIN KINASES (CIPKs; Shi et al., 1999; Halfter et al., 2000; Guo et al., 2001). The genetic linkage between these two families was established after identification of SOS2 from a genetic screen similar to the one that identified the sos3 mutant (Liu et al., 2000). SOS3 interacts with SOS2 in vivo and in vitro and activates SOS2 in a calcium-dependent manner in vitro (Halfter et al., 2000). The SOS3-SOS2 complex further activates SOS1, a plasma membrane (PM) Na+/H+ antiporter, by directly phosphorylating the SOS1 C terminus (Shi et al., 2000; Qiu et al., 2002; Quintero et al., 2002, 2011; Yu et al., 2010).In addition to the calcium-dependent activation of PKSs by SCaBP calcium sensors, two other regulatory mechanisms have been identified for these protein families. First, PKSs have a conserved 21-amino acid peptide (FISL motif) in their regulatory domain that is necessary for efficient interaction with the SCaBP calcium sensors (Guo et al., 2001; Albrecht et al., 2001; Gong et al., 2004). The PKS regulatory domain interacts with its kinase domain via the FISL motif to repress PKS activity; interaction of SCaBP with the PKS FISL motif releases the kinase domain inhibition allowing for kinase activity (Guo et al., 2001; Gong et al., 2004). Second, the PKSs phosphorylate a Ser residue in the conserved C-terminal PFPF motif of the SCaBP proteins. This phosphorylation enhances the interaction between the two proteins and fully activates the complex (Lin et al., 2009; Du et al., 2011; Hashimoto et al., 2012).In this study, we identified a novel PKS activation mechanism involving the calcium-independent activation of PKS24 by SCaBP1 and show that it requires binding of SCaBP1 to the FISL motif of PKS24 and the involvement of two Thr residues in the SCaBP1 C-terminal tail.  相似文献   

4.
The role of calcium-mediated signaling has been extensively studied in plant responses to abiotic stress signals. Calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs) constitute a complex signaling network acting in diverse plant stress responses. Osmotic stress imposed by soil salinity and drought is a major abiotic stress that impedes plant growth and development and involves calcium-signaling processes. In this study, we report the functional analysis of CIPK21, an Arabidopsis (Arabidopsis thaliana) CBL-interacting protein kinase, ubiquitously expressed in plant tissues and up-regulated under multiple abiotic stress conditions. The growth of a loss-of-function mutant of CIPK21, cipk21, was hypersensitive to high salt and osmotic stress conditions. The calcium sensors CBL2 and CBL3 were found to physically interact with CIPK21 and target this kinase to the tonoplast. Moreover, preferential localization of CIPK21 to the tonoplast was detected under salt stress condition when coexpressed with CBL2 or CBL3. These findings suggest that CIPK21 mediates responses to salt stress condition in Arabidopsis, at least in part, by regulating ion and water homeostasis across the vacuolar membranes.Drought and salinity cause osmotic stress in plants and severely affect crop productivity throughout the world. Plants respond to osmotic stress by changing a number of cellular processes (Xiong et al., 1999; Xiong and Zhu, 2002; Bartels and Sunkar, 2005; Boudsocq and Lauriére, 2005). Some of these changes include activation of stress-responsive genes, regulation of membrane transport at both plasma membrane (PM) and vacuolar membrane (tonoplast) to maintain water and ionic homeostasis, and metabolic changes to produce compatible osmolytes such as Pro (Stewart and Lee, 1974; Krasensky and Jonak, 2012). It has been well established that a specific calcium (Ca2+) signature is generated in response to a particular environmental stimulus (Trewavas and Malhó, 1998; Scrase-Field and Knight, 2003; Luan, 2009; Kudla et al., 2010). The Ca2+ changes are primarily perceived by several Ca2+ sensors such as calmodulin (Reddy, 2001; Luan et al., 2002), Ca2+-dependent protein kinases (Harper and Harmon, 2005), calcineurin B-like proteins (CBLs; Luan et al., 2002; Batistič and Kudla, 2004; Pandey, 2008; Luan, 2009; Sanyal et al., 2015), and other Ca2+-binding proteins (Reddy, 2001; Shao et al., 2008) to initiate various cellular responses.Plant CBL-type Ca2+ sensors interact with and activate CBL-interacting protein kinases (CIPKs) that phosphorylate downstream components to transduce Ca2+ signals (Liu et al., 2000; Luan et al., 2002; Batistič and Kudla, 2004; Luan, 2009). In several plant species, multiple members have been identified in the CBL and CIPK family (Luan et al., 2002; Kolukisaoglu et al., 2004; Pandey, 2008; Batistič and Kudla, 2009; Weinl and Kudla, 2009; Pandey et al., 2014). Involvement of specific CBL-CIPK pair to decode a particular type of signal entails the alternative and selective complex formation leading to stimulus-response coupling (D’Angelo et al., 2006; Batistič et al., 2010).Several CBL and CIPK family members have been implicated in plant responses to drought, salinity, and osmotic stress based on genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants (Zhu, 2002; Cheong et al., 2003, 2007; Kim et al., 2003; Pandey et al., 2004, 2008; D’Angelo et al., 2006; Qin et al., 2008; Tripathi et al., 2009; Held et al., 2011; Tang et al., 2012; Drerup et al., 2013; Eckert et al., 2014). A few CIPKs have also been functionally characterized by gain-of-function approach in crop plants such as rice (Oryza sativa), pea (Pisum sativum), and maize (Zea mays) and were found to be involved in osmotic stress responses (Mahajan et al., 2006; Xiang et al., 2007; Yang et al., 2008; Tripathi et al., 2009; Zhao et al., 2009; Cuéllar et al., 2010).In this report, we examined the role of the Arabidopsis CIPK21 gene in osmotic stress response by reverse genetic analysis. The loss-of-function mutant plants became hypersensitive to salt and mannitol stress conditions, suggesting that CIPK21 is involved in the regulation of osmotic stress response in Arabidopsis. These findings are further supported by an enhanced tonoplast targeting of the cytoplasmic CIPK21 through interaction with the vacuolar Ca2+ sensors CBL2 and CBL3 under salt stress condition.  相似文献   

5.
