首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A cellulose synthesis complex with a “rosette” shape is responsible for synthesis of cellulose chains and their assembly into microfibrils within the cell walls of land plants and their charophyte algal progenitors. The number of cellulose synthase proteins in this large multisubunit transmembrane protein complex and the number of cellulose chains in a microfibril have been debated for many years. This work reports a low resolution structure of the catalytic domain of CESA1 from Arabidopsis (Arabidopsis thaliana; AtCESA1CatD) determined by small-angle scattering techniques and provides the first experimental evidence for the self-assembly of CESA into a stable trimer in solution. The catalytic domain was overexpressed in Escherichia coli, and using a two-step procedure, it was possible to isolate monomeric and trimeric forms of AtCESA1CatD. The conformation of monomeric and trimeric AtCESA1CatD proteins were studied using small-angle neutron scattering and small-angle x-ray scattering. A series of AtCESA1CatD trimer computational models were compared with the small-angle x-ray scattering trimer profile to explore the possible arrangement of the monomers in the trimers. Several candidate trimers were identified with monomers oriented such that the newly synthesized cellulose chains project toward the cell membrane. In these models, the class-specific region is found at the periphery of the complex, and the plant-conserved region forms the base of the trimer. This study strongly supports the “hexamer of trimers” model for the rosette cellulose synthesis complex that synthesizes an 18-chain cellulose microfibril as its fundamental product.Cellulose, the most abundant biopolymer on Earth, is composed of linear chains of β-1,4 linked d-Glc monomers with repeating structural units of the disaccharide cellobiose. Numerous cellulose polymers cocrystallize to form microfibrils, which provide mechanical strength and rigidity to plants. Its natural abundance makes it an attractive target for many industrial applications, including paper and pulping, construction, and textile manufacture. More recently, cellulose has been used for production of biofuels, such as ethanol (Ragauskas et al., 2006; Langan et al., 2014), and in the form of nanocellulose as a component in advanced composite materials (Reddy et al., 2013; Habibi, 2014). Cellulose microfibrils are synthesized by a large membrane-bound protein complex. In the land plants and charophycean algae, the cellulose synthesis complex (CSC) has a “rosette” shape (Mueller et al., 1976; Mueller and Brown, 1980b; Kimura et al., 1999), and the entire CSC has reported diameters between 24 to 30 nm (Lerouxel et al., 2006). This structural information was revealed by freeze-fracture transmission electron microscopy, showing six lobes in a hexagonal arrangement at the point where the transmembrane helices of multiple cellulose synthase proteins (CESAs) cross the plasma membrane. Recently KORRIGAN, a protein with cellulase activity, has also been implicated as an integral component of the CSC (Vain et al., 2014).Vascular plants produce several different CESA isoforms. For example, Arabidopsis (Arabidopsis thaliana) has 10 different isoforms with 64% to 98% sequence identity (Holland et al., 2000; Richmond, 2000; McFarlane et al., 2014). The different CESA isoforms play specific roles in cellulose synthesis during plant development. In Arabidopsis, CESA1, CESA3, and CESA6 are required for primary cell wall synthesis, while CESA4, CESA7, and CESA8 are required for secondary cell wall synthesis (Gardiner et al., 2003; Taylor et al., 2003; Persson et al., 2007). CESA2, CESA5, and CESA9 play roles in tissue-specific processes and are partially redundant with CESA6, whereas CESA10 is closely related to AtCESA1 but evidently has a minor role in plant development (Somerville, 2006). The absolute number of CESA proteins present in a CSC remains a subject of much speculation, largely because the stoichiometry of the cellulose microfibril remains unresolved (Cosgrove, 2014). The traditional representation of the microfibril has 36 cellulose chains, and based on this, one would expect that each lobe of the rosette CSC contains six CESA proteins responsible for the synthesis of six glucan chains for a total of 36 CESA proteins per rosette CSC (Herth, 1983; Perrin, 2001; Doblin et al., 2002). However, recent studies using different analytical techniques combined with computation report 18 to 24 cellulose chains per microfibril (Fernandes et al., 2011; Thomas et al., 2013; Oehme et al., 2015). A study of cellulose from mung bean (Vigna radiata) primary cell walls, using x-ray diffraction, solid-state NMR, and computational analysis, supports an 18-chain model for a cellulose microfibril (Newman et al., 2013). This implies that the CSC is composed of fewer than 36 CESA proteins or that not all of the proteins in a CSC are simultaneously active. Further, it has been recently reported that the stoichiometry of CESAs 1, 3, and 6 and CESAs 4, 7, and 8 in the primary and secondary cell walls, respectively, is 1:1:1 (Gonneau et al., 2014; Hill et al., 2014). Together, these reports suggest a rosette CSC composed of 18 CESA proteins with three CESAs per lobe as the most likely composition of a rosette CSC to account for an 18-chain cellulose microfibril (Newman et al., 2013; Gonneau et al., 2014; Hill et al., 2014). In addition, it should also be noted that 24 CESA proteins in a rosette CSC with four proteins per lobe is incompatible with a 1:1:1 CESA stoichiometry.Numerous efforts to isolate active CESA proteins directly from plants or by recombinant expression have not been successful, preventing a detailed structural analysis of CESA proteins or the mechanism of plant cellulose synthesis. In contrast, the recently reported crystal structure of cellulose synthase from Rhodoobacter sphaeroides (Morgan et al., 2013) clearly showed that only a single cellulose synthase polypeptide is required for glucan polymerization and also identified the conserved sequence motifs responsible for catalysis. In addition, based on the presence of an 18-residue glucan chain in the protein tunnel, a mechanism for cellulose synthesis and translocation across a cytoplasmic membrane was proposed that also addressed how the alternate d-Glc molecules are inverted during polymer synthesis (Morgan et al., 2013; Omadjela et al., 2013). However, this structure cannot provide insight into the formation of microfibrils from the cellulose chains synthesized by single polypeptides of CESA.The CESA proteins of land plants and their charophycean algal relatives are multidomain single polypeptide chains of approximately 1000 amino acids. They are predicted to have eight transmembrane helices and to have their N- and C-terminal regions facing the cytoplasm (Pear et al., 1996). Although they share sequence similarity with the bacterial counterpart, they also have unique structural features not found in the bacterial enzymes. The N-terminal domain contains a Zn-binding site that may play a role in oligomerization of CESA proteins (Kurek et al., 2002). The putative cytosolic domain, which is flanked by a two-helix N-terminal transmembrane domain and a six-helix C-terminal transmembrane domain (McFarlane et al., 2014; Slabaugh et al., 2014), has D, D, D, QxxRW motifs that are conserved substrate binding and catalytic residues in the glycosyltransferase-2 superfamily (Nagahashi et al., 1995; Pear et al., 1996; Saxena and Brown, 1997; Yoshida et al., 2000). This domain also has a plant-conserved region (P-CR) and a class-specific region (CSR) that are only found in CESAs that form rosette CSCs. Although the roles of these regions are unknown, they are proposed to be involved in regulatory functions, such as interactions with other proteins and oligomerization to form the rosette shape. In the Arabidopsis CESAs, the sequence identity within the P-CR regions is greater than 80%, while in CSR regions, it is only about 40%. A recent computational model of the cytosolic domain of cotton (Gossypium hirsutum) CESA1 provides the first detailed structural model of the catalytic domain of CESA (Sethaphong et al., 2013). This model structure aligns well with the crystal structure of the bacterial cellulose synthase, indicating that a common mechanism exists for cellulose synthesis in bacteria and plants and that CESAs within rosette CSCs contain a single active synthetic site. In addition, this model made it possible to test possible configurations for the assembly of CESA monomers into a functional rosette CSC (Newman et al., 2013; Sethaphong et al., 2013).Our understanding of the mechanism of cellulose biosynthesis in plants at the molecular level is hampered by the lack of an atomic level CESA model. To gain deeper insight into the structure and role of the catalytic domain of CESA in rosette formation, we carried out a structural characterization of the cytosolic domain of Arabidopsis CESA1, a protein that is essential for cellulose synthesis in the primary cell wall (Arioli et al., 1998). The recombinant protein was purified from Escherichia coli in a two-step process that allowed us to obtain low-resolution structural information about the monomeric and trimeric forms of the recombinant protein using small-angle scattering (SAS) techniques. This study provides the first experimental evidence to support the self-assembly of CESAs into a stable trimer complex, revealing the possible role of the catalytic domain in the formation of the rosette CSC. Comparison of the size of the catalytic domain trimer with dimensions of rosette CSCs obtained from TEM studies strongly supports the “hexamer of trimers” model for rosette CSCs. Computational analysis of the scattering data suggested configurations for how the monomers, including the plant-specific P-CR and CSR domains, may be arranged in the trimeric lobes of the rosette CSC. Knowledge of how CESA proteins assemble in the CSC will enable approaches for rational genetic manipulation of plant cell wall synthesis, which offers enormous opportunities to improve feedstocks for the production of sustainable fuels and chemicals.  相似文献   

2.
