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 78 毫秒
1.
2.
3.
Diverse stresses such as high salt conditions cause an increase in reactive oxygen species (ROS), necessitating a redox stress response. However, little is known about the signaling pathways that regulate the antioxidant system to counteract oxidative stress. Here, we show that a Glycogen Synthase Kinase3 from Arabidopsis thaliana (ASKα) regulates stress tolerance by activating Glc-6-phosphate dehydrogenase (G6PD), which is essential for maintaining the cellular redox balance. Loss of stress-activated ASKα leads to reduced G6PD activity, elevated levels of ROS, and enhanced sensitivity to salt stress. Conversely, plants overexpressing ASKα have increased G6PD activity and low levels of ROS in response to stress and are more tolerant to salt stress. ASKα stimulates the activity of a specific cytosolic G6PD isoform by phosphorylating the evolutionarily conserved Thr-467, which is implicated in cosubstrate binding. Our results reveal a novel mechanism of G6PD adaptive regulation that is critical for the cellular stress response.  相似文献   

4.
5.
The phragmoplast, a plant-specific apparatus that mediates cytokinesis, mainly consists of microtubules (MTs) arranged in a bipolar fashion, such that their plus ends interdigitate at the equator. Membrane vesicles are thought to move along the MTs toward the equator and fuse to form the cell plate. Although several genes required for phragmoplast MT organization have been identified, the mechanisms that maintain the bipolarity of phragmoplasts remain poorly understood. Here, we show that engaging phragmoplast MTs in a bipolar fashion in protonemal cells of the moss Physcomitrella patens requires the conserved MT cross-linking protein MICROTUBULE-ASSOCIATED PROTEIN65 (MAP65). Simultaneous knockdown of the three MAP65s expressed in those cells severely compromised MT interdigitation at the phragmoplast equator after anaphase onset, resulting in the collapse of the phragmoplast in telophase. Cytokinetic vesicles initially localized to the anaphase midzone as normal but failed to further accumulate in the next several minutes, although the bipolarity of the MT array was preserved. Our data indicate that the presence of bipolar MT arrays is insufficient for vesicle accumulation at the equator and further suggest that MAP65-mediated MT interdigitation is a prerequisite for maintenance of bipolarity of the phragmoplast and accumulation and/or fusion of cell plate–destined vesicles at the equatorial plane.  相似文献   

6.
Regulation of membrane lipid biosynthesis is critical for cell function. We previously reported that disruption of PHOSPHATIDIC ACID PHOSPHOHYDROLASE1 (PAH1) and PAH2 stimulates net phosphatidylcholine (PC) biosynthesis and proliferation of the endoplasmic reticulum (ER) in Arabidopsis thaliana. Here, we show that this response is caused specifically by a reduction in the catalytic activity of the protein and positively correlates with an accumulation of its substrate, phosphatidic acid (PA). The accumulation of PC in pah1 pah2 is suppressed by disruption of CTP:PHOSPHOCHOLINE CYTIDYLYLTRANSFERASE1 (CCT1), which encodes a key enzyme in the nucleotide pathway for PC biosynthesis. The activity of recombinant CCT1 is stimulated by lipid vesicles containing PA. Truncation of CCT1, to remove the predicted C-terminal amphipathic lipid binding domain, produced a constitutively active enzyme. Overexpression of native CCT1 in Arabidopsis has no significant effect on PC biosynthesis or ER morphology, but overexpression of the truncated constitutively active version largely replicates the pah1 pah2 phenotype. Our data establish that membrane homeostasis is regulated by lipid composition in Arabidopsis and reveal a mechanism through which the abundance of PA, mediated by PAH activity, modulates CCT activity to govern PC content.  相似文献   

7.
Remodeling of actin filament arrays in response to biotic and abiotic stimuli is thought to require precise control over the generation and availability of filament ends. Heterodimeric capping protein (CP) is an abundant filament capper, and its activity is inhibited by membrane signaling phospholipids in vitro. How exactly CP modulates the properties of filament ends in cells and whether its activity is coordinated by phospholipids in vivo is not well understood. By observing directly the dynamic behavior of individual filament ends in the cortical array of living Arabidopsis thaliana epidermal cells, we dissected the contribution of CP to actin organization and dynamics in response to the signaling phospholipid, phosphatidic acid (PA). Here, we examined three cp knockdown mutants and found that reduced CP levels resulted in more dynamic activity at filament ends, and this significantly enhanced filament-filament annealing and filament elongation from free ends. The cp mutants also exhibited more dense actin filament arrays. Treatment of wild-type cells with exogenous PA phenocopied the actin-based defects in cp mutants, with an increase in the density of filament arrays and enhanced annealing frequency. These cytoskeletal responses to exogenous PA were completely abrogated in cp mutants. Our data provide compelling genetic evidence that the end-capping activity of CP is inhibited by membrane signaling lipids in eukaryotic cells. Specifically, CP acts as a PA biosensor and key transducer of fluxes in membrane signaling phospholipids into changes in actin cytoskeleton dynamics.  相似文献   

8.
9.
The mechanism underlying microtubule (MT) generation in plants has been primarily studied using the cortical MT array, in which fixed-angled branching nucleation and katanin-dependent MT severing predominate. However, little is known about MT generation in the endoplasm. Here, we explored the mechanism of endoplasmic MT generation in protonemal cells of Physcomitrella patens. We developed an assay that utilizes flow cell and oblique illumination fluorescence microscopy, which allowed visualization and quantification of individual MT dynamics. MT severing was infrequently observed, and disruption of katanin did not severely affect MT generation. Branching nucleation was observed, but it showed markedly variable branch angles and was occasionally accompanied by the transport of nucleated MTs. Cytoplasmic nucleation at seemingly random locations was most frequently observed and predominated when depolymerized MTs were regrown. The MT nucleator γ-tubulin was detected at the majority of the nucleation sites, at which a single MT was generated in random directions. When γ-tubulin was knocked down, MT generation was significantly delayed in the regrowth assay. However, nucleation occurred at a normal frequency in steady state, suggesting the presence of a γ-tubulin-independent backup mechanism. Thus, endoplasmic MTs in this cell type are generated in a less ordered manner, showing a broader spectrum of nucleation mechanisms in plants.  相似文献   

10.
