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.
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   

2.
3.
4.
5.
6.
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   

7.
8.
Salinity affects a significant portion of arable land and is particularly detrimental for irrigated agriculture, which provides one-third of the global food supply. Rice (Oryza sativa), the most important food crop, is salt sensitive. The genetic resources for salt tolerance in rice germplasm exist but are underutilized due to the difficulty in capturing the dynamic nature of physiological responses to salt stress. The genetic basis of these physiological responses is predicted to be polygenic. In an effort to address this challenge, we generated temporal imaging data from 378 diverse rice genotypes across 14 d of 90 mm NaCl stress and developed a statistical model to assess the genetic architecture of dynamic salinity-induced growth responses in rice germplasm. A genomic region on chromosome 3 was strongly associated with the early growth response and was captured using visible range imaging. Fluorescence imaging identified four genomic regions linked to salinity-induced fluorescence responses. A region on chromosome 1 regulates both the fluorescence shift indicative of the longer term ionic stress and the early growth rate decline during salinity stress. We present, to our knowledge, a new approach to capture the dynamic plant responses to its environment and elucidate the genetic basis of these responses using a longitudinal genome-wide association model.Nearly one-third of the 54 million ha of the highly saline soils in the world are located in South and Southeast Asia. Rice (Oryza sativa), which is the primary source of calories and protein for these two regions, is very sensitive to salinity stress, with even moderate salinity levels known to decrease yields by 50% (Zeng et al., 2002). Projected sea level rise due to climate change is expected to increase saltwater ingress in coastal rice-growing regions of South and Southeast Asia. Therefore, development of salt-tolerant rice cultivars is essential to maintain rice productivity in the salinity-affected regions globally.Salt tolerance, defined as the ability to maintain growth and productivity in saline conditions, is a complex polygenic trait that may be influenced by distinct physiological mechanisms (Munns et al., 1982; Munns and Termaat, 1986; Cheeseman, 1988; Munns and Tester, 2008; Horie et al., 2012; for a comprehensive review of genes involved in salinity tolerance in rice, see Negrão et al., 2011) At the cellular level, plants respond to saline conditions in two phases, namely an osmotic (shoot ion independent) and an ionic stress phase, which can occur in an overlapping manner with varying intensity during the course of salinity stress (Munns and Termaat, 1986; Munns, 2002; Munns and James, 2003; Munns and Tester, 2008; Horie et al., 2012). During the osmotic stress phase, which occurs soon after the onset of salinity, the reduction in external osmotic potential disrupts water uptake and impedes cell expansion, which, at the whole plant level, leads to reduced growth rate (Matsuda and Riazi, 1981; Munns and Passioura, 1984; Rawson and Munns, 1984; Azaizeh and Steudle, 1991; Fricke and Peters, 2002; Fricke, 2004; Boursiac et al., 2005). As salinity stress persists over several days and weeks, sodium ions (Na+) accumulate to toxic levels, resulting in cell death and precocious leaf senescence (Lutts and Bouharmont, 1996; Munns, 2002; Munns and James, 2003; Ghanem et al., 2008). This is typically observed during the ionic phase of the salinity response (Munns, 2002; Munns and James, 2003; Munns and Tester, 2008). Plants possess distinct mechanisms to adapt to these osmotic and ionic stresses that are controlled by a suite of genes (Apse et al., 1999; Carvajal et al., 1999; Halfter et al., 2000; Ishitani et al., 2000; Shi et al., 2000; Zeng and Shannon, 2000; Rus et al., 2001; Berthomieu et al., 2003; Martínez-Ballesta et al., 2003; Boursiac et al., 2005, 2008; Ren et al., 2005; Huang et al., 2006; Davenport et al., 2007; Obata et al., 2007; Székely et al., 2008; Horie et al., 2011; Rivandi et al., 2011; Asano et al., 2012; Munns et al., 2012; Latz et al., 2013; Schmidt et al., 2013; Campo et al., 2014; Choi et al., 2014; Liu et al., 2014). The genetic basis of temporal adaptive responses to salinity stress remains to be explored in rice and other crops. This