Plant growth and organ formation depend on the oriented deposition of load-bearing cellulose microfibrils in the cell wall. Cellulose is synthesized by a large relative molecular weight cellulose synthase complex (CSC), which comprises at least three distinct cellulose synthases. Cellulose synthesis in plants or bacteria also requires the activity of an endo-1,4-β-d-glucanase, the exact function of which in the synthesis process is not known. Here, we show, to our knowledge for the first time, that a leaky mutation in the Arabidopsis (Arabidopsis thaliana) membrane-bound endo-1,4-β-d-glucanase KORRIGAN1 (KOR1) not only caused reduced CSC movement in the plasma membrane but also a reduced cellulose synthesis inhibitor-induced accumulation of CSCs in intracellular compartments. This suggests a role for KOR1 both in the synthesis of cellulose microfibrils and in the intracellular trafficking of CSCs. Next, we used a multidisciplinary approach, including live cell imaging, gel filtration chromatography analysis, split ubiquitin assays in yeast (Saccharomyces cerevisiae NMY51), and bimolecular fluorescence complementation, to show that, in contrast to previous observations, KOR1 is an integral part of the primary cell wall CSC in the plasma membrane.Cellulose microfibrils are synthesized by a hexameric multiprotein complex at the plasma membrane called the cellulose synthase complex (CSC). Genetic analysis, expression data, and coimmunoprecipitation experiments have demonstrated that a functional CSC contains at least three different nonredundant cellulose synthase (CESA) isoforms (Höfte et al., 2007). CESA1, CESA3, and CESA6-like are involved in cellulose biosynthesis during primary cell wall deposition, whereas CESA4, CESA7, and CESA8 are essential for cellulose synthesis in the secondary cell wall (Taylor et al., 1999, 2000, 2003; Desprez et al., 2007; Persson et al., 2007). CSCs labeled by fluorescently tagged CESA proteins migrate in the plasma membrane along cortical microtubules (CMTs), propelled by the polymerization of the β-1,4-glucans (Paredez et al., 2006). Partial depolymerization of CMTs using oryzalin showed that the organized trajectories of CSCs depend on the presence of an intact CMT array. The CSC-microtubule interaction is mediated at least in part by a large protein, POMPOM2/CELLULOSE SYNTHASE INTERACTING1, that binds to both CESAs and microtubules (Lei et al., 2014). Interestingly, complete depolymerization of CMTs does not alter the velocity of the complexes, illustrating that CMTs are necessary for the guidance of CSCs but not for their movement (Paredez et al., 2006). The microtubule cytoskeleton also has a role in the secretion and internalization of CSCs (Crowell et al., 2009; Gutierrez et al., 2009)KORRIGAN1 (KOR1) is a membrane-bound endo-1,4-β-d-glucanase (EGase) that is also required for cellulose synthesis (Nicol et al., 1998). Enzymatic analysis of a recombinant and soluble form of the Brassica napus KOR1 homolog showed substrate specificity for low-substituted carboxymethyl cellulose and amorphous cellulose but no activity on crystalline cellulose, xyloglucans, or short cellulose oligomers (Mølhøj et al., 2001; Master et al., 2004). Fractionation of microsomes demonstrated that KOR1 is primarily present in plasma membrane fractions but also at low levels in a tonoplast-enriched fraction (Nicol et al., 1998). Similarly, the KOR1 ortholog from tomato (Solanum lycopersicum) was found in the plasma membrane and fractions enriched for the Golgi apparatus (Brummell et al., 1997). A GFP-KOR1 fusion protein expressed with the Cauliflower mosaic virus 35S promoter accumulated in the Golgi apparatus and post-Golgi compartments and the tonoplast (Robert et al., 2005). Surprisingly for an enzyme involved in cellulose synthesis, the protein could not be detected at the plasma membrane. Using this construct, it was also shown that KOR1 undergoes regulated intracellular cycling (Robert et al., 2005).Although numerous genetic studies indicate that KOR1 is required for cellulose synthesis in primary and secondary cell walls and during cell plate formation (Nicol et al., 1998; Peng et al., 2000; Zuo et al., 2000; Lane et al., 2001; Sato et al., 2001; Szyjanowicz et al., 2004), its precise role in the cellulose synthesis process remains unclear. It has been suggested that KOR1 might be a component of the CSC (Read and Bacic, 2002). However, until now there has been no experimental evidence for this in Arabidopsis (Arabidopsis thaliana), either with coprecipitation experiments or with localization studies (Szyjanowicz et al., 2004; Robert et al., 2005; Desprez et al., 2007). Numerous hypotheses have been proposed to explain the paradoxical role of KOR1 in cellulose synthesis (Robert et al., 2004). KOR1 might have a proofreading activity involved in hydrolyzing disordered amorphous cellulose to relieve stress generated during the assembly of glucan chains in cellulose microfibrils (Mølhøj et al., 2002). Alternatively, KOR1 may determine the length of individual cellulose chains, either during cellulose synthesis or once the microfibril has been incorporated in the wall. A third hypothesis is that KOR1 releases the cellulose microfibril from the CSC before the complex is internalized from the plasma membrane (Somerville, 2006). Studies in cotton (Gossypium hirsutum) fiber extracts identified sitosterol glucoside as a primer for the cellulose synthesis and suggested that KOR1 could be involved in their cleavage from the nascent glucan chain (Peng et al., 2002). However, this scenario is unlikely, since, at least for the bacterial CESA, which is homologous to plant CESAs, there is no evidence for the existence of lipid-linked precursors, as shown by the three-dimensional structure of an active complex (Morgan et al., 2013).In this study, we first confirmed previous observations (Paredez et al., 2008) that, in the leaky kor1-1 mutant, the velocity of the CSCs is reduced compared with that in a wild-type background but that, in addition, the mutation affects the ability of the cellulose synthesis inhibitor CGA325′615 (hereafter referred to as CGA) to induce the accumulation of GFP-CESA3 in a microtubule-associated compartment (MASC/small compartments carrying cellulose synthase complexes [SmaCCs]; Crowell et al., 2009; Gutierrez et al., 2009). This indicates that KOR1 plays a role both in the synthesis of cellulose and in the intracellular trafficking of the CSC. Using gel filtration approaches, we identified KOR1 in fractions of high molecular mass, suggesting that KOR1 is present in membranes as part of a protein complex. We next analyzed the dynamics of GFP-KOR1 expressed in the kor1-1 mutant background under the control of its endogenous promoter. GFP-KOR1 is found in discrete particles at the plasma membrane in the same cells as GFP-CESAs (Crowell et al., 2009). GFP-KOR1 plasma membrane particles migrate along linear trajectories with comparable velocities to those observed for GFP-CESAs. The organization of GFP-KOR1 at the plasma membrane also requires the presence of an intact microtubule array, suggesting that KOR1 and CESA trajectories in the plasma membrane are regulated in the same manner. GFP-KOR1 and mCherry-CESA1 partially colocalize in the plasma membrane, Golgi, and post-Golgi compartments. Finally, we provide evidence for direct interaction between KOR1 and primary cell wall CESA proteins using the membrane-based yeast (Saccharomyces cerevisiae NMY51) two-hybrid (MbYTH) system (Timmers et al., 2009) and bimolecular fluorescence complementation (BiFC). Our data support a new model in which KOR1 is an integral part of the CSC, where it plays a role not only in the synthesis of cellulose but also in the intracellular trafficking of the CSC.  相似文献   

3.