Yariv phenylglycosides [1,3,5-tri(p-glycosyloxyphenylazo)-2,4,6-trihydroxybenzene] are a group of chemical compounds that selectively bind to arabinogalactan proteins (AGPs), a type of plant proteoglycan. Yariv phenylglycosides are widely used as cytochemical reagents to perturb the molecular functions of AGPs as well as for the detection, quantification, purification, and staining of AGPs. However, the target structure in AGPs to which Yariv phenylglycosides bind has not been determined. Here, we identify the structural element of AGPs required for the interaction with Yariv phenylglycosides by stepwise trimming of the arabinogalactan moieties using combinations of specific glycoside hydrolases. Whereas the precipitation with Yariv phenylglycosides (Yariv reactivity) of radish (Raphanus sativus) root AGP was not reduced after enzyme treatment to remove α-l-arabinofuranosyl and β-glucuronosyl residues and β-1,6-galactan side chains, it was completely lost after degradation of the β-1,3-galactan main chains. In addition, Yariv reactivity of gum arabic, a commercial product of acacia (Acacia senegal) AGPs, increased rather than decreased during the repeated degradation of β-1,6-galactan side chains by Smith degradation. Among various oligosaccharides corresponding to partial structures of AGPs, β-1,3-galactooligosaccharides longer than β-1,3-galactoheptaose exhibited significant precipitation with Yariv in a radial diffusion assay on agar. A pull-down assay using oligosaccharides cross linked to hydrazine beads detected an interaction of β-1,3-galactooligosaccharides longer than β-1,3-galactopentaose with Yariv phenylglycoside. To the contrary, no interaction with Yariv was detected for β-1,6-galactooligosaccharides of any length. Therefore, we conclude that Yariv phenylglycosides should be considered specific binding reagents for β-1,3-galactan chains longer than five residues, and seven residues are sufficient for cross linking, leading to precipitation of the Yariv phenylglycosides.Arabinogalactan proteins (AGPs) are a type of plant proteoglycans consisting of a Hyp-rich core protein and large arabinogalactan (AG) moieties (Fincher et al., 1983; Nothnagel, 1997). Although there are many molecular species of AGP differentiated by their core proteins, the AG moieties commonly comprise β-1,3-galactan main chains and β-1,6-galactan side chains, to which l-Ara and other auxiliary sugars, such as GlcA, 4-O-methyl-GlcA, l-Fuc, l-Rha, and Xyl, are attached (Fincher et al., 1983; Nothnagel, 1997; Seifert and Roberts, 2007). A commercial product of AGPs prepared from the acacia (Acacia senegal) tree is known as gum arabic and utilized as a food stabilizer. In the Japanese herbal remedy Juzen-Taiho-To, AGs from Astragalus membranaceus are the active ingredient (Majewska-Sawka and Nothnagel, 2000; Kiyohara et al., 2002). In intact plants, AGPs are implicated in various physiological events and serve as extracellular constituents and signaling molecules. For instance, an AGP from stylar transmitting tissue attracts pollen tubes and stimulates their elongation in tobacco (Nicotiana tabacum; Cheung et al., 1995).Yariv phenylglycosides [1,3,5-tri(p-glycosyloxyphenylazo)-2,4,6-trihydroxybenzene] are a group of chemical compounds that were initially developed as carbohydrate antigens for the purification of anti-glycoside antibody and sugar-binding protein (Yariv et al., 1962, 1967a). It then turned out that Yariv phenylglycosides specifically precipitate AGPs (Yariv et al., 1967b; Jermyn and Yeow, 1975). The specific interaction of AGPs with Yariv phenylglycosides forming brown-red precipitate is called Yariv reactivity and has been recognized as an important criterion in the definition of AGPs, even though a number of AGPs do not exhibit Yariv reactivity. Nevertheless, the structure involved in the interaction with Yariv phenylglycoside is presumed to be conserved in many AGPs. The interaction of Yariv phenylglycosides with AGP depends on the glycosyl residues attached to the phenylazotrihydroxybenzene core. In particular, β-glucosyl Yariv phenylglycoside (β-Glc-Yariv) and β-galactosyl Yariv phenylglycoside (β-Gal-Yariv) bind to AGPs, whereas α-glucosyl Yariv and α-galactosyl Yariv (α-Gal-Yariv) do not bind to AGPs (Jermyn and Yeow, 1975; Larkin, 1977, 1978; Nothnagel and Lyon, 1986). Because of the specific interaction with the β-glycosyl Yariv phenylglycosides (β-Yarivs), AGPs were formerly called “β-lectins” (Jermyn and Yeow, 1975; Gleeson and Jermyn, 1979; Nothnagel and Lyon, 1986).The β-Yarivs are useful tools for staining, detection, and quantification of AGPs. Using β-Glc-Yariv, β-lectins were shown to exist in angiosperm, gymnosperm, fern, moss, and liverwort, illustrating the wide distribution of AGPs in the plant kingdom (Jermyn and Yeow, 1975; Clarke et al., 1978). In addition, β-Yarivs are also used as chemical reagents in the purification of AGPs. A nonclassical AGP, xylogen, which is a signaling molecule inducing the differentiation to tracheary elements, has been purified from the culture medium of zinnia (Zinnia elegans) cells by precipitation with β-Glc-Yariv (Motose et al., 2004). As the treatment with β-Yarivs causes the perturbation of various physiological processes in plants, β-Yarivs are reliable cytochemical reagents to explore AGP functions. Application of β-Yarivs to cultured cells of Arabidopsis (Arabidopsis thaliana) induced programmed cell death, demonstrating the involvement of AGPs in the determination of cell fate (Gao and Showalter, 1999). In tobacco cultured cells, the treatment with β-Yarivs has indicated a possible role of AGPs in the orientation of cortical microtubules and the polymerization of F-actin (Sardar et al., 2006).Although Yariv phenylglycosides have been extensively utilized in studies of AGPs over 40 years, the identification of the target structures on AGPs required for β-Yariv reactivity remains elusive (Nothnagel, 1997; Seifert and Roberts, 2007). It has been proposed that β-Yarivs bind to the Hyp-rich core protein, based on the observation that deglycosylation treatment with hydrogen fluoride did not abolish the Yariv reactivity of gum arabic and a tobacco AGP (Akiyama et al., 1984). To the contrary, other reports have asserted the importance of the carbohydrate moieties for Yariv reactivity (Komalavilas et al., 1991). However, with regard to the specific carbohydrate structure required for interaction with β-Yarivs, the results were not always consistent: neither α-l-arabinofuranosyl residues nor β-1,6-galactan side chains were found to be involved in Yariv reactivity of AGPs from Gladiolus spp., radish (Raphanus sativus), and grape (Vitis vinifera; Gleeson and Clarke, 1979; Tsumuraya et al., 1987; Saulnier et al., 1992); partial acid hydrolysis to remove α-l-arabinofuranosyl residues diminished Yariv reactivity of a rose (Rosa spp.) AGP (Komalavilas et al., 1991); and mugwort (Artemisia vulgaris) pollen O-glycans consisting of a β-1,6-galactan core and branched α-l-arabinofuranosyl side chains precipitated with β-Glc-Yariv (Léonard et al., 2005). Accordingly, it has also been suggested that Yariv reactivity depends on the overall physical and chemical properties rather than a specific structural feature of AGPs.In this study, we demonstrate that the peptide component of AGPs is not required for Yariv reactivity. By sequentially trimming the AG moieties of AGPs with sets of specific glycoside hydrolases, we show that β-Gal-Yariv binds to the β-1,3-galactan main chains of radish root AGP. We confirm that β-1,6-galactan side chains are not necessary for Yariv reactivity, we identify β-1,3-galactopentaose (β-1,3-Gal5) as the smallest carbohydrate structure to interact with β-Gal-Yariv, and we show that β-1,3-galactoheptaose (β-1,3-Gal7) or longer β-1,3-galactosyl chains are required for the formation of insoluble precipitate with Yariv phenylglycoside. Based on computational modeling, a possible interaction mechanism between β-Gal-Yariv and β-1,3-galactan is suggested.  相似文献   

11.