In epidermal and mesophyll cells of Arabidopsis (Arabidopsis thaliana) leaves, nuclei become relocated in response to strong blue light. We previously reported that nuclear positions both in darkness and in strong blue light are regulated by the blue light receptor phototropin2 in mesophyll cells. Here, we investigate the involvement of phototropin and the actin cytoskeleton in nuclear positioning in epidermal cells. Analysis of geometrical parameters revealed that, in darkness, nuclei were distributed near the center of the cell, adjacent to the inner periclinal wall, independent of cell shape. Dividing the anticlinal wall into concave, convex, and intermediate regions indicated that, in strong blue light, nuclei became relocated preferably to a concave region of the anticlinal wall, nearest the center of the cell. Mutant analyses verified that light-dependent nuclear positioning was regulated by phototropin2, while dark positioning of nuclei was independent of phototropin. Nuclear movement was inhibited by an actin-depolymerizing reagent, latrunculin B, but not by a microtubule-disrupting reagent, propyzamide. Imaging actin organization by immunofluorescence microscopy revealed that thick actin bundles, periclinally arranged parallel to the longest axis of the epidermal cell, were associated with the nucleus in darkness, whereas under strong blue light, the actin bundles, especially in the vicinity of the nucleus, became arranged close to the anticlinal walls. Light-dependent changes in the actin organization were clear in phot1 mutant but not in phot2 and phot1phot2 mutants. We propose that, in Arabidopsis, blue-light-dependent nuclear positioning is regulated by phototropin2-dependent reorganization of the actin cytoskeleton.Positioning organelles is essential for cellular activities. The nucleus changes its position in a programmatic way during development and the cell cycle (Britz, 1979; Nagai, 1993; Chytilova et al., 2000). For example, before asymmetrical divisions that give rise to the formation of root hair cells or guard mother cells, the nucleus migrates to the future division plane (Britz, 1979). In elongating root hair cells of Arabidopsis (Arabidopsis thaliana), the nucleus is maintained at a fixed distance from the apex (Ketelaar et al., 2002).While the nuclear migrations before mitosis and in root hairs are developmental, nuclear positioning is also regulated environmentally. In the fern, Adiantum capillus-veneris, nuclei in prothallial cells change their intracellular positions in response to light (Kagawa and Wada, 1993, 1995). The nuclei are located along the anticlinal walls in darkness and move toward the outer periclinal walls in weak light and to the anticlinal walls in strong light (Kagawa and Wada, 1993, 1995; Tsuboi et al., 2007). This response is called light-dependent nuclear positioning. Since the response is induced in cells that exhibit neither cell division nor expansion, it is believed to have a physiological role, distinct from the nuclear positioning associated with development.Recently, light-dependent nuclear positioning was reported in the spermatophyte Arabidopsis (Iwabuchi et al., 2007). In epidermal and mesophyll cells of dark-treated leaves, nuclei are distributed along the inner periclinal wall. Under strong light, they become located along the anticlinal walls. In mesophyll cells, nuclear movement from inner periclinal to anticlinal walls is induced repeatedly and specifically by blue light of high-fluence rate (more than 50 μ mol m−2 s−1) and is regulated by the blue light receptor phototropin2. Interestingly, mesophyll cells of the phot2 mutant have aberrantly positioned nuclei even in darkness. By contrast, the involvement of phototropins in nuclear positioning has not yet been examined for epidermal cells.Phototropin is a blue light receptor containing two light oxygen voltage domains at the N terminus, which bind an FMN chromophore, and a Ser/Thr kinase domain at the C terminus, which undergoes blue-light-dependent autophosphorylation (Briggs et al., 2001a; Christie, 2007). Arabidopsis possesses phototropins1 and 2 (Huala et al., 1997; Jarillo et al., 2001; Kagawa et al., 2001; Sakai et al., 2001). Phototropins are shown microscopically and biochemically to localize to the plasma membrane region (Briggs et al., 2001b; Sakamoto and Briggs, 2002; Kong et al., 2006) and mediate several responses, including phototropism (Liscum and Briggs, 1995; Sakai et al., 2001), stomatal opening (Kinoshita et al., 2001), and chloroplast movements (Jarillo et al., 2001; Kagawa et al., 2001; Sakai et al., 2001). In general, phototropin1 is more sensitive to light than its paralog and mediates low-fluence-rate light responses, whereas phototropin2 functions predominantly under higher fluence rates (Sakai et al., 2001).While the photoreceptor eliciting these nuclear movements has been revealed, the motile system responsible for moving the nuclei is still unknown. In general, organelle movements depend on the cytoskeleton, with the specific roles for actin and microtubules dependent on the organelle and species (Wada and Suetsugu, 2004). In land plants, the actin cytoskeleton plays a pivotal role in positioning organelles, including nuclei, chloroplasts, mitochondria, and peroxisomes (Wada and Suetsugu, 2004; Takagi et al., 2009).The role of the cytoskeleton in developmental nuclear movements has been investigated. In growing root hairs of Arabidopsis, the nuclear movements are driven along actin filaments (Ketelaar et al., 2002), whereas, in tobacco (Nicotiana tabacum) BY-2 cells, the cell-cycle-based nuclear migration before mitosis is found to depend on microtubules (Katsuta et al., 1990). In interphase Spirogyra crassa cells, centering of nuclei is regulated by both actin filaments and microtubules, but in distinct ways (Grolig, 1998). To the best of our knowledge, the cytoskeletal basis of environmentally induced nuclear movements in land plants has not been elucidated.The best-characterized organelle movements are the light-induced orientation movements of chloroplasts, and although exceptions have been reported, this movement depends on actin (Britz, 1979; Takagi, 2003; Wada et al., 2003). Under weak light, chloroplasts gather at the periclinal walls, perpendicular to the direction of light (accumulation response), whereas under strong light, they become positioned along the anticlinal walls, parallel to the direction of light (avoidance response). Recently, for Arabidopsis, Kadota et al. (2009) characterized the nature of the actin filaments probably involved in these movements. With the onset of either accumulation or avoidance response, short actin filaments appear at the leading edge of each chloroplast.In Arabidopsis, light-dependent nuclear positioning shows similarities to the chloroplast avoidance response, with regard to the direction of movement, relevant photoreceptor (phototropin2), and effective fluence rate (Iwabuchi and Takagi, 2008). On the other hand, nuclei are larger than chloroplasts and might require thicker, more rigid actin bundles for effective motility. Here, we investigate the involvement of the actin cytoskeleton as well as phototropin in regulatory system for nuclear positioning in epidermal cells of Arabidopsis leaves.  相似文献   

6.