is primarily due to challenges in capturing the dynamic physiological responses to salinity for a large number of genotypes in a nondestructive manner. Manual phenotyping to detect incremental changes in growth rate during the osmotic stress and slight shifts in leaf color due to ionic stress is difficult to quantify for a large number of genotypes.In rice, at least one major quantitative trait loci (QTL; saltol) for salinity tolerance has been characterized based on end point measurements of biomass, senescence/injury, and Na+ and K+ concentrations (Bonilla et al., 2002; Lin et al., 2004; Thomson et al., 2010). SHOOT K+ CONTENT1 (SKC1) is the causative gene underlying the saltol region. SKC1 encodes a Na+-selective high-affinity potassium transporter that regulates Na+/K+ homeostasis during salinity stress (Ren et al., 2005). High levels of Na+ displace cellular K+, an essential element for several enzymatic reactions and physiological processes (Gierth and Mäser, 2007). The ability to maintain cellular K+ levels during salinity through the action of Na+-selective potassium transporters or Na+/H+ antiporters is a well-characterized tolerance mechanism in cereals including rice (Ren et al., 2005; Sunarpi et al., 2005; Huang et al., 2006; Møller et al., 2009; Mian et al., 2011; Munns et al., 2012).Numerous studies have utilized conventional linkage mapping to identify QTL for morphological and physiological responses to salinity in rice using discrete end point measurements (Bonilla et al., 2002; Lin et al., 2004; Ren et al., 2005; Negrão et al., 2011; Wang et al., 2012). However, the physiological adaptation to saline conditions is a complex continuous process that develops over time. While some accessions will exhibit similar end point phenotypic values, the genetic and physiological mechanisms giving rise to the similar phenotypes may be very different and the growth trajectories throughout the experiment may be distinct. Although single time point studies have yielded important information regarding the genetic basis of salinity tolerance, such approaches are too simple to reveal the genetic architecture of stress adaptation. With the advent of high-throughput image-based phenotyping platforms, it is now feasible to quantify dynamic responses during the stress treatment for a large number of genotypes (Berger et al., 2010; Golzarian et al., 2011; Chen et al., 2014; Honsdorf et al., 2014).Image-based phenotyping has been combined with genome-wide association studies (GWAS) and linkage mapping to examine the genetic basis of complex developmental processes (Busemeyer et al., 2013; Moore et al., 2013; Topp et al., 2013; Slovak et al., 2014; Würschum et al., 2014; Yang et al., 2014; Bac-Molenaar et al., 2015). Moreover, the introduction of the time axis provides a better understanding of the physiological processes underlying complex stress and developmental responses compared with single end point measurements (Zhang et al., 2012; Moore et al., 2013; Brown et al., 2014; Chen et al., 2014; Slovak et al., 2014; Bac-Molenaar et al., 2015). However, to date, no studies have implemented an association mapping approach using image-derived phenotypes to address the genetic basis of dynamic stress responses in plants. Image-based phenotyping offers several advantages over conventional phenotyping: (1) quantitative measurements can be recorded over discrete time points to capture morphological and physiological responses in a nondestructive manner, and (2) the use of various types of spectral imaging address phenotypes that are not detectable to the human eye such as chlorophyll fluorescence and leaf water content. Integrating dynamic phenotypic data and association mapping has the potential to query genetic diversity across hundreds of accessions for complex traits and provides much higher resolution compared with conventional linkage mapping. Here, we explored the dynamic growth and chlorophyll responses to salinity of a diverse set of rice accessions using high-throughput visible and fluorescence imaging. To assess the genetic basis of plant growth in saline conditions, a logistic model was used to describe the temporal growth responses and was incorporated into the statistical framework necessary for association mapping. Coupled with temporal fluorescence imaging, we present, to our knowledge, new insights into the genetic architecture of osmotic and ionic responses during salinity stress in rice.  相似文献   

9.