Plants employ acentrosomal mechanisms to organize cortical microtubule arrays essential for cell growth and differentiation. Using structured illumination microscopy (SIM) adopted for the optimal documentation of Arabidopsis (Arabidopsis thaliana) hypocotyl epidermal cells, dynamic cortical microtubules labeled with green fluorescent protein fused to the microtubule-binding domain of the mammalian microtubule-associated protein MAP4 and with green fluorescent protein-fused to the alpha tubulin6 were comparatively recorded in wild-type Arabidopsis plants and in the mitogen-activated protein kinase mutant mpk4 possessing the former microtubule marker. The mpk4 mutant exhibits extensive microtubule bundling, due to increased abundance and reduced phosphorylation of the microtubule-associated protein MAP65-1, thus providing a very useful genetic tool to record intrabundle microtubule dynamics at the subdiffraction level. SIM imaging revealed nano-sized defects in microtubule bundling, spatially resolved microtubule branching and release, and finally allowed the quantification of individual microtubules within cortical bundles. Time-lapse SIM imaging allowed the visualization of subdiffraction, short-lived excursions of the microtubule plus end, and dynamic instability behavior of both ends during free, intrabundle, or microtubule-templated microtubule growth and shrinkage. Finally, short, rigid, and nondynamic microtubule bundles in the mpk4 mutant were observed to glide along the parent microtubule in a tip-wise manner. In conclusion, this study demonstrates the potential of SIM for superresolution time-lapse imaging of plant cells, showing unprecedented details accompanying microtubule dynamic organization.Plant cell growth and differentiation depend on dynamic cortical microtubule organization mechanisms (Ehrhardt, 2008). Such mechanisms include branched microtubule formation and release (Murata et al., 2005; Nakamura et al., 2010; Fishel and Dixit, 2013), microtubule-templated microtubule growth (Chan et al., 2009), angle-of-contact microtubule bundling or catastrophe induction (Dixit and Cyr, 2004; Tulin et al., 2012), severing at microtubule crossovers (Wightman and Turner, 2007), and unique dynamic behavior between steady-state treadmilling and dynamic instability (Shaw et al., 2003).Cortical microtubule dynamics have been studied in vivo and in vitro with total internal reflection microscopy (TIRFM; Vizcay-Barrena et al., 2011), variable-angle emission microscopy (VAEM; Wan et al., 2011), spinning-disc microscopy (SD; Shaw and Lucas, 2011), and confocal laser scanning microscopy (CLSM; Shaw et al., 2003). TIRFM and VAEM provide sufficient resolution and speed but at limited depth of imaging (approximately 200 nm; Martin-Fernandez et al., 2013) and inevitably a very narrow field of view when used for in vivo studies (Mattheyses et al., 2010). Dynamic CLSM imaging suffers from field-of-view limitations while also introducing phototoxicity to the imaged sample. Furthermore, CLSM is based on a speed-to-resolution tradeoff that will necessitate computational extrapolation to bring resolution to affordable levels (Rosero et al., 2014). Finally, SD can provide sufficient depth and speed but otherwise poor resolution, owing to aberrations arising from the sample and the properties of the optics commonly used (Shaw and Ehrhardt, 2013).Microtubule research evolved concomitant with optical microscopy and the development of fluorescent proteins markers, allowing the resolution of microtubule dynamics and organization at video rates (Marc et al., 1998; Shaw and Ehrhardt, 2013). However, the bulk of plant cells organized in tissues and the optical properties of cell walls hamper microscopic observations, so that the delineation of fine details of microtubule organization still relies on laborious transmission electron microscopy (Kang, 2010).Alternatively, in vitro assays using total internal reflection (TIRFM) or Allen’s video-enhanced contrast-differential interference contrast microscopy (Allen et al., 1981) with purified components have advanced the understanding of microtubule-microtubule-associated protein (MAP) interactions while providing mechanistic insight on the function of MAP65 proteins (Tulin et al., 2012; Portran et al., 2013; Stoppin-Mellet et al., 2013), kinesin motors (Song et al., 1997), katanin-mediated microtubule severing (Stoppin-Mellet et al., 2007), and microtubule dynamics (Moore et al., 1997). However, it is explicitly acknowledged that such in vitro assays should be addressed in biologically coherent systems with physiological relevance to microtubule dynamics (Gardner et al., 2013; Zanic et al., 2013). Thus, an ideal approach would be to address microtubule dynamics in the complex cellular environment at spatiotemporal resolutions achieved by in vitro assays.Subdiffraction optical microscopy techniques allow subcellular observations below Abbe’s resolution threshold (Verdaasdonk et al., 2014), complementing the use of transmission electron microscopy. Such approaches permit dynamic subcellular tracking of appropriately tagged structures within living cells (Tiwari and Nagai, 2013). Practically, two superresolution strategies exist. The first involves patterned light illumination, allowing superresolution acquisitions by two fundamentally different methods, stimulated emission depletion (STED; Hell, 2007) and structured illumination microscopy (SIM; Gustafsson, 2000). The second interrogates the precision of fluorophore localization and includes stochastic optical reconstruction microscopy (STORM; Kamiyama and Huang, 2012) and photoactivation localization microscopy (PALM; Sengupta et al., 2012). The above regimes differ in translational and axial resolution, and their temporal efficiency depends on the size of the imaged area. SIM is probably the best compromise for superresolution live imaging, as it offers reasonable lateral resolution (approximately 100 nm; Gustafsson, 2000), which may be reduced to 50 nm (Rego et al., 2012), and sufficient depth of imaging combined with a reasonable axial resolution (approximately 200 nm). SIM allows dynamic imaging in a broader field of view than STED, at biologically meaningful rates compared with PALM and STORM (Kner et al., 2009), and with deeper imaging capacity compared with other superresolution regimes and with TIRFM/VAEM (Leung and Chou, 2011). Superresolution approaches have received limited attention in the plant cell biology field (Fitzgibbon et al., 2010; Kleine-Vehn et al., 2011), and their resolution potential during live imaging was not quantified previously.Here, high-numerical aperture (NA) objectives were combined with SIM for the acquisition and systematic quantification of subdiffraction details of cortical microtubules labeled either with GFP fused to the microtubule-binding domain of mammalian MAP4 (GFP-MBD; Marc et al., 1998) or with GFP fused to alpha tubulin6 (GFP-TUA6; Shaw et al., 2003). For such studies, wild-type plants and a mitogen-activated protein kinase4 (mpk4) mutant, exhibiting extensive microtubule bundling due to the overexpression and underphosphorylation of MAP65-1 (Beck et al., 2010), were used.  相似文献   

12.
13.
The actin cytoskeleton is a major regulator of cell morphogenesis and responses to biotic and abiotic stimuli. The organization and activities of the cytoskeleton are choreographed by hundreds of accessory proteins. Many actin-binding proteins are thought to be stimulus-response regulators that bind to signaling phospholipids and change their activity upon lipid binding. Whether these proteins associate with and/or are regulated by signaling lipids in plant cells remains poorly understood. Heterodimeric capping protein (CP) is a conserved and ubiquitous regulator of actin dynamics. It binds to the barbed end of filaments with high affinity and modulates filament assembly and disassembly reactions in vitro. Direct interaction of CP with phospholipids, including phosphatidic acid, results in uncapping of filament ends in vitro. Live-cell imaging and reverse-genetic analyses of cp mutants in Arabidopsis (Arabidopsis thaliana) recently provided compelling support for a model in which CP activity is negatively regulated by phosphatidic acid in vivo. Here, we used complementary biochemical, subcellular fractionation, and immunofluorescence microscopy approaches to elucidate CP-membrane association. We found that CP is moderately abundant in Arabidopsis tissues and present in a microsomal membrane fraction. Sucrose density gradient separation and immunoblotting with known compartment markers were used to demonstrate that CP is enriched on membrane-bound organelles such as the endoplasmic reticulum and Golgi. This association could facilitate cross talk between the actin cytoskeleton and a wide spectrum of essential cellular functions such as organelle motility and signal transduction.The cellular levels of membrane-associated lipids undergo dynamic changes in response to developmental and environmental stimuli. Different species of phospholipids