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   

7.
Blue-light-induced chloroplast photorelocation movement is observed in most land plants. Chloroplasts move toward weak-light-irradiated areas to efficiently absorb light (the accumulation response) and escape from strong-light-irradiated areas to avoid photodamage (the avoidance response). The plant-specific kinase phototropin (phot) is the blue-light receptor for chloroplast movements. Although the molecular mechanisms for chloroplast photorelocation movement have been analyzed, the overall aspects of signal transduction common to land plants are still unknown. Here, we show that the liverwort Marchantia polymorpha exhibits the accumulation and avoidance responses exclusively induced by blue light as well as specific chloroplast positioning in the dark. Moreover, in silico and Southern-blot analyses revealed that the M. polymorpha genome encodes a single PHOT gene, MpPHOT, and its knockout line displayed none of the chloroplast photorelocation movements, indicating that the sole MpPHOT gene mediates all types of movement. Mpphot was localized on the plasma membrane and exhibited blue-light-dependent autophosphorylation both in vitro and in vivo. Heterologous expression of MpPHOT rescued the defects in chloroplast movement of phot mutants in the fern Adiantum capillus-veneris and the seed plant Arabidopsis (Arabidopsis thaliana). These results indicate that Mpphot possesses evolutionarily conserved regulatory activities for chloroplast photorelocation movement. M. polymorpha offers a simple and versatile platform for analyzing the fundamental processes of phototropin-mediated chloroplast photorelocation movement common to land plants.Light is not only an energy source for photosynthesis but it is also a signal that regulates numerous physiological responses for plants. Because chloroplasts are the important organelle for photosynthesis, most plant species possess a light-dependent mechanism to regulate the intracellular position of chloroplasts (chloroplast photorelocation movement). Intensive studies on chloroplast photorelocation movement have been performed since the 19th century (Böhm, 1856). Senn (1908) described the chloroplast distribution patterns under different light conditions in various plant species, including algae, liverworts, mosses, ferns, and seed plants, and revealed the general responses of chloroplasts to intensity and direction of light. Under low-light conditions, chloroplasts are positioned along the cell walls perpendicular to the direction of incident light (i.e. periclinal cell walls) to efficiently capture light for photosynthesis (the accumulation response). By contrast, under high-light conditions, chloroplasts are stacked along the cell walls parallel to the direction of incident light (i.e. anticlinal cell walls) to minimize total light absorption and to avoid photooxidative damage (the avoidance response). These chloroplast movements are induced primarily by blue light in most plant species (Suetsugu and Wada, 2007a). In some plant species, such as several ferns including Adiantum capillus-veneris, the moss Physcomitrella patens, and some charophycean green algae (Mougeotia scalaris and Mesotaenium caldariorum), red light is also effective to induce chloroplast movement (Suetsugu and Wada, 2007b). Analyses of chloroplast movement in response to irradiation with polarized light and/or a microbeam suggest that the photoreceptor for chloroplast movement is localized on or close to the plasma membrane (Haupt and Scheuerlein, 1990; Wada et al., 1993). In addition, chloroplasts assume their specific positions in the dark (dark positioning), although the patterns vary among plant species (Senn, 1908). For example, the chloroplasts are localized at the bottom of the cell in palisade cells of Arabidopsis (Arabidopsis thaliana; Suetsugu et al., 2005a) and on the anticlinal walls bordering neighboring cells in the prothallial cells of A. capillus-veneris (Kagawa and Wada, 1993; Tsuboi et al., 2007).Molecular mechanisms for chloroplast photorelocation movements have been revealed through molecular genetic analyses using Arabidopsis (Suetsugu and Wada, 2012). The light-activated kinase phototropin was identified as the blue-light receptor (Jarillo et al., 2001; Kagawa et al., 2001; Sakai et al., 2001). Phototropin consists of two functional regions: a photosensory domain at the N terminus and a Ser/Thr kinase domain at the C terminus (Christie, 2007). The N-terminal photosensory domain contains two light, oxygen, or voltage (LOV) domains, which belong to the Per/ARNT/Sim domain superfamily. Each LOV domain binds to one FMN and functions as a blue-light sensor (Christie et al., 1999). The LOV2 domain is essential for blue-light-dependent regulation of the activation of the C-terminal kinase domain (Christie et al., 2002; Harper et al., 2003).Arabidopsis has two phototropins: phot1 and phot2 (Christie, 2007). Besides chloroplast photorelocation movement, phototropin controls other photoresponses to optimize the photosynthetic efficiency in plants and improves growth responses such as phototropism, stomatal opening, and leaf flattening (Christie, 2007). Both phot1 and phot2 redundantly regulate the chloroplast accumulation response (Sakai et al., 2001), hypocotyl phototropism (Huala et al., 1997; Sakai et al., 2001), stomatal opening (Kinoshita et al., 2001), and leaf flattening (Sakai et al., 2001; Sakamoto and Briggs, 2002). Rapid inhibition of hypocotyl elongation is specifically mediated by phot1 (Folta and Spalding, 2001), whereas the chloroplast avoidance response (Jarillo et al., 2001; Kagawa et al., 2001) and palisade cell development (Kozuka et al., 2011) are mediated primarily by phot2.It is thought that the phototropin-regulated photoresponses are mediated by mechanisms in which gene expression is not involved primarily. For example, chloroplast photorelocation movement can be observed even in enucleated fern cells (Wada, 1988), and phototropins show only a minor contribution to blue-light-induced gene expression in Arabidopsis (Jiao et al., 2003; Ohgishi et al., 2004; Lehmann et al., 2011). Furthermore, both phot1 and phot2 are localized on the plasma membrane despite the absence of a transmembrane domain (Sakamoto and Briggs, 2002; Kong et al., 2006). During chloroplast movement, phototropins, in particular