Cytosolic Ca2+ in guard cells plays an important role in stomatal movement responses to environmental stimuli. These cytosolic Ca2+ increases result from Ca2+ influx through Ca2+-permeable channels in the plasma membrane and Ca2+ release from intracellular organelles in guard cells. However, the genes encoding defined plasma membrane Ca2+-permeable channel activity remain unknown in guard cells and, with some exceptions, largely unknown in higher plant cells. Here, we report the identification of two Arabidopsis (Arabidopsis thaliana) cation channel genes, CNGC5 and CNGC6, that are highly expressed in guard cells. Cytosolic application of cyclic GMP (cGMP) and extracellularly applied membrane-permeable 8-Bromoguanosine 3′,5′-cyclic monophosphate-cGMP both activated hyperpolarization-induced inward-conducting currents in wild-type guard cells using Mg2+ as the main charge carrier. The cGMP-activated currents were strongly blocked by lanthanum and gadolinium and also conducted Ba2+, Ca2+, and Na+ ions. cngc5 cngc6 double mutant guard cells exhibited dramatically impaired cGMP-activated currents. In contrast, mutations in CNGC1, CNGC2, and CNGC20 did not disrupt these cGMP-activated currents. The yellow fluorescent protein-CNGC5 and yellow fluorescent protein-CNGC6 proteins localize in the cell periphery. Cyclic AMP activated modest inward currents in both wild-type and cngc5cngc6 mutant guard cells. Moreover, cngc5 cngc6 double mutant guard cells exhibited functional abscisic acid (ABA)-activated hyperpolarization-dependent Ca2+-permeable cation channel currents, intact ABA-induced stomatal closing responses, and whole-plant stomatal conductance responses to darkness and changes in CO2 concentration. Furthermore, cGMP-activated currents remained intact in the growth controlled by abscisic acid2 and abscisic acid insensitive1 mutants. This research demonstrates that the CNGC5 and CNGC6 genes encode unique cGMP-activated nonselective Ca2+-permeable cation channels in the plasma membrane of Arabidopsis guard cells.Plants lose water via transpiration and take in CO2 for photosynthesis through stomatal pores. Each stomatal pore is surrounded by two guard cells, and stomatal movements are driven by the change of turgor pressure in guard cells. The intracellular second messenger Ca2+ functions in guard cell signal transduction (Schroeder and Hagiwara, 1989; McAinsh et al., 1990; Webb et al., 1996; Grabov and Blatt, 1998; Allen et al., 1999; MacRobbie, 2000; Mori et al., 2006; Young et al., 2006; Siegel et al., 2009; Chen et al., 2010; Hubbard et al., 2012). Plasma membrane ion channel activity and gene expression in guard cells are finely regulated by the intracellular free calcium concentration ([Ca2+]cyt; Schroeder and Hagiwara, 1989; Webb et al., 2001; Allen et al., 2002; Siegel et al., 2009; Kim et al., 2010; Stange et al., 2010). Ca2+-dependent protein kinases (CPKs) function as targets of the cytosolic Ca2+ signal, and several members of the CPK family have been shown to function in stimulus-induced stomatal closing, including the Arabidopsis (Arabidopsis thaliana) CPK3, CPK4, CPK6, CPK10, and CPK11 proteins (Mori et al., 2006; Zhu et al., 2007; Zou et al., 2010; Brandt et al., 2012; Hubbard et al., 2012). Further research found that several CPKs could activate the S-type anion channel SLAC1 in Xenopus laevis oocytes, including CPK21, CPK23, and CPK6 (Geiger et al., 2010; Brandt et al., 2012). At the same time, the Ca2+-independent protein kinase Open Stomata1 mediates stomatal closing and activates the S-type anion channel SLAC1 (Mustilli et al., 2002; Yoshida et al., 2002; Geiger et al., 2009; Lee et al., 2009; Xue et al., 2011), indicating that both Ca2+-dependent and Ca2+-independent pathways function in guard cells.Multiple essential factors of guard cell abscisic acid (ABA) signal transduction function in the regulation of Ca2+-permeable channels and [Ca2+]cyt elevations, including Abscisic Acid Insensitive1 (ABI1), ABI2, Enhanced Response to Abscisic Acid1 (ERA1), the NADPH oxidases AtrbohD and AtrbohF, the Guard Cell Hydrogen Peroxide-Resistant1 (GHR1) receptor kinase, as well as the Ca2+-activated CPK6 protein kinase (Pei et al., 1998; Allen et al., 1999, 2002; Kwak et al., 2003; Miao et al., 2006; Mori et al., 2006; Hua et al., 2012). [Ca2+]cyt increases result from both Ca2+ release from intracellular Ca2+ stores (McAinsh et al., 1992) and Ca2+ influx across the plasma membrane (Hamilton et al., 2000; Pei et al., 2000; Murata et al., 2001; Kwak et al., 2003; Hua et al., 2012). Electrophysiological analyses have characterized nonselective Ca2+-permeable channel activity in the plasma membrane of guard cells (Schroeder and Hagiwara, 1990; Hamilton et al., 2000; Pei et al., 2000; Murata et al., 2001; Köhler and Blatt, 2002; Miao et al., 2006; Mori et al., 2006; Suh et al., 2007; Vahisalu et al., 2008; Hua et al., 2012). However, the genetic identities of Ca2+-permeable channels in the plasma membrane of guard cells have remained unknown despite over two decades of