4.
Clathrin-mediated endocytosis (CME) is the best-characterized type of endocytosis in eukaryotic cells. Plants appear to possess all of the molecular components necessary to carry out CME; however, functional characterization of the components is still in its infancy. A yeast two-hybrid screen identified μ2 as a putative interaction partner of CELLULOSE SYNTHASE6 (CESA6). Arabidopsis (Arabidopsis thaliana) μ2 is homologous to the medium subunit 2 of the mammalian ADAPTOR PROTEIN COMPLEX2 (AP2). In mammals, the AP2 complex acts as the central hub of CME by docking to the plasma membrane while concomitantly recruiting cargo proteins, clathrin triskelia, and accessory proteins to the sites of endocytosis. We confirmed that μ2 interacts with multiple CESA proteins through the μ-homology domain of μ2, which is involved in specific interactions with endocytic cargo proteins in mammals. Consistent with its role in mediating the endocytosis of cargos at the plasma membrane, μ2-YELLOW FLUORESCENT PROTEIN localized to transient foci at the plasma membrane, and loss of μ2 resulted in defects in bulk endocytosis. Furthermore, loss of μ2 led to increased accumulation of YELLOW FLUORESCENT PROTEIN-CESA6 particles at the plasma membrane. Our results suggest that CESA represents a new class of CME cargo proteins and that plant cells might regulate cellulose synthesis by controlling the abundance of active CESA complexes at the plasma membrane through CME.Cellulose microfibrils, as the major load-bearing polymers in plant cell walls, are the predominant component that enforces asymmetric cell expansion (Green, 1962). In higher plants, cellulose is synthesized by multimeric rosettes, which are also referred to as cellulose synthase complexes (CSCs; Kimura et al., 1999). Genetic and coimmunoprecipitation studies have indicated that CELLULOSE SYNTHASE1 (CESA1), CESA3, and CESA6-like (CESA6, CESA2, CESA5, and CESA9) isoforms are constituents of CSCs during primary cell wall synthesis (Persson et al., 2005; Desprez et al., 2007; Persson et al., 2007; Wang et al., 2008), whereas CESA4, CESA7, and CESA8 are implicated in the cellulose synthesis of secondary cell walls (Taylor et al., 1999, 2003; Brown et al., 2005). Knowledge about cellulose synthesis has recently been enhanced by the development of a system whereby the dynamics of CESA can be imaged in living cells (Paredez et al., 2006; Desprez et al., 2007). In agreement with earlier transmission electron microscopy studies in which rosettes were visualized in Golgi cisternae, vesicles, and at the plasma membrane (Haigler and Brown, 1986), fluorescent protein tagging of CESA has identified CESA localization at the plasma membrane, in Golgi bodies, and in small intracellular compartments (Paredez et al., 2006; Desprez et al., 2007; Crowell et al., 2009; Gutierrez et al., 2009; Gu et al., 2010; Lei et al., 2012; Li et al., 2012b).Assuming that cellulose synthesis occurs solely at the plasma membrane, the trafficking of CSCs to and from the plasma membrane may act as a significant regulatory mechanism. Although the mechanistic details of CESA trafficking are lacking, live cell imaging has shown that CESA localizes to various subcellular compartments. A subset of CESAs colocalize with markers of the trans-Golgi network (TGN)/early endosome (EE), an organelle that is part of both the secretory and endocytic pathways in Arabidopsis (Arabidopsis thaliana; Dettmer et al., 2006; Lam et al., 2007; Crowell et al., 2009, 2010; Viotti et al., 2010). CESAs also localize to microtubule-associated cellulose synthase compartments (MASCs) and small CESA-containing compartments (SmaCCs). The exact function of SmaCCs/MASCs is unknown, but it has been proposed that SmaCCs/MASCs might result from the internalization of CSCs or might act in the delivery of CSCs to the plasma membrane (Crowell et al., 2009, 2010; Gutierrez et al., 2009).Clathrin-mediated endocytosis (CME) has been shown to be a major endocytic pathway in Arabidopsis (Holstein, 2002; Samaj et al., 2005; Dhonukshe et al., 2007; Kleine-Vehn and Friml, 2008; Chen et al., 2011; Beck et al., 2012; Wang et al., 2013), although there is also evidence of clathrin-independent endocytosis mechanisms (Bandmann and Homann, 2012). The function of many CME proteins has been extensively characterized in mammals (McMahon and Boucrot, 2011), and homologs of many CME components are encoded by the Arabidopsis genome, including multiple copies of clathrin H chain and clathrin light chain (CLC), all four subunits of the heterotetrameric ADAPTOR PROTEIN COMPLEX2 (AP2) complex, dynamin-related proteins, and accessory proteins such as AP180 (Holstein, 2002; Chen et al., 2011); however, many CME components have yet to be characterized in plants.It has been suggested that CME might also function in controlling cell wall metabolism. For example, dividing and growing cells internalize cross-linked cell wall pectins, which might allow for cell wall remodeling (Baluska et al., 2002, 2005; Samaj et al., 2004). Moreover, the importance of endocytosis for cell wall morphogenesis is apparent from the functional characterization of proteins involved in CME. A dynamin-related protein, DRP1A, plays a significant role in endocytosis and colocalizes with CLC (Collings et al., 2008; Konopka and Bednarek, 2008). Defective endocytosis in RADIAL SWELLING9 (rsw9) plants, which contain a mutation in DRP1A, results in cellulose deficiency and defects in cell elongation (Collings et al., 2008). A mutation in rice, brittle culm3 (bc3), was mapped to the dynamin-related gene OsDRP2A, which has been proposed to function in CME. The brittle-culm phenotype in this mutant was attributed to cellulose deficiency (Xiong et al., 2010). Although the abundance of OsCESA4 was also altered in bc3, it remains unclear whether the cellulose deficiency of either bc3 or rsw9 results directly from perturbations in CESA trafficking.To identify proteins involved in the regulation of cellulose biosynthesis, a yeast two-hybrid (Y2H) screen was performed in which the central domain of CESA6 (CESA6CD) was used as bait to screen an Arabidopsis complementary DNA library for potential interaction partners of CESA6 (Gu et al., 2010; Gu and Somerville, 2010). The Y2H screen identified μ2 as a putative interaction partner of CESA6CD. The mammalian homolog of μ2 is the medium subunit of the tetrameric AP2, which acts as the core of the CME machinery by docking to the plasma membrane while concomitantly recruiting cargo proteins, clathrin triskelia, and accessory proteins to the sites of endocytosis (Jackson et al., 2010; McMahon and Boucrot, 2011; Cocucci et al., 2012). In this study, we provide evidence that μ2 plays a role in CME in Arabidopsis, that CESAs are a new set of CME cargo proteins, and that plant cells might regulate cellulose synthesis by controlling the abundance of CSCs at the plasma membrane through CME. To our knowledge, this study is the first to show the affect of an AP2 complex component on endocytosis in plants and the first to visualize an AP2 complex component in living plant cells. Furthermore, our data suggest that the role of AP2 in plants may differ from what has been shown in animals.  相似文献   

5.