phot2, associate not only with the plasma membrane but also with the chloroplast outer membrane (Kong et al., 2013b). In addition, phot1 shows blue-light-dependent internalization into the cytoplasm (Sakamoto and Briggs, 2002; Knieb et al., 2004; Wan et al., 2008; Kaiserli et al., 2009), whereas phot2 exhibits a blue-light-dependent association with the Golgi apparatus (Kong et al., 2006).PHOT genes have been identified from various green plants and are indicated to be duplicated in respective lineages such as seed plants, ferns, lycophytes, and mosses (Li et al., 2014). In the fern A. capillus-veneris, chloroplast accumulation and avoidance responses are induced by both blue and red light (Yatsuhashi et al., 1985). This fern has three phototropin family proteins, two phototropins (Acphot1 and Acphot2; Kagawa et al., 2004), and one neochrome that possesses the chromophore-binding domain of phytochrome and complete phototropin domains (Nozue et al., 1998). Neochrome is the red-light receptor that mediates chloroplast movement (Kawai et al., 2003) and possibly blue-light-induced chloroplast movement through its LOV domains (Kanegae et al., 2006). Because the Acphot2 mutant is defective in the chloroplast avoidance response and dark positioning (Kagawa et al., 2004; Tsuboi et al., 2007), similar to the phot2 mutant in Arabidopsis (Jarillo et al., 2001; Kagawa et al., 2001; Suetsugu et al., 2005a), the function of phot2 in the regulation of chloroplast movement is highly conserved in these vascular plants. In the moss P. patens, in which chloroplast accumulation and avoidance responses are induced by both blue and red light (Kadota et al., 2000), seven phototropin genes are present in the draft genome sequences (Rensing et al., 2008). The phototropins encoded by four of these genes (PpphotA1, PpphotA2, PpphotB1, and PpphotB2) function in the blue-light-induced chloroplast movement (Kasahara et al., 2004). Moreover, red-light-induced chloroplast movements are mediated by both conventional phytochromes (Mittmann et al., 2004; Uenaka and Kadota, 2007) and phototropins (Kasahara et al., 2004). Because the direct association between phytochromes and phototropins is suggested to be involved in red-light-induced chloroplast movement (Jaedicke et al., 2012), phototropins should be essential components in the chloroplast movement signaling pathway (Kasahara et al., 2004).A single PHOT gene was isolated in a unicellular green alga, Chlamydomonas reinhardtii (Huang et al., 2002; Kasahara et al., 2002). When expressed in Arabidopsis phot1 phot2 double-mutant plants, C. reinhardtii phototropin rescued the defects in chloroplast photorelocation movement in phot1 phot2 plants (Onodera et al., 2005), indicating that the initial step of the phototropin-mediated signal transduction mechanism for chloroplast movements is conserved in the green plant lineage. Although the existence of only one PHOT gene is ideal for elucidation of phototropin-mediated responses, C. reinhardtii cells contain a single chloroplast and show no chloroplast photorelocation movement.Liverworts represent the most basal lineage of extant land plants and offer a valuable experimental system for elucidation of various physiological responses commonly seen in land plants (Bowman et al., 2007). Marchantia polymorpha has emerged as a model liverwort because molecular biological techniques, such as genetic transformation and gene-targeting technologies, have been established for the species (Ishizaki et al., 2008, 2013a; Kubota et al., 2013; Sugano et al., 2014). Furthermore, an ongoing M. polymorpha genome sequencing project under the Community Sequencing Program at the Joint Genome Institute has indicated that many biological mechanisms found in other groups of land plants are conserved in a much less complex form. Blue-light-induced chloroplast movement was briefly reported in M. polymorpha (Senn, 1908; Nakazato et al., 1999). However, information on chloroplast photorelocation movement in liverworts, including M. polymorpha, is very limited.In this study, we investigated chloroplast photorelocation movement in detail in M. polymorpha and analyzed the molecular mechanism underlying the photoreceptor system through molecular genetic analysis of M. polymorpha phototropin.  相似文献   

8.
9.
10.
11.
12.
Ca2+ and nitric oxide (NO) are essential components involved in plant senescence signaling cascades. In other signaling pathways, NO generation can be dependent on cytosolic Ca2+. The Arabidopsis (Arabidopsis thaliana) mutant dnd1 lacks a plasma membrane-localized cation channel (CNGC2). We recently demonstrated that this channel affects plant response to pathogens through a signaling cascade involving Ca2+ modulation of NO generation; the pathogen response phenotype of dnd1 can be complemented by application of a NO donor. At present, the interrelationship between Ca2+ and NO generation in plant cells during leaf senescence remains unclear. Here, we use dnd1 plants to present genetic evidence consistent with the hypothesis that Ca2+ uptake and NO production play pivotal roles in plant leaf senescence. Leaf Ca2+ accumulation is reduced in dnd1 leaves compared to the wild type. Early senescence-associated phenotypes (such as loss of chlorophyll, expression level of senescence-associated genes, H2O2 generation, lipid peroxidation, tissue necrosis, and increased salicylic acid levels) were more prominent in dnd1 leaves compared to the wild type. Application of a Ca2+ channel blocker hastened senescence of detached wild-type leaves maintained in the dark, increasing the rate of chlorophyll loss, expression of a senescence-associated gene, and lipid peroxidation. Pharmacological manipulation of Ca2+ signaling provides evidence consistent with genetic studies of the relationship between Ca2+ signaling and senescence with the dnd1 mutant. Basal levels of NO in dnd1 leaf tissue were lower than that in leaves of wild-type plants. Application of a NO donor effectively rescues many dnd1 senescence-related phenotypes. Our work demonstrates that the CNGC2 channel is involved in Ca2+ uptake during plant development beyond its role in pathogen defense response signaling. Work presented here suggests that this function of CNGC2 may impact downstream basal NO production in addition to its role (also linked to NO signaling) in pathogen defense responses and that this NO generation