research on these channel activities.The Arabidopsis genome includes 20 genes encoding cyclic nucleotide-gated channel (CNGC) homologs and 20 genes encoding homologs to animal Glu receptor channels (Lacombe et al., 2001; Kaplan et al., 2007; Ward et al., 2009), which have been proposed to function in plant cells as cation channels (Schuurink et al., 1998; Arazi et al., 1999; Köhler et al., 1999). Recent research has demonstrated functions of specific Glu receptor channels in mediating Ca2+ channel activity (Michard et al., 2011; Vincill et al., 2012). Previous studies have shown cAMP activation of nonselective cation currents in guard cells (Lemtiri-Chlieh and Berkowitz, 2004; Ali et al., 2007). However, only a few studies have shown the disappearance of a defined plasma membrane Ca2+ channel activity in plants upon mutation of candidate Ca2+ channel genes (Ali et al., 2007; Michard et al., 2011; Laohavisit et al., 2012; Vincill et al., 2012). Some CNGCs have been found to be involved in cation nutrient intake, including monovalent cation intake (Guo et al., 2010; Caballero et al., 2012), salt tolerance (Guo et al., 2008; Kugler et al., 2009), programmed cell death and pathogen responses (Clough et al., 2000; Balagué et al., 2003; Urquhart et al., 2007; Abdel-Hamid et al., 2013), thermal sensing (Finka et al., 2012; Gao et al., 2012), and pollen tube growth (Chang et al., 2007; Frietsch et al., 2007; Tunc-Ozdemir et al., 2013a, 2013b). Direct in vivo disappearance of Ca2+ channel activity in cngc disruption mutants has been demonstrated in only a few cases thus far (Ali et al., 2007; Gao et al., 2012). In this research, we show that CNGC5 and CNGC6 are required for a cyclic GMP (cGMP)-activated nonselective Ca2+-permeable cation channel activity in the plasma membrane of Arabidopsis guard cells.  相似文献   

10.
Insufficient accumulation levels of recombinant proteins in plants and the lack of efficient purification methods for recovering these valuable proteins have hindered the development of plant biotechnology applications. Hydrophobins are small and surface-active proteins derived from filamentous fungi that can be easily purified by a surfactant-based aqueous two-phase system. In this study, the hydrophobin HFBI sequence from Trichoderma reesei was fused to green fluorescent protein (GFP) and transiently expressed in Nicotiana benthamiana plants by Agrobacterium tumefaciens infiltration. The HFBI fusion significantly enhanced the accumulation of GFP, with the concentration of the fusion protein reaching 51% of total soluble protein, while also delaying necrosis of the infiltrated leaves. Furthermore, the endoplasmic reticulum-targeted GFP-HFBI fusion induced the formation of large novel protein bodies. A simple and scalable surfactant-based aqueous two-phase system was optimized to recover the HFBI fusion proteins from leaf extracts. The single-step phase separation was able to selectively recover up to 91% of the GFP-HFBI up to concentrations of 10 mg mL−1. HFBI fusions increased the expression levels of plant-made recombinant proteins while also providing a simple means for their subsequent purification. This hydrophobin fusion technology, when combined with the speed and posttranslational modification capabilities of plants, enhances the value of transient plant-based expression systems.As the amount of plant genome and proteome information increases, the need has arisen to develop technologies to rapidly overexpress these genes and to characterize the proteins at the structural and functional levels. Based on two decades of research, plant expression platforms are now recognized as a safe, effective, and inexpensive means of producing heterologous recombinant proteins (Ma et al., 2003).Agroinfiltration in Nicotiana benthamiana leaves (Kapila et al., 1997; Yang et al., 2000), when combined with the coexpression of a suppressor of gene silencing (Silhavy et al., 2002; Voinnet et al., 2003), has established itself as the most utilized transient expression system in plants. Agroinfiltration is a fast and convenient technique, producing recombinant protein within 2 to 5 d. This transient expression system is also flexible, as it allows for the expression of multiple genes simultaneously (Johansen and Carrington, 2001) and the transfer of relatively large genes (greater than 2 kb), which are genetically unstable in viral vectors (Porta and Lomonossoff, 1996). Although typically used for preliminary laboratory-scale analyses, agroinfiltration is now being scaled up for the rapid production of gram quantities of recombinant proteins in plants (Vézina et al., 2009).Despite the success of plant expression systems, two major challenges still limiting the economical production of plant-made recombinant proteins include inadequate accumulation levels and the lack of efficient purification methods. Thus, several protein fusion strategies have been developed to address these issues (Terpe, 2003). For example, the use of protein-stabilizing fusion partners, such as ubiquitin (Garbarino