target specific proteins and this often affects the activity and/or subcellular localization of these lipid-binding proteins. One such membrane lipid, phosphatidic acid (PA), serves as a second messenger and regulates multiple developmental processes in plants, including seedling development, root hair growth and pattern formation, pollen tube growth, leaf senescence, and fruit ripening. PA levels also change during various stress responses, including high salinity and dehydration, pathogen attack, and cold tolerance (Testerink and Munnik, 2005, 2011; Wang, 2005; Li et al., 2009). In mammalian cells, PA is critical for vesicle trafficking events, such as vesicle budding from the Golgi apparatus, vesicle transport, exocytosis, endocytosis, and vesicle fusion (Liscovitch et al., 2000; Freyberg et al., 2003; Jenkins and Frohman, 2005).The actin cytoskeleton and a plethora of actin-binding proteins (ABPs) are well-known targets and transducers of lipid signaling (Drøbak et al., 2004; Saarikangas et al., 2010; Pleskot et al., 2013). For example, several ABPs have the ability to bind phosphoinositide lipids, such as phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2]. The severing or actin filament depolymerizing proteins such as villin, cofilin, and profilin are inhibited when bound to PtdIns(4,5)P2. One ABP appears to be strongly regulated by another phospholipid; human gelsolin binds to lysophosphatidic acid and its filament severing and barbed-end capping activities are inhibited by this biologically active lipid (Meerschaert et al., 1998). Gelsolin is not, however, regulated by PA (Meerschaert et al., 1998), nor are profilin (Lassing and Lindberg, 1985), α-actinin (Fraley et al., 2003), or chicken CapZ (Schafer et al., 1996).The heterodimeric capping protein (CP) from Arabidopsis (Arabidopsis thaliana) also binds to and its activity is inhibited by phospholipids, including both PtdIns(4,5)P2 and PA (Huang et al., 2003, 2006). PA and phospholipase D activity have been implicated in the actin-dependent tip growth of root hairs and pollen tubes (Ohashi et al., 2003; Potocký et al., 2003; Samaj et al., 2004; Monteiro et al., 2005a; Pleskot et al., 2010). Exogenous application of PA causes an elevation of actin filament levels in suspension cells, pollen, and Arabidopsis epidermal cells (Lee et al., 2003; Potocký et al., 2003; Huang et al., 2006; Li et al., 2012; Pleskot et al., 2013). Capping protein (CP) binds to the barbed end of actin filaments with high (nanomolar) affinity, dissociates quite slowly, and prevents the addition of actin subunits at this end (Huang et al., 2003, 2006; Kim et al., 2007). In the presence of phospholipids, AtCP is not able to bind to the barbed end of actin filaments (Huang et al., 2003, 2006). Furthermore, capped filament ends are uncapped by the addition of PA, allowing actin assembly from a pool of profilin-actin (Huang et al., 2006). Collectively, these data lead to a simple model whereby CP, working in concert with profilin-actin, serves to maintain tight regulation of actin assembly at filament barbed ends (Huang et al., 2006; Blanchoin et al., 2010; Henty-Ridilla et al., 2013; Pleskot et al., 2013). Furthermore, the availability of CP for filament ends can be modulated by fluxes in signaling lipids. Genetic evidence for this model was recently obtained by analyzing the dynamic behavior of actin filament ends in living Arabidopsis epidermal cells after treatment with exogenous PA (Li et al., 2012). Specifically, changes in the architecture of cortical actin arrays and dynamics of individual actin filaments that are induced by PA treatment were found to be attenuated in cp mutant cells (Li et al., 2012; Pleskot et al., 2013).Structural characterization of chicken CapZ demonstrates that the α- and β-subunits of the heterodimer form a compact structure resembling a mushroom with pseudo-two-fold rotational symmetry (Yamashita et al., 2003). Actin- and phospholipid-binding sites are conserved on the C-terminal regions, sometimes referred to as tentacles, which comprise amphipathic α-helices (Cooper and Sept, 2008; Pleskot et al., 2012). Coarse-grained molecular dynamics (CG-MD) simulations recently revealed the mechanism of chicken and AtCP association with membranes (Pleskot et al., 2012). AtCP interacts specifically with lipid bilayers through interactions between PA and the amphipathic helix of the α-subunit tentacle. Extensive polar contacts between lipid headgroups and basic residues on CP (including K278, which is unique to plant CP), as well as partial embedding of nonpolar groups into the lipid bilayer, are observed (Pleskot et al., 2012). Moreover, a glutathione S-transferase fusion protein containing the C-terminal 38 amino acids from capping protein α subunit (CPA) is sufficient to bind PA-containing liposomes in vitro (Pleskot et al., 2012). Collectively, these findings lead us to predict that AtCP will behave like a membrane-associated protein in plant cells.Additional evidence from animal and microbial cells supports the association of CP with biological membranes. In Acanthamoeba castellanii, CP is localized primarily to the hyaline ectoplasm in a region of the cytoplasm just under the plasma membrane that contains a high concentration of actin filaments (Cooper et al., 1984). Localization of CP with regions rich in actin filaments and with membranes was supported by subcellular fractionation experiments, in which CP was associated with a crude membrane fraction that included plasma membrane (Cooper et al., 1984). Further evidence demonstrates that CP localizes to cortical actin patches at sites of new cell wall growth in budding yeast (Saccharomyces cerevisiae), including the site of bud emergence. By contrast, CP did not colocalize with actin cables in S. cerevisiae (Amatruda and Cooper, 1992). CP may localize to these sites by direct interactions with membrane lipids, through binding the ends of actin filaments, or by association with another protein different from actin. In support of this hypothesis, GFP-CP fusion proteins demonstrate that sites of actin assembling in living cells contain both CP and the actin-related protein2/3 (Arp2/3) complex, and CP is located in two types of structures: (1) motile regions of the cell periphery, which reflect movement of the edge of the lamella during extension and ruffling; and (2) dynamic spots within the lamella (Schafer et al., 1998). CP has been colocalized to the F-actin patches in fission yeast (Schizosaccharomyces pombe; Kovar et al., 2005), which promotes Arp2/3-dependent nucleation and branching and limits the extent of filament elongation (Akin and Mullins, 2008). These findings lend additional support for a model whereby CP cooperates with the Arp2/3 complex to regulate actin dynamics (Nakano and Mabuchi, 2006). Activities and localization of other plant ABPs are linked to membranes. Membrane association has been linked to the assembly status of the ARP2/3 complex, an actin filament nucleator, in Arabidopsis (Kotchoni et al., 2009). SPIKE1 (SPK1), a Rho of plants (Rop)-guanine nucleotide exchange factor (GEF) and peripheral membrane protein, maintains the homeostasis of the early secretory pathway and signal integration during morphogenesis through specialized domains in the endoplasmic reticulum (ER; Zhang et al., 2010). Furthermore, Nck-associated protein1 (NAP1), a component of the suppressor of cAMP receptor/WASP-family verprolin homology protein (SCAR/WAVE) complex, strongly associates with membranes and is particularly enriched in ER membranes (Zhang et al., 2013a). Finally, a superfamily of plant ABPs, called NETWORKED proteins, was recently discovered; these link the actin cytoskeleton to various cellular membranes (Deeks et al., 2012; Hawkins et al., 2014; Wang et al., 2014).In this work, we demonstrate that CP is a membrane-associated protein in Arabidopsis. To our knowledge, this is the first direct evidence for CP-membrane association in plants. This interaction likely targets CP to cellular compartments such as the ER and Golgi. This unique location may allow CP to remodel the actin cytoskeleton in the vicinity of endomembrane compartments and/or to respond rapidly to fluxes in signaling lipids.  相似文献   

14.
15.