acts as a negative regulator during plant leaf senescence signaling.Senescence can be considered as the final stage of a plant’s development. During this process, nutrients will be reallocated from older to younger parts of the plant, such as developing leaves and seeds. Leaf senescence has been characterized as a type of programmed cell death (PCD; Gan and Amasino, 1997; Quirino et al., 2000; Lim et al., 2003). During senescence, organelles such as chloroplasts will break down first. Biochemical changes will also occur in the peroxisome during this process. When the chloroplast disassembles, it is easily observed as a loss of chlorophyll. Mitochondria, the source of energy for cells, will be the last cell organelles to undergo changes during the senescence process (Quirino et al., 2000). At the same time, other catabolic events (e.g. protein and lipid breakdown, etc.) are occurring (Quirino et al., 2000). Hormones may also contribute to this process (Gepstein, 2004). From this information we can infer that leaf senescence is regulated by many signals.Darkness treatment can induce senescence in detached leaves (Poovaiah and Leopold, 1973; Chou and Kao, 1992; Weaver and Amasino, 2001; Chrost et al., 2004; Guo and Crawford, 2005; Ülker et al., 2007). Ca2+ can delay the senescence of detached leaves (Poovaiah and Leopold, 1973) and leaf senescence induced by methyl jasmonate (Chou and Kao, 1992); the molecular events that mediate this effect of Ca2+ are not well characterized at present.Nitric oxide (NO) is a critical signaling molecule involved in many plant physiological processes. Recently, published evidence supports NO acting as a negative regulator during leaf senescence (Guo and Crawford, 2005; Mishina et al., 2007). Abolishing NO generation in either loss-of-function mutants (Guo and Crawford, 2005) or transgenic Arabidopsis (Arabidopsis thaliana) plants expressing NO degrading dioxygenase (NOD; Mishina et al., 2007) leads to an early senescence phenotype in these plants compared to the wild type. Corpas et al. (2004) showed that endogenous NO is mainly accumulated in vascular tissues of pea (Pisum sativum) leaves. This accumulation is significantly reduced in senescing leaves (Corpas et al., 2004). Corpas et al. (2004) also provided evidence that NO synthase (NOS)-like activity (i.e. generation of NO from l-Arg) is greatly reduced in senescing leaves. Plant NOS activity is regulated by Ca2+/calmodulin (CaM; Delledonne et al., 1998; Corpas et al., 2004, 2009; del Río et al., 2004; Valderrama et al., 2007; Ma et al., 2008). These studies suggest a link between Ca2+ and NO that could be operating during senescence.In animal cells, all three NOS isoforms require Ca2+/CaM as a cofactor (Nathan and Xie, 1994; Stuehr, 1999; Alderton et al., 2001). Notably, animal NOS contains a CaM binding domain (Stuehr, 1999). It is unclear whether Ca2+/CaM can directly modulate plant NOS or if Ca2+/CaM impacts plant leaf development/senescence through (either direct or indirect) effects on NO generation. However, recent studies from our lab suggest that Ca2+/CaM acts as an activator of NOS activity in plant innate immune response signaling (Ali et al., 2007; Ma et al., 2008).Although Arabidopsis NO ASSOCIATED PROTEIN1 (AtNOA1; formerly named AtNOS1) was thought to encode a NOS enzyme, no NOS-encoding gene has yet been identified in plants (Guo et al., 2003; Crawford et al., 2006; Zemojtel et al., 2006). However, the AtNOA1 loss-of-function mutant does display reduced levels of NO generation, and several groups have used the NO donor sodium nitroprusside (SNP) to reverse some low-NO related phenotypes in Atnoa1 plants (Guo et al., 2003; Bright et al., 2006; Zhao et al., 2007). Importantly, plant endogenous NO deficiency (Guo and Crawford, 2005; Mishina et al., 2007) or abscisic acid/methyl jasmonate (Hung and Kao, 2003, 2004) induced early senescence can be successfully rescued by application of exogenous NO. Addition of NO donor can delay GA-elicited PCD in barley (Hordeum vulgare) aleurone layers as well (Beligni et al., 2002).It has been suggested that salicylic acid (SA), a critical pathogen defense metabolite, can be increased in natural (Morris et al., 2000; Mishina et al., 2007) and transgenic NOD-induced senescent Arabidopsis leaves (Mishina et al., 2007). Pathogenesis related gene1 (PR1) expression is up-regulated in transgenic Arabidopsis expressing NOD (Mishina et al., 2007) and in leaves of an early senescence mutant (Ülker et al., 2007).Plant cyclic nucleotide gated channels (CNGCs) have been proposed as candidates to conduct extracellular Ca2+ into the cytosol (Sunkar et al., 2000; Talke et al., 2003; Lemtiri-Chlieh and Berkowitz, 2004; Ali et al., 2007; Demidchik and Maathuis, 2007; Frietsch et al., 2007; Kaplan et al., 2007; Ma and Berkowitz, 2007; Urquhart et al., 2007; Ma et al., 2009a, 2009b). Arabidopsis “defense, no death” (dnd1) mutant plants have a null mutation in the gene encoding the plasma membrane-localized Ca2+-conducting CNGC2 channel. This mutant also displays no hypersensitive response to infection by some pathogens (Clough et al., 2000; Ali et al., 2007). In addition to involvement in pathogen-mediated Ca2+ signaling, CNGC2 has been suggested to participate in the process of leaf development/senescence (Köhler et al., 2001). dnd1 mutant plants have high levels of SA and expression of PR1 (Yu et al., 1998), and spontaneous necrotic lesions appear conditionally in dnd1 leaves (Clough et al., 2000; Jirage et al., 2001). Endogenous H2O2 levels in dnd1 mutants are increased from wild-type levels (Mateo et al., 2006). Reactive oxygen species molecules, such as H2O2, are critical to the PCD/senescence processes of plants (Navabpour et al., 2003; Overmyer et al., 2003; Hung and Kao, 2004; Guo and Crawford, 2005; Zimmermann et al., 2006). Here, we use the dnd1 mutant to evaluate the relationship between leaf Ca2+ uptake during plant growth and leaf senescence. Our results identify NO, as affected by leaf Ca2+ level, to be an important negative regulator of leaf senescence initiation. Ca2+-mediated NO production during leaf development could control senescence-associated gene (SAG) expression and the production of molecules (such as SA and H2O2) that act as signals during the initiation of leaf senescence programs.  相似文献   

13.
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   

14.