Cellulose biosynthesis is a common feature of land plants. Therefore, cellulose biosynthesis inhibitors (CBIs) have a potentially broad-acting herbicidal mode of action and are also useful tools in decoding fundamental aspects of cellulose biosynthesis. Here, we characterize the herbicide indaziflam as a CBI and provide insight into its inhibitory mechanism. Indaziflam-treated seedlings exhibited the CBI-like symptomologies of radial swelling and ectopic lignification. Furthermore, indaziflam inhibited the production of cellulose within <1 h of treatment and in a dose-dependent manner. Unlike the CBI isoxaben, indaziflam had strong CBI activity in both a monocotylonous plant (Poa annua) and a dicotyledonous plant (Arabidopsis [Arabidopsis thaliana]). Arabidopsis mutants resistant to known CBIs isoxaben or quinoxyphen were not cross resistant to indaziflam, suggesting a different molecular target for indaziflam. To explore this further, we monitored the distribution and mobility of fluorescently labeled CELLULOSE SYNTHASE A (CESA) proteins in living cells of Arabidopsis during indaziflam exposure. Indaziflam caused a reduction in the velocity of YELLOW FLUORESCENT PROTEIN:CESA6 particles at the plasma membrane focal plane compared with controls. Microtubule morphology and motility were not altered after indaziflam treatment. In the hypocotyl expansion zone, indaziflam caused an atypical increase in the density of plasma membrane-localized CESA particles. Interestingly, this was accompanied by a cellulose synthase interacting1-independent reduction in the normal coincidence rate between microtubules and CESA particles. As a CBI, for which there is little evidence of evolved weed resistance, indaziflam represents an important addition to the action mechanisms available for weed management.Cellulose is a composite polymer of β-1,4-linked glucan chains and is the main load-bearing structure of plant cell walls (Jarvis, 2013). Although cellulose is a relatively simple polysaccharide molecule, its synthesis is quite complex. The principle catalytic unit is a plasma membrane (PM)-localized protein complex referred to as the cellulose synthase complex (CSC; Davis, 2012). In plants, the CSC, visualized with freeze fracture microscopy, is a solitary, hexagonal rosette-shaped complex (Herth and Weber, 1984; Delmer, 1999) and at least three of the catalytic CELLULOSE SYNTHASE A (CESA) proteins are required in each CSC for the production of cellulose (Desprez et al., 2007; Persson et al., 2007). In addition to CESAs, several accessory proteins have been discovered to be necessary for the production and deposition of cellulose, such as KORRIGAN (Lane et al., 2001), COBRA (Roudier et al., 2005) and CELLULOSE SYNTHASE INTERACTING1 (CSI1; Gu et al., 2010), as well as several others that are yet to be identified. The loss of function in any of the aforementioned proteins causes complete or partial loss of anisotropic growth in cells undergoing expansion, resulting in radial swelling. Severe radial swelling in rapidly expanding tissue is also a common symptomology observed in seedlings treated with cellulose biosynthesis inhibitors (CBIs). Therefore, numerous potential herbicidal targets exist (mechanisms of action) for the broad group of known CBIs.Classification of an herbicide to the CBI designation was traditionally achieved by short-term [14C]radioisotope tracer studies focused on the incorporation of Glc into cellulose (Heim et al., 1990; Sabba and Vaughn, 1999). More recently, time-lapse confocal microscopy of reporter-tagged CESA proteins (Paredez et al., 2006) has been used to further classify CBIs. CBIs can be classified into at least three primary groups based on how treatment disrupts the normal tracking and localization of fluorescently labeled CESAs (for review, see Brabham and DeBolt, 2012). The disruption is, it can be assumed, the result of the inhibitory mechanism of the CBI. In the first group, isoxaben and numerous other compounds cause YELLOW FLUORESCENT PROTEIN YFP):CESAs to be depleted from the PM and concomitantly accumulate in cytosolic vesicles (called small CESA compartments or microtubule-associated cellulose synthase compartments; Paredez et al., 2006; Crowell et al., 2009; Gutierrez et al., 2009) The second group, consisting only of dichlobenil (DCB), causes YFP:CESAs to become immobilized and hyperaccumulated at distinct foci in the PM (Herth, 1987; DeBolt et al., 2007b). The third group influences CSC-microtubule (MT)-associated functions resulting in errant movement and localization of YFP:CESAs (DeBolt et al., 2007a; Yoneda et al., 2007). These different disruption processes suggest that each CBI group targets a different aspect of the complex cellulose biosynthetic process.A lack of evolved weed resistance in the field suggests that CBIs are potentially underutilized tools for weed control (Sabba and Vaughn, 1999; Heap, 2014). CBIs have also been useful research tools in decoding fundamental aspects of cellulose biosynthesis. An exogenous application of a CBI provides spatial and temporal inhibition of cellulose. Resistance screens to CBIs have uncovered key genes in cellulose biosynthesis (Scheible et al., 2001; Desprez et al., 2002). Furthermore, CBIs such as isoxaben have also been effective in linking accessory proteins with CESAs in the CSC (Robert et al., 2005; Gu et al., 2010). Therefore, it is important to extend our range of CBI compounds. Indaziflam (Fig. 1A), an herbicide introduced by Bayer Crop Science, was recently proposed to be a CBI and was reported to have a photosystem II inhibition value of 9.4 (Meyer et al., 2009; Dietrich and Laber, 2012). Indaziflam is labeled for use in turf, for perennial crops, and for nonagricultural situations for preemergent control of grasses and broadleaf weeds (Meyer et al., 2009; Brosnan et al., 2011). The aim herein was to investigate indaziflam as a CBI and to characterize its inhibitory effect on cellulose biosynthesis.Open in a separate windowFigure 1.Indaziflam is a fluoroalkytriazine-containing compound that inhibits elongation in seedlings of P. annua and Arabidopsis. A, Chemical structure of indaziflam. B to D, Images of 7-d-old seedlings treated with increasing concentrations of indaziflam. B shows light-grown P. annua seedlings (indaziflam concentrations from left to right are 0, 100, 250, 500, 1,000, 5,000, and 10,000 pm). C and D show light-grown and dark-grown Arabidopsis seedlings, respectively (indaziflam concentrations from left to right are 0, 100, 250, 500, 1,000, and 2,500 pm). Indaziflam treatment induced swollen cells. E, Representative images of the primary root of P. annua grown in plates for 4 d with and without 10 nm indaziflam. F, Transgenic Arabidopsis seedlings expressing GFP:PIP2 were examined by laser scanning confocal microscopy and images represent visualization of the primary root grown vertically for 7-d plates without and with 250 pm indaziflam. PIP2, Plasma membrane intrinsic protein2. Bar = 10 mm in B, 5 mm in C and D, 2 mm in E, and 50 μm in F.  相似文献   

6.
In higher plants, cellulose is synthesized by plasma membrane–localized cellulose synthase complexes (CSCs). Arabidopsis thaliana GH9A1/KORRIGAN1 is a membrane-bound, family 9 glycosyl hydrolase that is important for cellulose synthesis in both primary and secondary cell walls. Most previously identified korrigan1 mutants show severe phenotypes such as embryo lethality; therefore, the role of GH9A1 in cellulose synthesis remains unclear. Here, we report a novel A577V missense mutation, designated jiaoyao1 (jia1), in the second of the glycosyl hydrolase family 9 active site signature motifs in GH9A1. jia1 is defective in cell expansion in dark-grown hypocotyls, roots, and adult plants. Consistent with its defect in cell expansion, this mutation in GH9A1 resulted in reduced cellulose content and reduced CSC velocity at the plasma membrane. Green fluorescent protein–GH9A1 is associated with CSCs at multiple locations, including the plasma membrane, Golgi, trans-Golgi network, and small CESA-containing compartments or microtubule-associated cellulose synthase compartments, indicating a tight association between GH9A1 and CSCs. GH9A1A577V abolishes the endoglucanase activity of GH9A1 in vitro but does not affect its interaction with CESAs in vitro, suggesting that endoglucanase activity is important for cellulose synthesis. Interestingly, jia1 results in both cellulose microfibril and microtubule disorganization. Our study establishes the important role of endoglucanase in cellulose synthesis and cellulose microfibril organization in plants.  相似文献   

7.