Abscisic acid (ABA) is a key plant hormone involved in diverse physiological and developmental processes, including abiotic stress responses and the regulation of stomatal aperture and seed germination. Abscisic acid glucosyl ester (ABA-GE) is a hydrolyzable ABA conjugate that accumulates in the vacuole and presumably also in the endoplasmic reticulum. Deconjugation of ABA-GE by the endoplasmic reticulum and vacuolar β-glucosidases allows the rapid formation of free ABA in response to abiotic stress conditions such as dehydration and salt stress. ABA-GE further contributes to the maintenance of ABA homeostasis, as it is the major ABA catabolite exported from the cytosol. In this work, we identified that the import of ABA-GE into vacuoles isolated from Arabidopsis (Arabidopsis thaliana) mesophyll cells is mediated by two distinct membrane transport mechanisms: proton gradient-driven and ATP-binding cassette (ABC) transporters. Both systems have similar Km values of approximately 1 mm. According to our estimations, this low affinity appears nevertheless to be sufficient for the continuous vacuolar sequestration of ABA-GE produced in the cytosol. We further demonstrate that two tested multispecific vacuolar ABCC-type ABC transporters from Arabidopsis exhibit ABA-GE transport activity when expressed in yeast (Saccharomyces cerevisiae), which also supports the involvement of ABC transporters in ABA-GE uptake. Our findings suggest that the vacuolar ABA-GE uptake is not mediated by specific, but rather by several, possibly multispecific, transporters that are involved in the general vacuolar sequestration of conjugated metabolites.Abscisic acid (ABA) is a major plant hormone involved in various physiological and developmental processes. ABA signaling is fundamental in plant responses to abiotic stresses, including drought, cold, osmotic, and salt stress (Cutler et al., 2010). The best-characterized function of ABA is the regulation of stomatal aperture in response to environmental signals, such as soil and air humidity, temperature, and CO2 concentration (Nilson and Assmann, 2007; Kim et al., 2010). However, ABA also has important functions in seed development, dormancy, and germination (Holdsworth et al., 2008), lateral root formation (Galvan-Ampudia and Testerink, 2011), and leaf senescence (Lim et al., 2007). Besides, ABA is not restricted only to plants; it was also identified to have functions in species from all kingdoms, including humans, and may even have universal functions (e.g. in UV-B stress response; Tossi et al., 2012).ABA is synthesized de novo from the carotenoid zeaxanthin, whereby the first ABA-specific biosynthetic step occurs in the plastid and the final two steps take place in the cytosol (Nambara and Marion-Poll, 2005). The catabolism of ABA is mediated via oxidative and Glc conjugation pathways (Nambara and Marion-Poll, 2005). The ABA 8′-hydroxylation catalyzed by P450 cytochromes of the CYP707A subfamily represents the predominant catabolic pathway of ABA and has been demonstrated to be a key regulatory step in ABA action (Kushiro et al., 2004). The major oxidative ABA catabolites, phaseic acid (PA) and dihydroxyphaseic acid (DPA), exhibit lower and no biological activity, respectively (Sharkey and Raschke, 1980; Kepka et al., 2011). The conjugation of ABA and its oxidative catabolites PA and DPA with Glc catalyzed by UDP-glucosyltransferases represents the other mechanism of ABA inactivation. Abscisic acid glucosyl ester (ABA-GE) appears to be the major conjugate, which was found in various organs of different plant species (Piotrowska and Bajguz, 2011). In contrast to the oxidative pathway, the inactivation of ABA by Glc conjugation is reversible, and hydrolysis of ABA-GE catalyzed by β-glucosidases results in free ABA (Dietz et al., 2000; Lee et al., 2006; Xu et al., 2012). ABA-GE levels were shown to substantially increase during dehydration and specific seed developmental and germination stages (Boyer and Zeevaart, 1982; Hocher et al., 1991; Chiwocha et al., 2003). Furthermore, ABA-GE is present in the xylem sap, where it was shown to increase under drought, salt, and osmotic stress (Sauter et al., 2002). Apoplastic ABA β-glucosidases in leaves have been suggested to mediate the release of free ABA from xylem-borne ABA-GE (Dietz et al., 2000). Therefore, ABA-GE was proposed to be a root-to-shoot signaling molecule. However, under drought stress, ABA-mediated stomatal closure occurs independently of root ABA biosynthesis (Christmann et al., 2007). Thus, the involvement of ABA-GE in root-to-shoot signaling of water stress conditions remains to be revealed (Goodger and Schachtman, 2010).The intracellular compartmentalization of ABA and its catabolites is important for ABA homeostasis (Xu et al., 2013). Free ABA, PA, and DPA mainly occur in the extravacuolar compartments. In contrast to these oxidative ABA catabolites, ABA-GE has been reported to accumulate in vacuoles (Bray and Zeevaart, 1985; Lehmann and Glund, 1986). Since the sequestered ABA-GE can instantaneously provide ABA via a one-step hydrolysis, this conjugate and its compartmentalization may be of importance in the maintenance of ABA homeostasis. The identification of the endoplasmic reticulum (ER)-localized β-glucosidase AtBG1 that specifically hydrolyzes ABA-GE suggests that ABA-GE is also present in the ER (Lee et al., 2006). Plants lacking functional AtBG1 exhibit pronounced ABA-deficiency phenotypes, including sensitivity to dehydration, impaired stomatal closure, earlier germination, and lower ABA levels. Hydrolysis of ER-localized ABA-GE, therefore, represents an alternative pathway for the generation of free cytosolic ABA (Lee et al., 2006; Bauer et al., 2013). This finding raised the question of whether vacuolar ABA-GE also has an important function as an ABA reservoir. This hypothesis was supported by recent identifications of two vacuolar β-glucosidases that hydrolyze vacuolar ABA-GE (Wang et al., 2011; Xu et al., 2013). The vacuolar AtBG1 homolog AtBG2 forms high molecular weight complexes, which are present at low levels under normal conditions but significantly accumulate under dehydration stress. AtBG2 knockout plants displayed a similar, although less pronounced, phenotype to AtBG1 mutants: elevated sensitivity to drought and salt stress, while overexpression of AtBG2 resulted in exactly the opposite effect (i.e. increased drought tolerance). The other identified vacuolar ABA-GE glucosidase, BGLU10, exhibits comparable mutant phenotypes to AtBG2 (Wang et al., 2011). This redundancy may explain the less pronounced mutant phenotypes of vacuolar ABA-GE glucosidases compared with the ER-localized AtBG1. Moreover, the fact that overexpression of the vacuolar AtBG2 is able to phenotypically complement AtBG1 deletion mutants indicates an important role of vacuolar ABA-GE as a pool for free ABA during the abiotic stress response (Xu et al., 2012).The described accumulation and functions of vacuolar ABA-GE raise the question of by which mechanisms ABA-GE is sequestered into the vacuoles. To answer this question, we synthesized radiolabeled ABA-GE and characterized the ABA-GE transport into isolated mesophyll vacuoles. We showed that the vacuole comprises two distinct transport systems involved in the accumulation of ABA-GE: proton gradient-dependent and directly energized ATP-binding cassette (ABC)-type transport. In a targeted approach, we furthermore show that the Arabidopsis (Arabidopsis thaliana) ABC transporters AtABCC1 and AtABCC2 exhibit ABA-GE transport activity in vitro.  相似文献   

16.
Proteins decorated with arabinogalactan (AG) have important roles in cell wall structure and plant development, yet the structure and biosynthesis of this polysaccharide are poorly understood. To facilitate the analysis of biosynthetic mutants, water-extractable arabinogalactan proteins (AGPs) were isolated from the leaves of Arabidopsis (Arabidopsis thaliana) plants and the structure of the AG carbohydrate component was studied. Enzymes able to hydrolyze specifically AG were utilized to release AG oligosaccharides. The released oligosaccharides were characterized by high-energy matrix-assisted laser desorption ionization-collision-induced dissociation mass spectrometry and polysaccharide analysis by carbohydrate gel electrophoresis. The Arabidopsis AG is composed of a β-(1→3)-galactan backbone with β-(1→6)-d-galactan side chains. The β-(1→6)-galactan side chains vary in length from one to over 20 galactosyl residues, and they are partly substituted with single α-(1→3)-l-arabinofuranosyl residues. Additionally, a substantial proportion of the β-(1→6)-galactan side chain oligosaccharides are substituted at the nonreducing termini with single 4-O-methyl-glucuronosyl residues via β-(1→6)-linkages. The β-(1→6)-galactan side chains are occasionally substituted with α-l-fucosyl. In the fucose-deficient murus1 mutant, AGPs lack these fucose modifications. This work demonstrates that Arabidopsis mutants in AGP structure can be identified and characterized. The detailed structural elucidation of the AG polysaccharides from the leaves of Arabidopsis is essential for insights into the structure-function relationships of these molecules and will assist studies on their biosynthesis.Arabinogalactans (AGs) are structurally complex large-branched polysaccharides attached to Hyp residues of many plant cell wall polypeptides. Most proteins glycosylated with AGs (AGPs) have both AG glycosylated domains (glycomodules) and structural or enzymatic domains. However, typical AGPs commonly contain less than 10% protein, suggesting that the AG is the functional part of the molecule (Clarke et al., 1979; Fincher et al., 1983; Kieliszewski and Lamport, 1994; Borner et al., 2003; Xu et al., 2008). Hyp is the most characteristic amino acid present at the glycosylated domain of the AGP, but other amino acids such as