et al., 1995; Hondred et al., 1999; Mishra et al., 2006), β-glucuronidase (Gil et al., 2001; Dus Santos et al., 2002), cholera toxin B subunit (Arakawa et al., 2001; Kim et al., 2004; Molina et al., 2004), viral coat proteins (Canizares et al., 2005), and human IgG α-chains (Obregon et al., 2006), are common approaches for enhancing recombinant protein accumulation in plants. To simplify purification, recombinant proteins are often fused translationally to small affinity tags or proteins with defined binding characteristics, such as the StrepII tag, Arg tag, His tag, FLAG tag, c-myc tag, glutathione S-transferase tag, calmodulin-binding peptide, maltose-binding protein, and cellulose-binding domain (Terpe, 2003; Witte et al., 2004; Lichty et al., 2005; Rubio et al., 2005; Streatfield 2007). However, these affinity chromatography methods are often ineffective when purifying proteins from the complex plant proteome and are costly and difficult to scale up for industrial applications (Waugh, 2005).More recently, elastin-like polypeptide (ELP) and Zera protein fusions have been shown to significantly enhance recombinant protein accumulation in the leaves of plants (Patel et al., 2007; Floss et al., 2008; Conley et al., 2009c; Torrent et al., 2009) while also providing a means for their purification. ELPs are thermally responsive synthetic biopolymers composed of a repeating pentapeptide (VPGXG) sequence (Urry, 1988) that are valuable for the simple nonchromatographic “inverse transition cycling” bioseparation of recombinant proteins (Meyer and Chilkoti, 1999; Lin et al., 2006). However, the purity and recovery efficiency are rather low when using inverse transition cycling for the purification of plant-made proteins that accumulate to low levels, so expensive and tedious affinity chromatography steps are still needed in these cases (Conley et al., 2009a; Joensuu et al., 2009). Alternatively, Zera, the Pro-rich domain derived from the maize (Zea mays) seed storage protein γ-zein, can facilitate the recovery and purification of fused recombinant proteins by density-based separation methods, but this technique is difficult to scale up (Torrent et al., 2009). Interestingly, both of these protein fusions, derived from taxonomically distinct kingdoms, have been shown to induce the formation of novel endoplasmic reticulum (ER)-derived protein bodies (PBs; Conley et al., 2009b; Torrent et al., 2009). These PBs are physiologically inert and allow for the stable storage of large amounts of recombinant protein within the cell. To overcome the current limitations of the ELP and Zera purification schemes, we chose to investigate hydrophobins as fusion partners for the expression and purification of plant-made recombinant proteins, since they share many interesting physicochemical properties with ELP and Zera.Hydrophobins are small surface-active fungal proteins that have a characteristic pattern of eight conserved Cys residues, which form four intramolecular disulfide bridges and are responsible for stabilizing the protein''s structure (Hakanpaa et al., 2004). In nature, hydrophobins contribute to surface hydrophobicity and function to coat various fungal structures important for growth and development (Linder, 2009). Hydrophobins have a propensity to self-assemble into an amphipathic protein membrane at hydrophilic-hydrophobic interfaces (Wösten and de Vocht, 2000; Paananen et al., 2003; Wang et al., 2005). Because of these unique properties, hydrophobins have numerous potential applications, including the ability to interface proteins with nonbiological surfaces, to alter the wettability of different materials, to act as biosurfactants and oil stabilizers, and to form medical and technical coatings (Wessels, 1997; Askolin et al., 2001; Linder et al., 2005; Linder, 2009).Hydrophobins are also capable of altering the hydrophobicity of their respective fusion partners, thus enabling efficient purification using a surfactant-based aqueous two-phase system (ATPS; Linder et al., 2004). The ATPS concentrates the hydrophobin fusions inside micellar structures and partitions them toward the surfactant phase (Lahtinen et al., 2008). ATPSs offer several benefits, since they are simple, rapid, and inexpensive while providing volume reduction, high capacity, and fast separations (Persson et al., 1999). Most importantly, the one-step ATPS purification is particularly attractive because it can be easily and effectively scaled up for industrial-scale protein purification (Linder et al., 2004; Selber et al., 2004).Here, we used agroinfiltration to study the effect of a hydrophobin fusion on the accumulation of GFP and the commercially valuable enzyme Glc oxidase (GOx). We also determined the capability of hydrophobins for purifying recombinant proteins from leaf extracts using an ATPS. The hydrophobin fusion partner significantly enhanced the production yield of GFP while also providing a simple, efficient, and inexpensive approach for the purification of recombinant proteins from plants.  相似文献   

11.