On the Inside     
Cellulose synthase complexes (CSCs) at the plasma membrane (PM) are aligned with cortical microtubules (MTs) and direct the biosynthesis of cellulose. The mechanism of the interaction between CSCs and MTs, and the cellular determinants that control the delivery of CSCs at the PM, are not yet well understood. We identified a unique small molecule, CESA TRAFFICKING INHIBITOR (CESTRIN), which reduces cellulose content and alters the anisotropic growth of Arabidopsis (Arabidopsis thaliana) hypocotyls. We monitored the distribution and mobility of fluorescently labeled cellulose synthases (CESAs) in live Arabidopsis cells under chemical exposure to characterize their subcellular effects. CESTRIN reduces the velocity of PM CSCs and causes their accumulation in the cell cortex. The CSC-associated proteins KORRIGAN1 (KOR1) and POM2/CELLULOSE SYNTHASE INTERACTIVE PROTEIN1 (CSI1) were differentially affected by CESTRIN treatment, indicating different forms of association with the PM CSCs. KOR1 accumulated in bodies similar to CESA; however, POM2/CSI1 dissociated into the cytoplasm. In addition, MT stability was altered without direct inhibition of MT polymerization, suggesting a feedback mechanism caused by cellulose interference. The selectivity of CESTRIN was assessed using a variety of subcellular markers for which no morphological effect was observed. The association of CESAs with vesicles decorated by the trans-Golgi network-localized protein SYNTAXIN OF PLANTS61 (SYP61) was increased under CESTRIN treatment, implicating SYP61 compartments in CESA trafficking. The properties of CESTRIN compared with known CESA inhibitors afford unique avenues to study and understand the mechanism under which PM-associated CSCs are maintained and interact with MTs and to dissect their trafficking routes in etiolated hypocotyls.Plant cell expansion and anisotropic cell growth are driven by vacuolar turgor pressure and cell wall extensibility, which in a dynamic and restrictive manner direct cell morphogenesis (Baskin, 2005). Cellulose is the major load-bearing component of the cell wall and is thus a major determinant for anisotropic growth (Baskin, 2001). Cellulose is made up of β-1,4-linked glucan chains that may aggregate to form microfibrils holding 18 to 36 chains (Somerville, 2006; Fernandes et al., 2011; Jarvis, 2013; Newman et al., 2013; Thomas et al., 2013). In contrast to cell wall structural polysaccharides, including pectin and hemicellulose, which are synthesized by Golgi-localized enzymes, cellulose is synthesized at the plasma membrane (PM) by cellulose synthase complexes (CSCs; Somerville, 2006; Scheller and Ulvskov, 2010; Atmodjo et al., 2013). The cellulose synthases (CESAs) are the principal catalytic units of cellulose biosynthesis and in higher plants are organized into globular rosettes (Haigler and Brown, 1986). For their biosynthetic function, each primary cell wall CSC requires a minimum of three catalytic CESA proteins (Desprez et al., 2007; Persson et al., 2007).On the basis of observations that cellulose microfibrils align with cortical microtubules (MTs) and that MT disruption leads to a loss of cell expansion, it was hypothesized that cortical MTs guide the deposition and, therefore, the orientation of cellulose (Green, 1962; Ledbetter and Porter, 1963; Baskin, 2001; Bichet et al., 2001; Sugimoto et al., 2003; Baskin et al., 2004; Wasteneys and Fujita, 2006). Confocal microscopy of CESA fluorescent fusions has advanced our understanding of CESA trafficking and dynamics. CSCs are visualized as small particles moving within the plane of the PM, with an average velocity of approximately 200 to 400 nm min−1. Their movement in linear tracks along cortical MTs (Paredez et al., 2006) supports the MT-cellulose alignment hypothesis.Our current understanding of cellulose synthesis suggests that CESAs are assembled into CSCs in either the endoplasmic reticulum (ER) or the Golgi apparatus and trafficked by vesicles to the PM (Bashline et al., 2014; McFarlane et al., 2014). The presence of CESAs in isolated Golgi and vesicles from the trans-Golgi network (TGN) has been established by proteomic studies (Dunkley et al., 2006; Drakakaki et al., 2012; Nikolovski et al., 2012; Parsons et al., 2012; Groen et al., 2014). Their localization at the TGN has been corroborated by electron microscopy and colocalization with TGN markers, such as vacuolar H+-ATP synthase subunit a1 (VHA-a1), and the Soluble NSF Attachment Protein Receptor (SNARE) protein SYNTAXIN OF PLANTS41 (SYP41), SYP42, and SYP61 (Crowell et al., 2009; Gutierrez et al., 2009; Drakakaki et al., 2012). A population of post-Golgi compartments carrying CSCs, referred to as microtubule-associated cellulose synthase compartments (MASCs) or small cellulose synthase compartments (SmaCCs), may be associated with MTs or actin filaments and are thought to be directly involved in either CSC delivery to, or internalization from, the PM (Crowell et al., 2009; Gutierrez et al., 2009).In addition to the CESAs, auxiliary proteins have been identified that play a vital role in the cellulose-synthesizing machinery. These include COBRA (Roudier et al., 2005), the endoglucanase KORRIGAN1 (KOR1; Lane et al., 2001; Lei et al., 2014b; Vain et al., 2014), and the recently identified POM-POM2/CELLULOSE SYNTHASE INTERACTIVE PROTEIN1 (POM2/CSI1; Gu et al., 2010; Bringmann et al., 2012). The latter protein functions as a linker between the cortical MTs and CSCs, as genetic lesions in POM2/CSI1 result in a lower incidence of coalignment between CSCs and cortical MTs (Bringmann et al., 2012). Given the highly regulated process of cellulose biosynthesis and deposition, it can be expected that many more accessory proteins participate in the delivery of CSCs and their interaction with MTs. Identification of these unique CSC-associated proteins can ultimately provide clues for the mechanisms behind cell growth and cell shape formation.Arabidopsis (Arabidopsis thaliana) mutants with defects in the cellulose biosynthetic machinery exhibit a loss of anisotropic growth, which results in organ swelling. This phenotype may be used as a diagnostic tool in genetic screens to identify cellulose biosynthetic and CSC auxiliary proteins (Mutwil et al., 2008). Chemical inhibitors complement genetic lesions to perturb, study, and control the cellular and physiological function of proteins (Drakakaki et al., 2009). A plethora of bioactive small molecules have been identified, and their analytical use contributes to our understanding of cellulose biosynthesis and CESA subcellular behavior (for review, see Brabham and Debolt, 2012). Small molecule treatment can induce distinct characteristic subcellular CESA patterns that can be broadly grouped into three categories (Brabham and Debolt, 2012). The first is characterized by the depletion of CESAs from the PM and their accumulation in cytosolic compartments, as observed for the herbicide isoxaben {N-[3-(1-ethyl-1-methylpropyl)-5-isoxazolyl]-2,6-dimethyoxybenzamide}, CGA 325615 [1-cyclohexyl-5-(2,3,4,5,6-pentafluorophe-noxyl)-1λ4,2,4,6-thiatriazin-3-amine], thaxtomin A (4-nitroindol-3-yl containing 2,5-dioxopiperazine), AE F150944 [N2-(1-ethyl-3-phenylpropyl)-6-(1-fluoro-1-methylethyl)-1,3,5-triazine-2,4-di-amine], and quinoxyphen [4-(2-bromo-4,5-dimethoxyphenyl)-3,4-dihydro-1H-benzo-quinolin-2-one]; (Paredez et al., 2006; Bischoff et al., 2009; Crowell et al., 2009; Gutierrez et al., 2009; Harris et al., 2012). The second displays hyperaccumulation of CESAs at the PM, as seen for the herbicides dichlobenil (2,6-dichlorobenzonitrile) and indaziflam {N-[(1R,2S)-2,3-dihydro-2,6-dimethyl-1H-inden-1-yl)-6-(1-fluoroethyl]-1,3,5-triazine-2,4-diamine} (Herth, 1987; DeBolt et al., 2007b; Brabham et al., 2014). The third exhibits disturbance of both CESAs and MTs and alters CESA trajectories at the PM, as exemplified by morlin (7-ethoxy-4-methylchromen-2-one; DeBolt et al., 2007a). Unique compounds inducing a phenotype combining CESA accumulation in intermediate compartments and disruption of CSC-MT interactions can contribute to both the identification of the accessory proteins linking CSCs with MTs and the vesicular delivery mechanisms of CESAs.In this study, we identified and characterized a unique cellulose deposition inhibitor, the small molecule CESA TRAFFICKING INHIBITOR (CESTRIN), which affects the localization pattern of CSCs and their interacting proteins in a unique way. The induction of cytoplasmic CESTRIN bodies might provide further clues for trafficking routes that carry CESAs to the PM.  相似文献   

8.