Ser, Ala, and Thr are also very common. Type II AG polysaccharides share common structural features based on a β-(1→3)-galactan backbone with β-(1→6)-linked galactan side chains and can be found both on AGPs and rhamnogalacturonan-I (RG-I) pectin (Renard et al., 1991). The galactopyranosyl (Galp) residues can be further substituted with l-arabinofuranosyl (l-Araf) and occasionally also l-rhamnosyl (l-Rha), l-fucosyl (l-Fuc), and glucuronosyl (GlcA; with or without 4-O-methylation) residues (Tsumuraya et al., 1988; Tan et al., 2004; Tryfona et al., 2010). (Sugars mentioned in this work belong to the D-series unless otherwise stated.)The structure of AGs is poorly characterized, and this is mainly due to the great heterogeneity of glycan structures, not only between different AGPs but also even on the same peptide sequence in the same tissue (Estévez et al., 2006). The glycan structure can also be different depending on the developmental stage and tissue type (Tsumuraya et al., 1988), adding to the great heterogeneity of these molecules and therefore limiting their detailed characterization. Molecular and biochemical evidence has indicated that AGPs have specific functions during root formation, promotion of somatic embryogenesis (van Hengel et al., 2002), and attraction of pollen tubes to the style (Cheung et al., 1995). In addition, enhanced secretion efficiency or stability in the cell wall are properties that the AG may confer on the glycosylated protein (Borner et al., 2003). However, it has been difficult to differentiate one species of AGP from another in plant tissues and to assign specific roles to individual AGPs.l-Fuc is present in AGPs in Arabidopsis (Arabidopsis thaliana; van Hengel et al., 2002), radish (Raphanus sativus; Nakamura et al., 1984; Tsumuraya et al., 1984a, 1984b, 1988), and several other dicot plants such as thyme (Thymus vulgaris; Chun et al., 2001) and celery (Apium graveolens; Lin et al., 2011). Reduction in l-Fuc by 40% in roots of murus1 (mur1) plants resulted in a decrease of 50% in root cell elongation, and eel lectin binding assays suggested that the phenotype was the result of alterations in the composition of root AGPs (van Hengel and Roberts, 2002). An α-(1→2)-fucosyltransferase (FUT) activity for radish primary root AGPs has been described, where an α-l-Araf-(1→3)-β-Galp-(1→6)-Galp trisaccharide was used as exogenous substrate acceptor to mimic an AG polysaccharide in the enzymatic assay (Misawa et al., 1996). Linkage analysis, reactivity with eel lectin, and digestion with α-(1→2)-fucosidase indicated that the l-Fuc residues added are terminal and attached via an α-linkage to the C-2 position of an adjacent l-Araf residue (Nakamura et al., 1984; Tsumuraya et al., 1984a, 1984b, 1988). Recently, Wu et al. (2010) identified AtFUT4 and AtFUT6 genes encoding FUT proteins specific to AGPs, but the structures of the fucosylated AG generated have not been fully characterized.To gain insights into the synthesis and function of plant AGPs, it would be useful to have mutants altered in their carbohydrate moieties. However, no AG-specific biosynthetic mutants have been characterized, and this, among other reasons, is due to the very limited knowledge of the structure of Arabidopsis AGs (Qu et al., 2008). Moreover, characterization of AG in candidate mutants remains challenging. Even though the structures of some AGs have been proposed using NMR and sugar linkage analyses, the complete structural elucidation of a native AG still remains a formidable task, because NMR spectroscopy and methylation analysis have been largely used to provide information regarding the amount and type of linkages between adjacent glycosyl residues, and AG heterogeneity can confound attempts to build complete structural models. Recently, a modular structure was proposed for AGs on heterologously expressed proteins in tobacco (Nicotiana tabacum; Tan et al., 2010). Tan et al. (2010) proposed that approximately 15-residue repeating blocks of decorated β-(1→3)-trigalactosyl subunits connected by β-(1→6)-linkages were the building blocks of type II AG polysaccharides and concluded that these molecules are far less complex than commonly supposed. Most characterized β-(1→6)-galactan side chains in AGs are reported to be short, of one or two residues (Neukom and Markwalder, 1975; Gane et al., 1995; Gaspar et al., 2001). On the contrary, there are reports of long β-(1→6)-galactan side chains in radish root AGPs (Haque et al., 2005). Similarly, we recently found evidence that wheat (Triticum aestivum) flour endosperm AGP extracts contained long β-(1→6)-galactan side chains heavily substituted with l-Araf at C-3 (Tryfona et al., 2010). This partial structure of the carbohydrate component of wheat flour AGP isolated from water extracts of wheat endosperm was elucidated utilizing a combination of analytical approaches, such as the use of enzymes able to release oligosaccharides specifically from AGs, high-energy matrix-assisted laser desorption ionization (MALDI)-collision-induced dissociation (CID) mass spectrometry (MS), and polysaccharide analysis by carbohydrate gel electrophoresis (PACE; Tryfona et al., 2010). In this work, we applied these techniques to study the carbohydrate component of Arabidopsis leaf AGPs. AG-specific enzyme digestion products were analyzed by PACE and MS, allowing a partial structure to be proposed. We show that endogenous Arabidopsis leaf AG is composed of a β-(1→3)-galactan backbone with β-(1→6)-galactan side chains. These side chains are substituted with l-Araf residues via α-(1→3)-linkages and can vary in length from one up to at least 20 Galp residues. We also found that the β-(1→6)-galactan side chains are substituted mainly with 4-O-methyl-glucuronosyl (4-O-Me-GlcA) at their nonreducing termini, while occasional l-Fuc substitutions were also present via α-(1→2)-linkages on l-Araf residues. In addition, AG oligosaccharides from leaves of the mur1 mutant were identified, and their structures were compared with those isolated from wild-type plants.  相似文献   

17.
Heterotrimeric G proteins have been shown to transmit ultraviolet B (UV-B) signals in mammalian cells, but whether they also transmit UV-B signals in plant cells is not clear. In this paper, we report that 0.5 W m−2 UV-B induces stomatal closure in Arabidopsis (Arabidopsis thaliana) by eliciting a cascade of intracellular signaling events including Gα protein, hydrogen peroxide (H2O2), and nitric oxide (NO). UV-B triggered a significant increase in H2O2 or NO levels associated with stomatal closure in the wild type, but these effects were abolished in the single and double mutants of AtrbohD and AtrbohF or in the Nia1 mutants, respectively. Furthermore, we found that UV-B-mediated H2O2 and NO generation are regulated by GPA1, the Gα-subunit of heterotrimeric G proteins. UV-B-dependent H2O2 and NO accumulation were nullified in gpa1 knockout mutants but enhanced by overexpression of a constitutively active form of GPA1 (cGα). In addition, exogenously applied H2O2 or NO rescued the defect in UV-B-mediated stomatal closure in gpa1 mutants, whereas cGα AtrbohD/AtrbohF and cGα nia1 constructs exhibited a similar response to AtrbohD/AtrbohF and Nia1, respectively. Finally, we demonstrated that Gα activation of NO production depends on H2O2. The mutants of AtrbohD and AtrbohF had impaired NO generation in response to UV-B, but UV-B-induced H2O2 accumulation was not impaired in Nia1. Moreover, exogenously applied NO rescued the defect in UV-B-mediated stomatal closure in the mutants of AtrbohD and AtrbohF. These findings establish a signaling pathway leading to UV-B-induced stomatal closure that involves GPA1-dependent activation of H2O2 production and subsequent Nia1-dependent NO accumulation.Heterotrimeric G proteins, composed of α-, β-, and γ-subunits, are a key intracellular signaling molecule in both mammalian and plant systems. Classically, upon signal reception by a receptor coupled to the heterotrimer, the Gα-subunit separates from the Gβγ dimer, and either Gα or the Gβγ dimer can act as a functional unit and induce downstream signaling (Oldham and Hamm, 2008). In contrast to mammalian cells, where multiple α, β, and γ genes exist, there is only one prototypical Gα (GPA1), one Gβ (AGB1), and two known Gγ (AGG1 and AGG2) genes in Arabidopsis (Arabidopsis thaliana; Temple and Jones, 2007). Despite the comparative simplicity of players, G proteins have been shown to participate in multiple signaling pathways in Arabidopsis, including developmental processes, phytohormone responses, and responses to biotic and abiotic environmental signals such as pathogens, ozone, drought, and light (Assmann, 2005; Temple and Jones, 2007; Warpeha et al., 2007; Okamoto et al., 2009; Nilson and Assmann, 2010).Depletion of the stratospheric ozone layer results in increased levels of the sun’s UV-B radiation (280–315 nm) at the Earth’s surface. Although this influx of shortwave photons with high energy implies serious effects for all living organisms (Frohnmeyer and Staiger, 2003), UV-B is