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   

12.
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.
Although cytosolic free Ca2+ mobilization induced by microbe/pathogen-associated molecular patterns is postulated to play a pivotal role in innate immunity in plants, the molecular links between Ca2+ and downstream defense responses still remain largely unknown. Calcineurin B-like proteins (CBLs) act as Ca2+ sensors to activate specific protein kinases, CBL-interacting protein kinases (CIPKs). We here identified two CIPKs, OsCIPK14 and OsCIPK15, rapidly induced by microbe-associated molecular patterns, including chitooligosaccharides and xylanase (Trichoderma viride/ethylene-inducing xylanase [TvX/EIX]), in rice (Oryza sativa). Although they are located on different chromosomes, they have over 95% nucleotide sequence identity, including the surrounding genomic region, suggesting that they are duplicated genes. OsCIPK14/15 interacted with several OsCBLs through the FISL/NAF motif in yeast cells and showed the strongest interaction with OsCBL4. The recombinant OsCIPK14/15 proteins showed Mn2+-dependent protein kinase activity, which was enhanced both by deletion of their FISL/NAF motifs and by combination with OsCBL4. OsCIPK14/15-RNAi transgenic cell lines showed reduced sensitivity to TvX/EIX for the induction of a wide range of defense responses, including hypersensitive cell death, mitochondrial dysfunction, phytoalexin biosynthesis, and pathogenesis-related gene expression. On the other hand, TvX/EIX-induced cell death was enhanced in OsCIPK15-overexpressing lines. Our results suggest that OsCIPK14/15 play a crucial role in the microbe-associated molecular pattern-induced defense signaling pathway in rice cultured cells.Calcium ions regulate diverse cellular processes in plants as a ubiquitous internal second messenger, conveying signals received at the cell surface to the inside of the cell through spatial and temporal concentration changes that are decoded by an array of Ca2+ sensors (Reddy, 2001; Sanders et al., 2002; Yang and Poovaiah, 2003). Several families of Ca2+ sensors have been identified in higher plants. The best known are calmodulins (CaMs) and CaM-related proteins, which typically contain four EF-hand domains for Ca2+ binding (Zielinski, 1998). Unlike mammals, which possess single molecular species of CaM, plants have at least three distinct molecular species of CaM playing diverse physiological functions and whose expression is differently regulated (Yamakawa et al., 2001; Luan et al., 2002; Karita et al., 2004; Takabatake et al., 2007). The second major class is exemplified by the Ca2+-dependent protein kinases, which contain CaM-like Ca2+-binding domains and a kinase domain in a single protein (Harmon et al., 2000). In addition, a new family of Ca2+ sensors was identified as calcineurin B-like (CBL) proteins, which consists of proteins similar to both the regulatory β-subunit of calcineurin and the neuronal Ca2+ sensor in animals (Liu and Zhu, 1998; Kudla et al., 1999).Unlike CaMs, which interact with a large variety of target proteins, CBLs specifically target a family of protein kinases referred to as CBL-interacting protein kinases (CIPKs) or SnRK3s (for sucrose nonfermenting 1-related protein kinases type 3), which are most similar to the SNF family protein kinases in yeast (Luan et al., 2002). A database search of the Arabidopsis (Arabidopsis thaliana) genome sequence revealed 10 CBL and 25 CIPK homologues (Luan et al., 2002). Expression patterns of these Ca2+ sensors and protein kinases suggest their diverse functions in different signaling processes, including