Plants are constantly subjected to various biotic and abiotic stresses and have evolved complex strategies to cope with these stresses. For example, plant cells endocytose plasma membrane material under stress and subsequently recycle it back when the stress conditions are relieved. Cellulose biosynthesis is a tightly regulated process that is performed by plasma membrane-localized cellulose synthase (CESA) complexes (CSCs). However, the regulatory mechanism of cellulose biosynthesis under abiotic stress has not been well explored. In this study, we show that small CESA compartments (SmaCCs) or microtubule-associated cellulose synthase compartments (MASCs) are critical for fast recovery of CSCs to the plasma membrane after stress is relieved in Arabidopsis thaliana. This SmaCC/MASC-mediated fast recovery of CSCs is dependent on CELLULOSE SYNTHASE INTERACTIVE1 (CSI1), a protein previously known to represent the link between CSCs and cortical microtubules. Independently, AP2M, a core component in clathrin-mediated endocytosis, plays a role in the formation of SmaCCs/MASCs. Together, our study establishes a model in which CSI1-dependent SmaCCs/MASCs are formed through a process that involves endocytosis, which represents an important mechanism for plants to quickly regulate cellulose synthesis under abiotic stress.  相似文献   

9.
CELLULOSE SYNTHASE5 (CESA5) synthesizes cellulose necessary for seed mucilage adherence to seed coat epidermal cells of Arabidopsis (Arabidopsis thaliana). The involvement of additional CESA proteins in this process and details concerning the manner in which cellulose is deposited in the mucilage pocket are unknown. Here, we show that both CESA3 and CESA10 are highly expressed in this cell type at the time of mucilage synthesis and localize to the plasma membrane adjacent to the mucilage pocket. The isoxaben resistant1-1 and isoxaben resistant1-2 mutants affecting CESA3 show defects consistent with altered mucilage cellulose biosynthesis. CESA3 can interact with CESA5 in vitro, and green fluorescent protein-tagged CESA5, CESA3, and CESA10 proteins move in a linear, unidirectional fashion around the cytoplasmic column of the cell, parallel with the surface of the seed, in a pattern similar to that of cortical microtubules. Consistent with this movement, cytological evidence suggests that the mucilage is coiled around the columella and unwinds during mucilage extrusion to form a linear ray. Mutations in CESA5 and CESA3 affect the speed of mucilage extrusion and mucilage adherence. These findings imply that cellulose fibrils are synthesized in an ordered helical array around the columella, providing a distinct structure to the mucilage that is important for both mucilage extrusion and adherence.The epidermal cells of Arabidopsis (Arabidopsis thaliana) seed coats produce two distinct secondary cell walls: pectin-rich mucilage and cellulose-rich columellae (Western et al., 2000). When seeds are hydrated, mucilage expands rapidly, rupturing the outer tangential cell wall and forming a mucilage capsule that surrounds the seed. Seed coat mucilage is composed primarily of rhamnogalacturonan I (RG I) and also contains homogalacturonan (HG), hemicelluloses (such as xylans and glucomannans), and cellulose (for review, see Haughn and Western, 2012). Extruded mucilage consists of an outer, nonadherent fraction and an inner, adherent fraction (Western et al., 2000, 2001; Macquet et al., 2007a). The adherent and nonadherent mucilage layers differ in the amount of methylesterified HG (Rautengarten et al., 2008; Saez-Aguayo et al., 2013; Voiniciuc et al., 2013), galactans (Dean et al., 2007; Macquet et al., 2007b), arabinans (Arsovski et al., 2009), mannans (Yu et al., 2014), and cellulose (Harpaz-Saad et al., 2011; Mendu et al., 2011; Sullivan et al., 2011), all of which influence the physical properties of the layers.Adherent mucilage has a distinct structure, which can be examined using cell wall dyes and antibodies. When treated with cellulose-specific dyes, densely stained rays extend from the top of each columella to the outer edge of the adherent layer, many cell lengths above the seed surface (Mendu et al., 2011; Sullivan et al., 2011). Cytological evidence indicates that cellulose, pectins, and mannans are components of the ray (Haughn and Western, 2012; Griffiths et al., 2014; North et al., 2014; Yu et al., 2014), although the exact manner in which they are assembled is unknown.Cellulose is abundant in mucilage rays and mediates adherence. Loss-of-function mutations in CELLULOSE SYNTHASE5 (CESA5) result in reduced cellulose levels and increased detachment of mucilage from the seed (Harpaz-Saad et al., 2011; Mendu et al., 2011; Sullivan et al., 2011; Griffiths et al., 2014). How a reduction in cellulose results in a loss of adherence is still unknown, but it likely involves interaction with other mucilage components such as pectin and arabinogalactan proteins (Griffiths et al., 2014). Since cesa5 mutants still have some cellulose in the rays of the adherent mucilage halo (Mendu et al., 2011; Sullivan et al., 2011), additional cellulose synthases must be involved in mucilage cellulose biosynthesis.The Arabidopsis genome encodes 10 different CESAs (Delmer, 1999; Richmond and Somerville, 2000). Multiple lines of evidence suggest that three different CESAs are required to form one active cellulose synthase complex (CSC; for review, see Somerville, 2006). CSCs are membrane-bound protein complexes that synthesize cellulose microfibrils in the apoplast (for review, see Somerville, 2006; Endler and Persson, 2011; Lei et al., 2012). CESA1, CESA3, and CESA6 are considered the core components of the primary wall CSC (Desprez et al., 2007; Persson et al., 2007). CESA2, CESA5, and CESA9 are partially redundant to CESA6 in primary wall biosynthesis, and genetic evidence suggests that each of these CESA polypeptides can form a functional CSC with CESA3 and CESA1 (Desprez et al., 2007; Persson et al., 2007). CESA10 is expressed in young plants, stems, floral tissue, and the base of rosette leaves (Beeckman et al., 2002; Doblin et al., 2002), but its function in cellulose biosynthesis is unclear. Other cesa mutant lines have been examined for altered mucilage phenotypes (cesa1, radially swollen1 [Burn et al., 2002; Sullivan et al., 2011], cesa2, cesa6, and cesa9 [Mendu et al., 2011]; CESA3, je5 [Sullivan et al., 2011] and cesa10-1 [Sullivan et al., 2011]); to date, only CESA5 has been shown to be required for cellulose biosynthesis during mucilage deposition.Two mutant alleles of CESA3, isoxaben resistant1-1 (ixr1-1) and ixr1-2, were isolated in a screen for resistance to the herbicide isoxaben (Scheible et al., 2001). Isoxaben inhibits the incorporation of Glc into the emerging cellulose polymer and is considered a potent and specific inhibitor of cellulose biosynthesis (Heim et al., 1990). Homozygous ixr1-1 and ixr1-2 lines show increased resistance to the herbicide, and the mutations causing this resistance were mapped to the genomic locus of CESA3 (Heim et al., 1990; Scheible et al., 2001). The ixr1-1 and ixr1-2 mutations cause amino acid substitutions near the C terminus of the CESA3 protein. ixr1-1 causes a Gly-to-Asn substitution (G998A) located in a transmembrane domain, while ixr1-2 contains a Thr-to-Ile substitution (T942I) in an apoplastic region of the protein between two transmembrane domains (Scheible et al., 2001). Recently, the ixr1-2 allele was shown to affect the velocity of CSCs in the plasma membrane, which consequently modifies cellulose crystallinity in the cell wall (Harris et al., 2012). It is not exactly clear how the ixr1-1 mutation affects cellulose biosynthesis. The effects of either of these mutations on seed coat mucilage have not been investigated.Since mucilage is composed primarily of pectins with smaller amounts of cellulose, seed coat epidermal cells represent an excellent system to study cellulose biosynthesis and interactions between cellulose and other wall components in muro. In this study, we investigated how cellulose is synthesized and deposited in seed coat epidermal cells. We show that at least three different CESA proteins are highly expressed in the seed coat epidermis during mucilage biosynthesis. These CESAs are oriented and move in a linear fashion around the cytoplasmic column of each cell in an identical pattern to cortical microtubules. In addition, we provide evidence that the adherent mucilage has a helical structure that expands and unwinds during extrusion to form the mucilage ray. We propose that during seed coat epidermal cell development, the biosynthesis of cellulose predetermines the structure of rays in the adherent mucilage layer.  相似文献   

10.