In higher plants, blue light (BL) phototropism is primarily controlled by the phototropins, which are also involved in stomatal movement and chloroplast relocation. These photoresponses are mediated by two phototropins, phot1 and phot2. Phot1 mediates responses with higher sensitivity than phot2, and phot2 specifically mediates chloroplast avoidance and dark positioning responses. Here, we report the isolation and characterization of a Nonphototropic seedling1 (Nps1) mutant of tomato (Solanum lycopersicum). The mutant is impaired in low-fluence BL responses, including chloroplast accumulation and stomatal opening. Genetic analyses show that the mutant locus is dominant negative in nature. In dark-grown seedlings of the Nps1 mutant, phot1 protein accumulates at a highly reduced level relative to the wild type and lacks BL-induced autophosphorylation. The mutant harbors a single glycine-1484-to-alanine transition in the Hinge1 region of a phot1 homolog, resulting in an arginine-to-histidine substitution (R495H) in a highly conserved A′α helix proximal to the light-oxygen and voltage2 domain of the translated gene product. Significantly, the R495H substitution occurring in the Hinge1 region of PHOT1 abolishes its regulatory activity in Nps1 seedlings, thereby highlighting the functional significance of the A′α helix region in phototropic signaling of tomato.Being sessile in nature, plants have developed diverse sets of sensory mechanisms, integrating external cues such as light, water, and temperature to adapt their growth and development to the ambient environment. Plants have evolved a cohort of photoreceptors such as red/far-red light-sensing phytochromes (Chen and Chory, 2011), UV-A/blue light (BL)-sensing phototropins (Christie, 2007; Holland et al., 2009; Suetsugu and Wada, 2013), cryptochromes (Yu et al., 2010; Liu et al., 2011), Zeitlupe (ZTL)/Flavin-binding, Kelch repeat, F-box protein1/light-oxygen and voltage (LOV)-kelch protein2 members of the ZTL/ADAGIO putative family of photoreceptors (Suetsugu and Wada, 2013), and UV-B light-sensing UV-B resistance8 (Heijde and Ulm, 2012), enabling them to sense nearly the full range of the solar spectrum. One of the most visually obvious photoresponses of flowering plants involves the growth and orientation of organs toward or away from light, particularly during the early stages of growth and the establishment of seedlings (Iino, 1990) and during gap-filling situations in dense canopy conditions (Ballaré, 1999) for optimizing photosynthesis and interspecies/intraspecies competition. Several studies involving the relative effectiveness of different wavelengths of the solar spectrum as well as monitoring of lateral differences in light intensity revealed that the directional growth of plants is specifically mediated by the UV-A/blue region of the visible spectrum. Molecular genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants inhibited in hypocotyl curvature toward BL revealed that, among the UV-A light-/BL-specific photoreceptors, the phototropins perceive ambient light as a cue for directional growth (Liscum and Briggs, 1995; Kagawa et al., 2001).Phototropins have been identified in several plant species, ranging from the green alga Chlamydomonas reinhardtii to higher plants (Briggs et al., 2001). To date, two members of the phototropins have been reported from higher plants, phot1 and phot2, which share sequence homology (Sakai et al., 2001). Physiological analyses with Arabidopsis mutants lacking phot1 and phot2 have revealed that, in addition to regulating the hypocotyl curvature of seedlings toward BL (Huala et al., 1997; Christie et al., 1998), phototropins also regulate a diverse range of responses in flowering plants (Christie and Murphy, 2013; Hohm et al., 2013). These responses include chloroplast movements (Sakai et al., 2001), nuclear positioning (Iwabuchi et al., 2007), stomatal opening (Kinoshita et al., 2001), sun tracking (Inoue et al., 2008b), leaf expansion (Ohgishi et al., 2004), leaf movements (Inoue et al., 2005), leaf photomorphogenesis (Kozuka et al., 2011), leaf flattening (Sakamoto and Briggs, 2002), and the rapid inhibition of the growth of etiolated hypocotyls (Folta and Spalding, 2001).While both phot1 and phot2 overlap in function in regulating phototropism, chloroplast accumulation, leaf expansion, and stomatal opening, they also exhibit differential photosensitivity to BL, where phot1 is more sensitive to low-fluence BL than phot2. Both phot1 and phot2 redundantly regulate the chloroplast accumulation toward low-fluence BL, and phot2 exclusively regulates the chloroplast avoidance from high-fluence BL (Jarillo et al., 2001; Kagawa et al., 2001), while phot1 solely mediates the rapid inhibition of the elongation of etiolated hypocotyls (Folta and Spalding, 2001). Analysis of mutants downstream of blue light perception by phototropins revealed that the phototropin signaling branches out at an early step, and phot1 and phot2 trigger distinct photoresponses recruiting multiple signaling partners (Christie and Murphy, 2013; Hohm et al., 2013).Molecular characterizations have shown that phototropins are plasma membrane-associated Ser/Thr kinases containing a photosensory domain (Briggs and Christie, 2002) in the N-terminal region composed of two LOV domains (LOV1 and LOV2) and the kinase domain at the C-terminal end. The LOV1 and LOV2 domains bind the FMN as chromophore and are responsible for BL sensing by phototropin. Although phototropins characteristically possess two LOV domains, the photoregulation of phototropin activity is predominantly mediated by LOV2 (Christie, 2007). The exposure to BL also causes adduct formation between the FMN and the Cys residue in LOV domains and leads to the phosphorylation of phototropin, which is believed to be the primary step in the transmission of phototropic signals (Christie et al., 1998; Sakai et al., 2000). To decipher the functions of different domains of phototropins, many different substitution mutants of phototropins have been generated, which have enabled the elucidation of the functional significance of the different domains (Matsuoka and Tokutomi, 2005; Jones et al., 2007; Kong et al., 2007; Inoue et al., 2008a). Inoue et al. (2008a) showed that the BL-induced autophosphorylation of Ser-851 in the C-terminal kinase domain of phototropin is the primary step for initiating stomatal opening, phototropism, chloroplast accumulation, and leaf flattening. Mutational studies also revealed that the photosensory N-terminal domain of phototropin acts as a kinase inhibitor, where the LOV2 domain inhibits the activity of kinase domain by binding to it, and BL exposure is required for the dissociation of the LOV2 domain, enabling phosphorylation of the kinase domain (Matsuoka and Tokutomi, 2005; Jones et al., 2007).While our current understanding of phototropism has been greatly facilitated by the isolation of phototropins and their signaling mutants, the phot mutants identified to date are loss-of-function alleles. The lack of dominant-negative alleles or alleles with increased sensitivity to phototropic stimulus has hindered exploration into the roles of different domains of phot proteins in regulating phototropic signaling. In addition, the dearth of mutants defective in phototropin or phototropin-mediated responses has been a major bottleneck in furthering our understanding of the function of phototropins in crop species. Although phototropin homologs have been identified from a variety of crop species, including oat (Avena sativa; Zacherl et al., 1998), rice (Oryza sativa; Kanegae et al., 2000), and tomato (Solanum lycopersicum; Sharma et al., 2007; Sharma and Sharma, 2007), only the coleoptile phototropism1 mutant of rice has been isolated, which is defective in BL phototropism (Haga et al., 2005).Here, we report that in a mutant screen for nonphototropic seedlings under continuous BL, we recovered a strong dominant-negative mutation of phot1. The dominant-negative mutations are useful to elucidate redundant functions, as mutant protein in addition to suppressing its own functions can also suppress the function of its partners. The characterization of this new phot1 mutant revealed that the dominant activity is caused by the substitution of an Arg residue located in the A′α helix in the Hinge1 region between the LOV1 and LOV2 domains. Our study shows the functional importance of the A′α helix (Halavaty and Moffat, 2007) in regulating phot1-mediated signaling in tomato.  相似文献   

15.