light, hormone, sugar, and stress responses (Batistic and Kudla, 2004). AtCBL4/Salt Overly Sensitive3 (SOS3) and AtCIPK24/SOS2 have been shown to play a key role in Ca2+-mediated salt stress adaptation (Zhu, 2002). The CBL-CIPK system has been shown to be involved in signaling pathways of abscisic acid (Kim et al., 2003a), sugar (Gong et al., 2002a), gibberellins (Hwang et al., 2005), salicylic acid (Mahajan et al., 2006), and K+ channel regulation (Li et al., 2006; Lee et al., 2007; for review, see Luan, 2009; Batistic and Kudla, 2009). However, physiological functions of most of the family members still remain largely unknown.Plants respond to pathogen attack by activating a variety of defense responses, including the generation of reactive oxygen species (ROS), synthesis of phytoalexins, expression of pathogenesis-related (PR) genes, cell cycle arrest, and mitochondrial dysfunction followed by a form of hypersensitive cell death known as the hypersensitive response (Nürnberger and Scheel, 2001; Greenberg and Yao, 2004; Kadota et al., 2004b). Transient membrane potential changes and Ca2+ influx are involved at the initial stage of defense responses (Kuchitsu et al., 1993; Pugin et al., 1997; Blume et al., 2000; Kadota et al., 2004a). Many kinds of defense responses are prevented when Ca2+ influx is compromised by Ca2+ chelators (Nürnberger and Scheel, 2001; Lecourieux et al., 2002). Since complex spatiotemporal patterns of cytosolic free Ca2+ concentration have been suggested to play pivotal roles in defense signaling (Nürnberger and Scheel, 2001; Sanders et al., 2002), multiple Ca2+ sensor proteins and their effectors should function in the defense signaling pathways. Although possible involvement of some CaM isoforms (Heo et al., 1999; Yamakawa et al., 2001), Ca2+-dependent protein kinases (Romeis et al., 2000, 2001; Ludwig et al., 2005; Kobayashi et al., 2007; Yoshioka et al., 2009), as well as Ca2+ regulation of EF-hand-containing enzymes such as ROS-generating NADPH oxidase (Ogasawara et al., 2008) have been suggested, other Ca2+-regulated signaling components still remain to be identified. No CBLs or CIPKs have so far been implicated as signaling components in defense signaling.N-Acetylchitooligosaccharides, chitin fragments, are microbe-associated molecular patterns (MAMPs) that are recognized by plasma membrane receptors (Kaku et al., 2006; Miya et al., 2007) and induce a variety of defense responses, such as membrane depolarization (Kuchitsu et al., 1993; Kikuyama et al., 1997), ion fluxes (Kuchitsu et al., 1997), ROS production (Kuchitsu et al., 1995), phytoalexin biosynthesis (Yamada et al., 1993), and induction of PR genes (Nishizawa et al., 1999), without hypersensitive cell death in rice (Oryza sativa) cells. In contrast, a fungal proteinaceous elicitor, xylanase from Trichoderma viride (TvX)/ethylene-inducing xylanase (EIX), which is recognized by two putative plasma membrane receptors, LeEix1 and LeEix2 (Ron and Avni, 2004), triggers hypersensitive cell death along with different kinetics of ROS production and activation of a mitogen-activated protein kinase, OsMPK6, previously named as OsMPK2 or OsMAPK6, in rice cells (Kurusu et al., 2005). These two fungal MAMPs thus provide excellent model systems to study innate immunity in rice cells.This study identified two CIPKs involved in various MAMP-induced layers of defense responses, including PR gene expression, phytoalexin biosynthesis, mitochondrial dysfunction, and cell death, in rice. Molecular characterization of these CIPKs, including interaction with the putative Ca2+ sensors as well as their physiological functions, is discussed.  相似文献   

15.