Anisotropic plant cell growth depends on the coordination between the orientation of cortical microtubules and the orientation of nascent cellulose microfibrils. CELLULOSE SYNTHASE INTERACTIVE1 (CSI1) is a key scaffold protein that guides primary cellulose synthase complexes (CSCs) along cortical microtubules during cellulose biosynthesis. Here, we investigated the function of the CSI1-like protein, CSI3, in Arabidopsis thaliana. Similar to CSI1, CSI3 associates with primary CSCs in vitro, colocalizes with CSCs in vivo, and exhibits the same plasma membrane localization and bidirectional motility as CSI1. However, ProCSI1:GFP-CSI3 cannot complement the anisotropic cell growth defect in csi1 mutants, suggesting that CSI3 is not functionally equivalent to CSI1. Also, the colocalization ratio between CSI1 and CSI3 is low, which may suggest heterogeneity within the CSC population. csi1 csi3 double mutants showed an enhanced cell expansion defect as well as an additive reduction of CSC velocities, and CSI3 dynamics are dependent on CSI1 function. We propose that CSI3 is an important regulator of plant cellulose biosynthesis and plant anisotropic cell growth that modulates the velocity of CSCs in both a microtubule-dependent and microtubule-independent manner.  相似文献   

11.
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   

12.
The actin and microtubule cytoskeletons regulate cell shape across phyla, from bacteria to metazoans. In organisms with cell walls, the wall acts as a primary constraint of shape, and generation of specific cell shape depends on cytoskeletal organization for wall deposition and/or cell expansion. In higher plants, cortical microtubules help to organize cell wall construction by positioning the delivery of cellulose synthase (CesA) complexes and guiding their trajectories to orient newly synthesized cellulose microfibrils. The actin cytoskeleton is required for normal distribution of CesAs to the plasma membrane, but more specific roles for actin in cell wall assembly and organization remain largely elusive. We show that the actin cytoskeleton functions to regulate the CesA delivery rate to, and lifetime of CesAs at, the plasma membrane, which affects cellulose production. Furthermore, quantitative image analyses revealed that actin organization affects CesA tracking behavior at the plasma membrane and that small CesA compartments were associated with the actin cytoskeleton. By contrast, localized insertion of CesAs adjacent to cortical microtubules was not affected by the actin organization. Hence, both actin and microtubule cytoskeletons play important roles in regulating CesA trafficking, cellulose deposition, and organization of cell wall biogenesis.Plant cells are surrounded by a flexible yet durable extracellular matrix that makes up the cell wall. This structure offers mechanical strength that counters osmotically driven turgor pressure, is an important factor for water movement in plants, acts as a physical barrier against pathogens (Somerville et al., 2004), and is a determining factor for plant cell morphogenesis. Hence, the cell wall plays a central role in plant biology.Two main types of cell walls can typically be distinguished: the primary and the secondary cell wall. The major load-bearing component in both of these cell walls is the β-1,4-linked glucan polymer cellulose (Somerville et al., 2004). Cellulose polymers are synthesized by plasma membrane (PM)-localized cellulose synthase (CesA) complexes (Mueller and Brown, 1980), which contain several CesA subunits with similar amino acid sequences (Mutwil et al., 2008a). The primary wall CesA complexes are believed to be assembled in the Golgi and are subsequently delivered to the PM via vesicular trafficking (Gutierrez et al., 2009), sometimes associated with Golgi pausing (Crowell et al., 2009). Furthermore, the primary wall CesA complexes are preferentially inserted into the PM at sites that coincide with cortical microtubules (MTs), which subsequently guide cellulose microfibril deposition (Gutierrez et al., 2009). Hence, the cortical MT array is a determinant for multiple aspects of primary wall cellulose production.The actin cytoskeleton plays a crucial role in organized deposition of cell wall polymers in many cell types, including cellulose-related polymers and pectins in tip-growing cells, such as pollen tubes and root hairs (Hu et al., 2003; Chen et al., 2007). Thus, actin-depolymerizing drugs and genetic manipulation of ACTIN genes impair directed expansion of tip-growing cells and long-distance transport of Golgi bodies with vesicles to growing regions (Ketelaar et al., 2003; Szymanski, 2005). In diffusely growing cells in roots and hypocotyls, loss of anisotropic growth has also been observed in response to mutations to vegetative ACTIN genes and to actin-depolymerizing and -stabilizing drugs (Baluska et al., 2001; Kandasamy et al., 2009). While actin is clearly important for cell wall assembly, it is less clear what precise roles it plays.One well-known function of actin in higher plants is to support intracellular movement of cytoplasmic organelles via actomyosin-based motility (Geisler et al., 2008; Szymanski, 2009). During primary wall synthesis in interphase cells, treatment with the actin assembly inhibitor latrunculin B (LatB) led to inhibition of Golgi motility and pronounced inhomogenities in CesA density at the PM (Crowell et al., 2009; Gutierrez et al., 2009) that coincided with the density of underlying and immobile Golgi bodies (Gutierrez et al., 2009). These results suggested that Golgi motility is important for CesA distribution (Gutierrez et al., 2009). The actin cytoskeleton also appears to be important for secondary wall cellulose microfibril deposition. For example, longitudinal actin filaments (AFs) define the movement of secondary wall CesA-containing Golgi bodies in developing xylem vessels (Wightman and Turner, 2008). In addition, it has been proposed that the AFs also can regulate the delivery of the secondary wall CesA complex to the PM via pausing of the Golgi (Wightman and Turner, 2008). It is therefore clear that actin organization is important for CesA distribution and for the pattern of cellulose microfibril deposition.Despite the above findings, very few reports have undertaken detailed studies to elucidate the role of the actin cytoskeleton in the distribution and trafficking of specific proteins in plant cells. Here, we have investigated the intracellular trafficking of CesA-containing vesicles and delivery of CesAs to the PM, in the context of the actin cytoskeleton. We quantitatively demonstrate that the organization of the actin cytoskeleton regulates CesA-containing Golgi distribution and the exocytic and endocytic rate of the CesAs. However, actin organization has no effect on the localized insertion of CesAs at sites of MTs at the PM.  相似文献   

13.
The current dogma for cell wall polysaccharide biosynthesis is that cellulose (and callose) is synthesized at the plasma membrane (PM), whereas matrix phase polysaccharides are assembled in the Golgi apparatus. We provide evidence that (1,3;1,4)-β-d-glucan (mixed-linkage glucan [MLG]) does not conform to this paradigm. We show in various grass (Poaceae) species that MLG-specific antibody labeling is present in the wall but absent over Golgi, suggesting it is assembled at the PM. Antibodies to the MLG synthases, cellulose synthase-like F6 (CSLF6) and CSLH1, located CSLF6 to the endoplasmic reticulum, Golgi, secretory vesicles, and the PM and CSLH1 to the same locations apart from the PM. This pattern was recreated upon expression of VENUS-tagged barley (Hordeum vulgare) CSLF6 and CSLH1 in Nicotiana benthamiana leaves and, consistent with our biochemical analyses of native grass tissues, shown to be catalytically active with CSLF6 and CSLH1 in PM-enriched and PM-depleted membrane fractions, respectively. These data support a PM location for the synthesis of MLG by CSLF6, the predominant enzymatically active isoform. A model is proposed to guide future experimental approaches to dissect the molecular mechanism(s) of MLG assembly.  相似文献   

14.