16.
17.
18.
19.
Plant trichomes play important protective functions and may have a major influence on leaf surface wettability. With the aim of gaining insight into trichome structure, composition, and function in relation to water-plant surface interactions, we analyzed the adaxial and abaxial leaf surface of holm oak (Quercus ilex) as a model. By measuring the leaf water potential 24 h after the deposition of water drops onto abaxial and adaxial surfaces, evidence for water penetration through the upper leaf side was gained in young and mature leaves. The structure and chemical composition of the abaxial (always present) and adaxial (occurring only in young leaves) trichomes were analyzed by various microscopic and analytical procedures. The adaxial surfaces were wettable and had a high degree of water drop adhesion in contrast to the highly unwettable and water-repellent abaxial holm oak leaf sides. The surface free energy and solubility parameter decreased with leaf age, with higher values determined for the adaxial sides. All holm oak leaf trichomes were covered with a cuticle. The abaxial trichomes were composed of 8% soluble waxes, 49% cutin, and 43% polysaccharides. For the adaxial side, it is concluded that trichomes and the scars after trichome shedding contribute to water uptake, while the abaxial leaf side is highly hydrophobic due to its high degree of pubescence and different trichome structure, composition, and density. Results are interpreted in terms of water-plant surface interactions, plant surface physical chemistry, and plant ecophysiology.Plant surfaces have an important protecting function against multiple biotic and abiotic stress factors (Riederer, 2006). They may, for example, limit the attack of insects (Eigenbrode and Jetter, 2002) or pathogenic fungi (Gniwotta et al., 2005; Łaźniewska et al., 2012), avoid damage caused by high intensities of UV and visible radiation (Reicosky and Hanover, 1978; Karabourniotis and Bormann, 1999), help to regulate leaf temperature (Ehleringer and Björkman, 1978; Ripley et al., 1999), and chiefly prevent plant organs from dehydration (Riederer and Schreiber, 2001).The epidermis of plants has been found to have a major degree of physical and chemical variability and may often contain specialized cells such as trichomes or stomata (Roth-Nebelsick et al., 2009; Javelle et al., 2011). Most aerial organs are covered with an extracellular and generally lipid-rich layer named the cuticle, which is typically composed of waxes embedded in (intracuticular waxes) or deposited on (epicuticular waxes) a biopolymer matrix of cutin, forming a network of cross-esterified hydroxy C16 and/or C18 fatty acids, and/or cutan, with variable amounts of polysaccharides and phenolics (Domínguez et al., 2011; Yeats and Rose, 2013). Different nano- and/or microscale levels of plant surface sculpturing have been observed by scanning electron microscopy (SEM), generally in relation to the topography of epicuticular waxes, cuticular folds, and epidermal cells (Koch and Barthlott, 2009). Such surface features together with their chemical composition (Khayet and Fernández, 2012) may lead to a high degree of roughness and hydrophobicity (Koch and Barthlott, 2009; Konrad et al., 2012). The interactions of plant surfaces with water have been addressed in some investigations (Brewer et al., 1991; Brewer and Smith, 1997; Pandey and Nagar, 2003; Hanba et al., 2004; Dietz et al., 2007; Holder, 2007a, 2007b; Fernández et al., 2011, 2014; Roth-Nebelsick et al., 2012; Wen et al., 2012; Urrego-Pereira et al., 2013) and are a topic of growing interest for plant ecophysiology (Helliker and Griffiths, 2007; Aryal and Neuner, 2010; Limm and Dawson, 2010; Kim and Lee, 2011; Berry and Smith, 2012; Berry et al., 2013; Rosado and Holder, 2013; Helliker, 2014). On the other hand, the mechanisms of foliar uptake of water and solutes by plant surfaces are still not fully understood (Fernández and Eichert, 2009; Burkhardt and Hunsche, 2013), but they may play an important ecophysiological role (Limm et al., 2009; Johnstone and Dawson, 2010; Adamec, 2013; Berry et al., 2014).The importance of trichomes and pubescent layers on water drop-plant surface interactions and on the subsequent potential water uptake into the organs has been analyzed in some investigations (Fahn, 1986; Brewer et al., 1991; Grammatikopoulos and Manetas, 1994; Brewer and Smith, 1997; Pierce et al., 2001; Kenzo et al., 2008; Fernández et al., 2011, 2014; Burrows et al., 2013). Trichomes are unicellular or multicellular and glandular or nonglandular appendages, which originate from epidermal cells only and develop outwards on the surface of plant organs (Werker, 2000). Nonglandular trichomes are categorized according to their morphology and exhibit a major variability in size, morphology, and function. On the other hand, glandular trichomes are classified by the secretory materials they excrete, accumulate, or absorb (Johnson, 1975; Werker, 2000; Wagner et al., 2004). Trichomes can be often found in xeromorphic leaves and in young organs (Fahn, 1986; Karabourniotis et al., 1995). The occurrence of protecting leaf trichomes has been also reported for Mediterranean species such as holm oak (Quercus ilex; Karabourniotis et al., 1995, 1998; Morales et al., 2002; Karioti et al., 2011; Camarero et al., 2012). There is limited information about the nature of the surface of trichomes, but they are also covered with a cuticle similarly to other epidermal cell types (Fernández et al., 2011, 2014).In this study and using holm oak as a model, we assessed, for the first time, the leaf surface-water relations of the abaxial (always pubescent) versus the adaxial (only pubescent in developing leaves and for a few months) surface, including their capacity to absorb surface-deposited water drops. Based on membrane science methodologies (Fernández et al., 2011; Khayet and Fernández, 2012) and following a new integrative approach, the chemical, physical, and anatomical properties of holm oak leaf surfaces and trichomes were analyzed, with the aim of addressing the following questions. Are young and mature adaxial and abaxial leaf surfaces capable of absorbing water deposited as drops on to the surfaces? Are young and mature abaxial and adaxial leaf surfaces similar in relation to their wettability, hydrophobicity, polarity, work of adhesion (Wa) for water, solubility parameter (δ), and surface free energy (γ)? What is the physical and chemical nature of the adaxial versus the abaxial trichomes, chiefly in relation to young leaves?  相似文献   

20.