Calcium plays an essential role in pollen tube tip growth. However, little is known concerning the molecular basis of the signaling pathways involved. Here, we identified Arabidopsis (Arabidopsis thaliana) CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASE19 (CIPK19) as an important element to pollen tube growth through a functional survey for CIPK family members. The CIPK19 gene was specifically expressed in pollen grains and pollen tubes, and its overexpression induced severe loss of polarity in pollen tube growth. In the CIPK19 loss-of-function mutant, tube growth and polarity were significantly impaired, as demonstrated by both in vitro and in vivo pollen tube growth assays. Genetic analysis indicated that disruption of CIPK19 resulted in a male-specific transmission defect. Furthermore, loss of polarity induced by CIPK19 overexpression was associated with elevated cytosolic Ca2+ throughout the bulging tip, whereas LaCl3, a Ca2+ influx blocker, rescued CIPK19 overexpression-induced growth inhibition. Our results suggest that CIPK19 may be involved in maintaining Ca2+ homeostasis through its potential function in the modulation of Ca2+ influx.In flowering plants, fertilization is mediated by pollen tubes that extend directionally toward the ovule for sperm delivery (Krichevsky et al., 2007; Johnson, 2012). The formation of these elongated tubular structures is dependent on extreme polar growth (termed tip growth), in which cell expansion occurs exclusively in the very apical area (Yang, 2008; Rounds and Bezanilla, 2013). As this type of tip growth is amenable to genetic manipulation and cell biological analysis, the pollen tube is an excellent model system for the functional analysis of essential genes involved in polarity control and fertilization (Yang, 2008; Qin and Yang, 2011; Bloch and Yalovsky, 2013).It is well established that Ca2+ plays a critical role in pollen germination and tube growth (Konrad et al., 2011; Hepler et al., 2012). A steep tip-focused Ca2+ gradient has been detected at the tip of elongating pollen tubes (Rathore et al., 1991; Pierson et al., 1994; Hepler, 1997). In previous studies, artificial dissipation of the Ca2+ gradient seriously inhibited tip growth of pollen tubes, whereas elevation of internal Ca2+ level induced bending of the growth axis toward the zone of higher Ca2+. These studies suggest that Ca2+ not only controls pollen tube elongation but also modulates growth orientation (Miller et al., 1992; Malho et al., 1994; Malho and Trewavas, 1996; Hepler, 1997). These Ca2+ signatures are perceived and relayed to downstream responses by a complex toolkit of Ca2+-binding proteins that function as Ca2+ sensors (Yang and Poovaiah, 2003; Harper et al., 2004; Dodd et al., 2010).To date, four major Ca2+ sensor families have been identified in Arabidopsis (Arabidopsis thaliana), including calcium-dependent protein kinase, calmodulin (CaM), calmodulin-like (CML), and CALCINEURIN B-LIKE (CBL) proteins (Luan et al., 2002, 2009; Yang and Poovaiah, 2003; Harper et al., 2004). Calcium-dependent protein kinase family members comprise a kinase domain and a CaM-like domain in a single protein; thus, they act not only as a Ca2+ sensor but also as an effector, designated as sensor responders (Cheng et al., 2002). In contrast, CaM, CML, and CBL proteins do not have any enzymatic domains but transmit Ca2+ signals to downstream targets via Ca2+-dependent protein-protein interactions. Therefore, they have been designated as sensor relays (McCormack et al., 2005). While CaM and CML proteins interact with a diverse array of target proteins, it is generally accepted that CBLs interact specifically with a group of Ser/Thr protein kinases termed CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASEs (CIPKs; Luan et al., 2002; Kolukisaoglu et al., 2004).In Arabidopsis, several CBLs coupled with their target CIPKs have been demonstrated to function in the regulation of ion homeostasis and stress responses (Luan et al., 2009). Under salt stress, SALT OVERLY SENSITIVE3 (SOS3)/CBL4-SOS2/CIPK24 regulate SOS1 at the plasma membrane for Na+ exclusion, whereas CBL10-CIPK24 complexes appear to regulate Na+ sequestration at the tonoplast (Liu et al., 2000; Qiu et al., 2002; Kim et al., 2007; Quan et al., 2007). For low-K+ stress, CBL1 and CBL9, with 87% amino acid sequence identity, interact with CIPK23, which regulates a voltage-gated ion channel (ARABIDOPSIS K+ TRANSPORTER1) to mediate the uptake of K+ in root hairs (Li et al., 2006; Xu et al., 2006; Cheong et al., 2007). In addition, CBL1 integrates plant responses to cold, drought, salinity, and hyperosmotic stresses (Albrecht et al., 2003; Cheong et al., 2003), and CBL9 is involved in abscisic acid signaling and biosynthesis during seed germination (Pandey et al., 2004). Over the past decade, the functions of CBL-CIPK complexes in abiotic stress tolerance have been studied extensively, but only limited studies focus on CBL family members in pollen tube growth. For example, CBL3 overexpression caused a defective phenotype in pollen tube growth (Zhou et al., 2009). Overexpression of CBL1 or its closest homolog CBL9 inhibited pollen germination and perturbed tube growth at high external K+, whereas disruption of CBL1 and CBL9 leads to a significantly reduced growth rate of pollen tubes under low-K+ conditions (Mähs et al., 2013). The potential roles of CIPKs in pollen tubes so far appear to be completely unknown.In this study, we demonstrated that Arabidopsis CIPK19, a CIPK specifically expressed in pollen grains and pollen tubes, functions in pollen tube tip growth, providing a new insight into the function of the CBL-CIPK network in the control of growth polarity during pollen tube extension in fertilization.  相似文献   

16.
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   

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