also a key environmental signal that initiates diverse responses in a range of organisms (Jansen and Bornman, 2012). Thus, understanding the mechanism of UV-B signal transduction in cells is very important. In recent years, significant progress has been made in identifying the molecular players and understanding the early mechanisms and functions of the UV-B perception and signaling pathway in plants. The perception of UV-B by UV RESISTANCE LOCUS8 (UVR8) followed by the interaction among UVR8, CONSTITUTIVELY PHOTOMORPHOGENIC1 (COP1), and ELONGATED HYPOCOTYL5 (HY5) has emerged as a primary mechanism of the UV-B response that is crucial for UV-B acclimation and tolerance (Rizzini et al., 2011; Christie et al., 2012; Heijde and Ulm, 2012; Jansen and Bornman, 2012). However, few of the molecular players involved in UV-B signal transduction are currently known. In mammalian cells, heterotrimeric G proteins have been shown to mediate various UV-B-induced cellular responses, such as secretion of heparin-binding epidermal growth factor (HB-EGF), activation of mitogen-activated protein kinases, cyclooxygenase2 expression, and apoptosis in human keratinocytes (Seo et al., 2004, 2007; Seo and Juhnn, 2010), suggesting that G proteins are important molecular players in UV-B signal transduction. However, at present, whether G proteins participate in the responses of plant cells to UV-B is not known.Stomata embedded in the epidermis of terrestrial plants are important for CO2 absorption and water transpiration and are possible points of entry for pathogens. Thus, the regulation of stomatal apertures is extremely important for the survival of plants. Phenotypic analyses of Arabidopsis mutants lacking the Gα- or Gβ-subunit show that these G proteins are involved in stomatal movement regulated by abscisic acid (ABA; Wang et al., 2001; Pandey and Assmann, 2004; Liu et al., 2007; Fan et al., 2008; Zhang et al., 2011), sphingosine-1-P (Coursol et al., 2003, 2005), phosphatidic acid (PA; Mishra et al., 2006), extracellular calmodulin (ExtCaM; Chen et al., 2004; Li et al., 2009), extracellular ATP (Hao et al., 2012), and the pathogen-associated molecular pattern flg22 (Zhang et al., 2008), suggesting that plant G proteins respond to various stimuli as key regulators of stomatal movement. On exposure to UV-B radiation, many plant species exhibit decreases in stomatal conductance and/or aperture under growth chamber, greenhouse, and field conditions (Musil and Wand, 1993; Nogués et al., 1999; Jansen and Noort, 2000). However, in some species, UV-B has been reported to induce either stomatal opening or stomatal closure, perhaps depending on the metabolic state of guard cells (Jansen and Noort, 2000). Furthermore, UV-B-inhibited photosynthesis is partially caused by stomatal limitation (He et al., 2004). Thus, understanding the mechanism of stomatal movement regulated by UV-B is extremely important for improving the resistance of plants to enhanced UV-B radiation, but, to date, it is poorly understood.Recently, compelling evidence emerged that hydrogen peroxide (H2O2) and nitric oxide (NO) function as signaling molecules in plants, mediating a range of responses to environmental stress including UV-B radiation (Neill et al., 2002; Qiao and Fan, 2008; Wilson et al., 2008). Increasing evidence also points to the role for H2O2 and NO as essential components in guard cell signaling. For example, both H2O2 and NO have been implicated in ABA-, salicylic acid (SA)-, ethylene-, ExtCaM-, and darkness-induced stomatal closure. Furthermore, several main cellular players in stomatal movement, such as mitogen-activated protein kinases, protein phosphatases, cytoskeleton, and ion channels, have already been identified as likely targets downstream of H2O2 or NO (Neill et al., 2008; Wang and Song, 2008; Huang et al., 2009; Li et al., 2009; Wilkins et al., 2011; Yemets et al., 2011). G protein signaling to the membrane-bound H2O2 synthetic enzyme, NADPH oxidase, has been implicated in the development of disease resistance and the apoptotic hypersensitive response in rice (Oryza sativa; Suharsono et al., 2002). Production of reactive oxygen species (ROS) in response to the air pollutant ozone is also impaired in a mutant lacking the Gα subunit (Joo et al., 2005). The heterotrimeric G proteins also participate in ROS metabolism in plant cells (Wei et al., 2008; Zhao et al., 2010). During stomatal movement, G proteins mediate H2O2 production induced by ABA (Zhang et al., 2011), ExtCaM (Chen et al., 2004; Li et al., 2009), and extracellular ATP (Hao et al., 2012) as well as NO production induced by ExtCaM in guard cells (Li et al., 2009). In addition, phospholipase Dα and its product PA, which interact with GPA1 during ABA inhibition of stomatal opening (Mishra et al., 2006), also promote ABA-induced ROS production (Zhang et al., 2009). These observations suggest that G proteins may be key regulators of H2O2 and NO production in plant cells, including guard cells. With regard to the stomatal movement regulated by UV-B radiation, our previous studies showed that H2O2 and NO generation are required for UV-B-induced stomatal closure (He et al., 2005, 2011a, 2011b). However, whether the UV-B-induced production of H2O2 and NO in guard cells is also regulated by G proteins remains unknown.In this study, we use Arabidopsis mutants (e.g. GPA1 null mutants gpa1-1 and gpa1-2; Nia1-2, Nia2-1, and Nia1-2/Nia2-5, which are defective in NO production; and AtrbohD, AtrbohF, and AtrbohD/AtrbohF, which are defective in producing H2O2) and pharmacological reagents to show that the G protein is involved in the regulation of UV-B-induced stomatal closure in Arabidopsis via sequential elucidation of H2O2 and NO, two key regulators of UV-B regulation of stomatal movements. Our results establish a linear signaling cascade in which the Gα protein transmits UV-B signals to elicit H2O2, which then elicits NO in guard cells to regulate UV-B-dependent stomatal closure.  相似文献   

18.
Xyloglucan constitutes most of the hemicellulose in eudicot primary cell walls and functions in cell wall structure and mechanics. Although Arabidopsis (Arabidopsis thaliana) xxt1 xxt2 mutants lacking detectable xyloglucan are viable, they display growth defects that are suggestive of alterations in wall integrity. To probe the mechanisms underlying these defects, we analyzed cellulose arrangement, microtubule patterning and dynamics, microtubule- and wall-integrity-related gene expression, and cellulose biosynthesis in xxt1 xxt2 plants. We found that cellulose is highly aligned in xxt1 xxt2 cell walls, that its three-dimensional distribution is altered, and that microtubule patterning and stability are aberrant in etiolated xxt1 xxt2 hypocotyls. We also found that the expression levels of microtubule-associated genes, such as MAP70-5 and CLASP, and receptor genes, such as HERK1 and WAK1, were changed in xxt1 xxt2 plants and that cellulose synthase motility is reduced in xxt1 xxt2 cells, corresponding with a reduction in cellulose content. Our results indicate that loss of xyloglucan affects both the stability of the microtubule cytoskeleton and the production and patterning of cellulose in primary cell walls. These findings establish, to our knowledge, new links between wall integrity, cytoskeletal dynamics, and wall synthesis in the regulation of plant morphogenesis.The primary walls of growing plant cells are largely constructed of cellulose and noncellulosic matrix polysaccharides that include hemicelluloses and pectins (Carpita and Gibeaut, 1993; Somerville et al., 2004; Cosgrove, 2005). Xyloglucan (XyG) is the most abundant hemicellulose in the primary walls of eudicots and is composed of a β-1,4-glucan backbone with side chains containing Xyl, Gal, and Fuc (Park and Cosgrove, 2015). XyG is synthesized in the Golgi apparatus before being secreted to the apoplast, and its biosynthesis requires several glycosyltransferases, including β-1,4-glucosyltransferase, α-1,6-xylosyltransferase, β-1,2-galactosyltransferase, and α-1,2-fucosyltransferase activities (Zabotina, 2012). Arabidopsis (Arabidopsis thaliana) XYLOGLUCAN XYLOSYLTRANSFERASE1 (XXT1) and XXT2 display xylosyltransferase activity in vitro (Faik et al., 2002; Cavalier and Keegstra, 2006), and strikingly, no XyG is detectable in the walls of xxt1 xxt2 double mutants (Cavalier et al., 2008; Park and Cosgrove, 2012a), suggesting that the activity of XXT1 and XXT2 are required for XyG synthesis, delivery, and/or stability.Much attention has been paid to the interactions between cellulose and XyG over the past 40 years. Currently, there are several hypotheses concerning the nature of these interactions (Park and Cosgrove, 2015). One possibility is that XyGs bind directly to cellulose microfibrils (CMFs). Recent data indicating that crystalline cellulose cores are surrounded with hemicelluloses support this hypothesis (Dick-Pérez et al., 2011). It is also possible that XyG acts as a spacer-molecule to prevent CMFs from aggregating in cell walls (Anderson et al., 2010) or as an adapter to