The threat to global food security of stagnating yields and population growth makes increasing crop productivity a critical goal over the coming decades. One key target for improving crop productivity and yields is increasing the efficiency of photosynthesis. Central to photosynthesis is Rubisco, which is a critical but often rate-limiting component. Here, we present full Rubisco catalytic properties measured at three temperatures for 75 plants species representing both crops and undomesticated plants from diverse climates. Some newly characterized Rubiscos were naturally “better” compared to crop enzymes and have the potential to improve crop photosynthetic efficiency. The temperature response of the various catalytic parameters was largely consistent across the diverse range of species, though absolute values showed significant variation in Rubisco catalysis, even between closely related species. An analysis of residue differences among the species characterized identified a number of candidate amino acid substitutions that will aid in advancing engineering of improved Rubisco in crop systems. This study provides new insights on the range of Rubisco catalysis and temperature response present in nature, and provides new information to include in models from leaf to canopy and ecosystem scale.In a changing climate and under pressure from a population set to hit nine billion by 2050, global food security will require massive changes to the way food is produced, distributed, and consumed (Ort et al., 2015). To match rising demand, agricultural production must increase by 50 to 70% in the next 35 years, and yet the gains in crop yields initiated by the green revolution are slowing, and in some cases, stagnating (Long and Ort, 2010; Ray et al., 2012). Among a number of areas being pursued to increase crop productivity and food production, improving photosynthetic efficiency is a clear target, offering great promise (Parry et al., 2007; von Caemmerer et al., 2012; Price et al., 2013; Ort et al., 2015). As the gatekeeper of carbon entry into the biosphere and often acting as the rate-limiting step of photosynthesis, Rubisco, the most abundant enzyme on the planet (Ellis, 1979), is an obvious and important target for improving crop photosynthetic efficiency.Rubisco is considered to exhibit comparatively poor catalysis, in terms of catalytic rate, specificity, and CO2 affinity (Tcherkez et al., 2006; Andersson, 2008), leading to the suggestion that even small increases in catalytic efficiency may result in substantial improvements to carbon assimilation across a growing season (Zhu et al., 2004; Parry et al., 2013; Galmés et al., 2014a; Carmo-Silva et al., 2015). If combined with complimentary changes such as optimizing other components of the Calvin Benson or photorespiratory cycles (Raines, 2011; Peterhansel et al., 2013; Simkin et al., 2015), optimized canopy architecture (Drewry et al., 2014), or introducing elements of a carbon concentrating mechanism (Furbank et al., 2009; Lin et al., 2014a; Hanson et al., 2016; Long et al., 2016), Rubisco improvement presents an opportunity to dramatically increase the photosynthetic efficiency of crop plants (McGrath and Long, 2014; Long et al., 2015; Betti et al., 2016). A combination of the available strategies is essential for devising tailored solutions to meet the varied requirements of different crops and the diverse conditions under which they are typically grown around the world.Efforts to engineer an improved Rubisco have not yet produced a “super Rubisco” (Parry et al., 2007; Ort et al., 2015). However, advances in engineering precise changes in model systems continue to provide important developments that are increasing our understanding of Rubisco catalysis (Spreitzer et al., 2005; Whitney et al., 2011a, 2011b; Morita et al., 2014; Wilson et al., 2016), regulation (Andralojc et al., 2012; Carmo-Silva and Salvucci, 2013; Bracher et al., 2015), and biogenesis (Saschenbrecker et al., 2007; Whitney and Sharwood, 2008; Lin et al., 2014b; Hauser et al., 2015; Whitney et al., 2015).A complementary approach is to understand and exploit Rubisco natural diversity. Previous characterization of Rubisco from a limited number of species has not only demonstrated significant differences in the underlying catalytic parameters, but also suggests that further undiscovered diversity exists in nature and that the properties of some of these enzymes could be beneficial if present in crop plants (Carmo-Silva et al., 2015). Recent studies clearly illustrate the variation possible among even closely related species (Galmés et al., 2005, 2014b, 2014c; Kubien et al., 2008; Andralojc et al., 2014; Prins et al., 2016).Until recently, there have been relatively few attempts to characterize the consistency, or lack thereof, of temperature effects on in vitro Rubisco catalysis (Sharwood and Whitney, 2014), and often studies only consider a subset of Rubisco catalytic properties. This type of characterization is particularly important for future engineering efforts, enabling specific temperature effects to be factored into any attempts to modify crops for a future climate. In addition, the ability to coanalyze catalytic properties and DNA or amino acid sequence provides the opportunity to correlate sequence and biochemistry to inform engineering studies (Christin et al., 2008; Kapralov et al., 2011; Rosnow et al., 2015). While the amount of gene sequence information available grows rapidly with improving technology, knowledge of the corresponding biochemical variation resulting has yet to be determined (Cousins et al., 2010; Carmo-Silva et al., 2015; Sharwood and Whitney, 2014; Nunes-Nesi et al., 2016).This study aimed to characterize the catalytic properties of Rubisco from diverse species, comprising a broad range of monocots and dicots from diverse environments. The temperature dependence of Rubisco catalysis was evaluated to tailor Rubisco engineering for crop improvement in specific environments. Catalytic diversity was analyzed alongside the sequence of the Rubisco large subunit gene, rbcL, to identify potential catalytic switches for improving photosynthesis and productivity. In vitro results were compared to the average temperature of the warmest quarter in the regions where each species grows to investigate the role of temperature in modulating Rubisco catalysis.  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