link cellulose with other cell wall components, such as pectin (Cosgrove, 2005; Cavalier et al., 2008). XyG can be covalently linked to pectin (Thompson and Fry, 2000; Popper and Fry, 2005, 2008), and NMR data demonstrate that pectins and cellulose might interact to a greater extent than XyG and cellulose in native walls (Dick-Pérez et al., 2011). Alternative models exist for how XyG-cellulose interactions influence primary wall architecture and mechanics. One such model posits that XyG chains act as load-bearing tethers that bind to CMFs in primary cell walls to form a cellulose-XyG network (Carpita and Gibeaut, 1993; Pauly et al., 1999; Somerville et al., 2004; Cosgrove, 2005). However, results have been accumulating against this tethered network model, leading to an alternative model in which CMFs make direct contact, in some cases mediated by a monolayer of xyloglucan, at limited cell wall sites dubbed “biomechanical hotspots,” which are envisioned as the key sites of cell wall loosening during cell growth (Park and Cosgrove, 2012a; Wang et al., 2013; Park and Cosgrove, 2015). Further molecular, biochemical, and microscopy experiments are required to help distinguish which aspects of the load-bearing, spacer/plasticizer, and/or hotspot models most accurately describe the functions of XyG in primary walls.Cortical microtubules (MTs) direct CMF deposition by guiding cellulose synthase complexes in the plasma membrane (Baskin et al., 2004; Paredez et al., 2006; Emons et al., 2007; Sánchez-Rodriguez et al., 2012), and the patterned deposition of cellulose in the wall in turn can help determine plant cell anisotropic growth and morphogenesis (Baskin, 2005). Disruption of cortical MTs by oryzalin, a MT-depolymerizing drug, alters the alignment of CMFs, suggesting that MTs contribute to CMF organization (Baskin et al., 2004). CELLULOSE SYNTHASE (CESA) genes, including CESA1, CESA3, and CESA6, are required for normal CMF synthesis in primary cell walls (Kohorn et al., 2006; Desprez et al., 2007), and accessory proteins such as COBRA function in cellulose production (Lally et al., 2001). Live-cell imaging from double-labeled YFP-CESA6; CFP-ALPHA-1 TUBULIN (TUA1) Arabidopsis seedlings provides direct evidence that cortical MTs determine the trajectories of cellulose synthesis complexes (CSCs) and patterns of cellulose deposition (Paredez et al., 2006). Additionally, MT organization affects the rotation of cellulose synthase trajectories in the epidermal cells of Arabidopsis hypocotyls (Chan et al., 2010). Recently, additional evidence for direct guidance of CSCs by MTs has been provided by the identification of CSI1/POM2, which binds to both MTs and CESAs (Bringmann et al., 2012; Li et al., 2012). MICROTUBULE ORGANIZATION1 (MOR1) is essential for cortical MT organization (Whittington et al., 2001), but disruption of cortical MTs in the mor1 mutant does not greatly affect CMF organization (Sugimoto et al., 2003), and oryzalin treatment does not abolish CSC motility (Paredez et al., 2006).Conversely, the organization of cortical MTs can be affected by cellulose synthesis. Treatment with isoxaben, a cellulose synthesis inhibitor, results in disorganized cortical MTs in tobacco cells, suggesting that inhibition of cellulose synthesis affects MT organization (Fisher and Cyr, 1998), and treatment with 2,6-dichlorobenzonitrile, another cellulose synthesis inhibitor, alters MT organization in mor1 plants (Himmelspach et al., 2003). Cortical MT orientation in Arabidopsis roots is also altered in two cellulose synthesis-deficient mutants, CESA652-isx and kor1-3, suggesting that CSC activity can affect MT arrays (Paredez et al., 2008). Together, these results point to a bidirectional relationship between cellulose synthesis/patterning and MT organization.MTs influence plant organ morphology, but the detailed mechanisms by which they do so are incompletely understood. The dynamics and stability of cortical MTs are also affected by MT-associated proteins (MAPs). MAP18 is a MT destabilizing protein that depolymerizes MTs (Wang et al., 2007), MAP65-1 functions as a MT crosslinker, and MAP70-1 functions in MT assembly (Korolev et al., 2005; Lucas et al., 2011). MAP70-5 stabilizes existing MTs to maintain their length, and its overexpression induces right-handed helical growth (Korolev et al., 2007); likewise, MAP20 overexpression results in helical cell twisting (Rajangam et al., 2008). CLASP promotes microtubule stability, and its mutant is hypersensitive to microtubule-destabilizing drug oryzalin (Ambrose et al., 2007). KATANIN1 (KTN1) is a MT-severing protein that can sever MTs into short fragments and promote the formation of thick MT bundles that ultimately depolymerize (Stoppin-Mellet et al., 2006), and loss of KTN1 function results in reduced responses to mechanical stress (Uyttewaal et al., 2012). In general, cortical MT orientation responds to mechanical signals and can be altered by applying force directly to the shoot apical meristem (Hamant et al., 2008). The application of external mechanical pressure to Arabidopsis leaves also triggers MT bundling (Jacques et al., 2013). Kinesins, including KINESIN-13A (KIN-13A) and FRAGILE FIBER1 (FRA1), have been implicated in cell wall synthesis (Cheung and Wu, 2011; Fujikura et al., 2014). The identification of cell wall receptors and sensors is beginning to reveal how plant cell walls sense and respond to external signals (Humphrey et al., 2007; Ringli, 2010); some of them, such as FEI1, FEI2, THESEUS1 (THE1), FERONIA (FER), HERCULES RECEPTOR KINASE1 (HERK1), WALL ASSOCIATED KINASE1 (WAK1), WAK2, and WAK4, have been characterized (Lally et al., 2001; Decreux and Messiaen, 2005; Kohorn et al., 2006; Xu et al., 2008; Guo et al., 2009; Cheung and Wu, 2011). However, the relationships between wall integrity, cytoskeletal dynamics, and wall synthesis have not yet been fully elucidated.In this study, we analyzed CMF patterning, MT patterning and dynamics, and cellulose biosynthesis in the Arabidopsis xxt1 xxt2 double mutant that lacks detectable XyG and displays altered growth (Cavalier et al., 2008; Park and Cosgrove, 2012a). To investigate whether and how XyG deficiency affects the organization of CMFs and cortical MTs, we observed CMF patterning in xxt1 xxt2 mutants and Col (wild-type) controls using atomic force microscopy (AFM), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), and confocal microscopy (Hodick and Kutschera, 1992; Derbyshire et al., 2007; Anderson et al., 2010; Zhang et al., 2014). We also generated transgenic Col and xxt1 xxt2 lines expressing GFP-MAP4 (Marc et al., 1998) and GFP-CESA3 (Desprez et al., 2007), and analyzed MT arrays and cellulose synthesis using live-cell imaging. Our results show that the organization of CMFs is altered, that MTs in xxt1 xxt2 mutants are aberrantly organized and are more sensitive to external mechanical pressure and the MT-depolymerizing drug oryzalin, and that cellulose synthase motility and cellulose content are decreased in xxt1 xxt2 mutants. Furthermore, real-time quantitative RT-PCR measurements indicate that the enhanced sensitivity of cortical MTs to mechanical stress and oryzalin in xxt1 xxt2 plants might be due to altered expression of MT-stabilizing and wall receptor genes. Together, these data provide insights into the connections between the functions of XyG in wall assembly, the mechanical integrity of the cell wall, cytoskeleton-mediated cellular responses to deficiencies in wall biosynthesis, and cell and tissue morphogenesis.  相似文献   

19.
Triacylglycerol (TAG) metabolism is a key aspect of intracellular lipid homeostasis in yeast and mammals, but its role in vegetative tissues of plants remains poorly defined. We previously reported that PHOSPHOLIPID:DIACYLGLYCEROL ACYLTRANSFERASE1 (PDAT1) is crucial for diverting fatty acids (FAs) from membrane lipid synthesis to TAG and thereby protecting against FA-induced cell death in leaves. Here, we show that overexpression of PDAT1 enhances the turnover of FAs in leaf lipids. Using the trigalactosyldiacylglycerol1-1 (tgd1-1) mutant, which displays substantially enhanced PDAT1-mediated TAG synthesis, we demonstrate that disruption of SUGAR-DEPENDENT1 (SDP1) TAG lipase or PEROXISOMAL TRANSPORTER1 (PXA1) severely decreases FA turnover, leading to increases in leaf TAG accumulation, to 9% of dry weight, and in total leaf lipid, by 3-fold. The membrane lipid composition of tgd1-1 sdp1-4 and tgd1-1 pxa1-2 double mutants is altered, and their growth and development are compromised. We also show that two Arabidopsis thaliana lipin homologs provide most of the diacylglycerol for TAG synthesis and that loss of their functions markedly reduces TAG content, but with only minor impact on eukaryotic galactolipid synthesis. Collectively, these results show that Arabidopsis lipins, along with PDAT1 and SDP1, function synergistically in directing FAs toward peroxisomal β-oxidation via TAG intermediates, thereby maintaining membrane lipid homeostasis in leaves.  相似文献   

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