Interactions between cell wall polymers are critical for establishing cell wall integrity and cell-cell adhesion. Here, we exploit the Arabidopsis (Arabidopsis thaliana) seed coat mucilage system to examine cell wall polymer interactions. On hydration, seeds release an adherent mucilage layer strongly attached to the seed in addition to a nonadherent layer that can be removed by gentle agitation. Rhamnogalacturonan I (RG I) is the primary component of adherent mucilage, with homogalacturonan, cellulose, and xyloglucan constituting minor components. Adherent mucilage contains rays composed of cellulose and pectin that extend above the center of each epidermal cell. CELLULOSE SYNTHASE5 (CESA5) and the arabinogalactan protein SALT-OVERLY SENSITIVE5 (SOS5) are required for mucilage adherence through unknown mechanisms. SOS5 has been suggested to mediate adherence by influencing cellulose biosynthesis. We, therefore, investigated the relationship between SOS5 and CESA5. cesa5-1 seeds show reduced cellulose, RG I, and ray size in adherent mucilage. In contrast, sos5-2 seeds have wild-type levels of cellulose but completely lack adherent RG I and rays. Thus, relative to each other, cesa5-1 has a greater effect on cellulose, whereas sos5-2 mainly affects pectin. The double mutant cesa5-1 sos5-2 has a much more severe loss of mucilage adherence, suggesting that SOS5 and CESA5 function independently. Double-mutant analyses with mutations in MUCILAGE MODIFIED2 and FLYING SAUCER1 that reduce mucilage release through pectin modification suggest that only SOS5 influences pectin-mediated adherence. Together, these findings suggest that SOS5 mediates adherence through pectins and does so independently of but in concert with cellulose synthesized by CESA5.Cellulosic cell walls are a defining feature of land plants. Primary cell walls are composed of three major classes of polysaccharides: cellulose, hemicelluloses, and pectins. In addition, approximately 10% of the primary cell wall is composed of protein (Burton et al., 2010). Cell walls provide mechanical support for the cell, and cell wall carbohydrates in the middle lamellae mediate cell-cell adhesion (Caffall and Mohnen, 2009). Current models of cell wall structure depict a cellulose-hemicellulose network embedded in an independent pectin gel (for review, see Albersheim et al., 2011). These components are believed to interact through both covalent and noncovalent bonds to provide structure and strength to the cell wall, although the relative importance of pectin and its interactions with the hemicellulose-cellulose network remain unclear (for review, see Cosgrove, 2005).Another gap in our understanding of cell wall structure and assembly is the role of arabinogalactan proteins (AGPs). AGPs are a family of evolutionarily conserved secreted proteins highly glycosylated with type II arabinogalactans, and they can be localized to the plasma membrane by a C-terminal glycophosphatidylinositol (GPI) lipid anchor (for review, see Schultz et al., 2000; Showalter, 2001; Johnson et al., 2003; Seifert and Roberts, 2007; Ellis et al., 2010). AGPs can be extensively modified in the cell wall; many glycosyl hydrolases can affect AGP function by cleaving their glycosyl side chains (Sekimata et al., 1989; Cheung et al., 1995; Wu et al., 1995; Kotake et al., 2005). The GPI anchor can also be cleaved, releasing the AGPs from the membrane into the cell wall (Schultz et al., 2000). Although their exact roles are still unclear, AGPs have been proposed to interact with cell wall polysaccharides, initiate intracellular signaling cascades, and influence a wide variety of biological processes (for review, see Seifert and Roberts, 2007; Ellis et al., 2010; Tan et al., 2013).Many fasciclin-like AGPs (FLAs), which contain at least one fasciclin domain (FAS) associated with protein-protein interactions, have been suggested to influence cellulose biosynthesis or organization (Seifert and Roberts, 2007; Li et al., 2010; MacMillan et al., 2010). FLA3 RNA interference lines have reduced intine cell wall biosynthesis and loss of Calcofluor white (a fluorescent dye specific for glycan molecules) staining in aborted pollen grains (Li et al., 2010). A fla11 fla12 double mutant was shown to have reduced cellulose deposition, altered cellulose microfibril angle, and reduced cell wall integrity (MacMillan et al., 2010). The fla11 fla12 double mutant also had reductions in arabinans, galactans, and rhamnose (MacMillan et al., 2010). FLA4/SALT-OVERLY SENSITIVE5 (SOS5) was identified in a screen for salt sensitivity in roots. The SOS5 gene encodes an FLA protein with a GPI anchor, two AGP-like domains, and two FAS domains (Shi et al., 2003). Plants homozygous for the loss-of-function conditional allele sos5-1 have thinner root cell walls that appear less organized (Shi et al., 2003). The presence of the two FAS domains has led to the suggestion that SOS5 may interact with other proteins, forming a network that strengthens the cell wall (Shi et al., 2003). SOS5 is involved in regulation of cell wall rheology through a pathway involving two Leu-rich repeat receptor-like kinases, FEI1 and FEI2 (Xu et al., 2008). SOS5 and FEI2 are also required for normal seed coat mucilage adherence and hypothesized to do so by influencing cellulose biosynthesis (Harpaz-Saad et al., 2011, 2012).Arabidopsis (Arabidopsis thaliana) seed coat mucilage is a powerful model for studying cell wall biosynthesis and polysaccharide interactions (Arsovski et al., 2010; Haughn and Western, 2012). Seed coat epidermal cells sequentially produce two distinct types of secondary cell walls with unique morphologies and properties (Western et al., 2000; Windsor et al., 2000). Between approximately 5 and 9 d approximate time of fertilization (DPA), seed coat epidermal cells synthesize mucilage and deposit it in the apoplast, creating a donut-shaped mucilage pocket that surrounds a central cytoplasmic column (Western et al., 2000, 2004; Haughn and Chaudhury, 2005). From 9 to 13 DPA, the cytoplasmic column is gradually replaced by a cellulose-rich, volcano-shaped secondary cell wall called the columella (Beeckman et al., 2000; Western et al., 2000; Windsor et al., 2000; Stork et al., 2010; Mendu et al., 2011).Seed mucilage is composed primarily of relatively unbranched rhamnogalacturonan I (RG I) with minor amounts of homogalacturonan (HG), cellulose, and hemicelluloses (for review, see Haughn and Western, 2012). When mucilage is hydrated, it expands rapidly from the apoplastic pocket, forming a halo that surrounds the seed. Mucilage separates into two fractions: a loose nonadherent fraction and an inner adherent fraction that can only be released by vigorous shaking, strong bases, or glycosidases (for review, see North et al., 2014). Galactans and arabinans are also present in mucilage, and their regulation by glycosidases is required for correct mucilage hydration (Dean et al., 2007; Macquet et al., 2007b; Arsovski et al., 2009). For example, β-XYLOSIDASE1 encodes a bifunctional β-d-xylosidase/α-l-arabinofuranosidase required for arabinan modification in mucilage, and β-xylosidase1 mutant seeds have a delayed mucilage release phenotype (Arsovski et al., 2009). MUCILAGE MODIFIED2 (MUM2) encodes a β-d-galactosidase, and mum2 seeds fail to release mucilage when hydrated in water (Dean et al., 2007; Macquet et al., 2007b). MUM2 is believed to modify RG I galactan side chains but may also affect the galactan component of other mucilage components (Dean et al., 2007; Macquet et al., 2007b). Galactans are capable of binding to cellulose in vitro and could affect mucilage hydration through pectin-cellulose interactions (Zykwinska et al., 2005, 2007a, 2007b; Dick-Pérez et al., 2011; Wang et al., 2012), although carbohydrate linkage analysis suggests that the galactan side chains are very short.Several studies indicate that seed mucilage extrusion and expansion are also influenced by methylesterification of HG. For example, both SUBTILISIN-LIKE SER PROTEASE1.7 and PECTIN METHYLESTERASE INHIBITOR6 are required for proper methyl esterification of mucilage (Rautengarten et al., 2008; Saez-Aguayo et al., 2013). Mutations in another gene, FLYING SAUCER1 (FLY1; a transmembrane E3 ubiquitin ligase), reduce the degree of pectin methylesterification in mucilage and cause increased mucilage adherence and defective mucilage extrusion (Voiniciuc et al., 2013). fly1 seeds have disc-like structures at the edge of the mucilage halo, which are outer primary cell wall fragments that detach from the columella during extrusion and are difficult to separate from the adherent mucilage (Voiniciuc et al., 2013).Recently, CELLULOSE SYNTHASE5 (CESA5) and SOS5 were proposed to facilitate cellulose-mediated mucilage adherence (Harpaz-Saad et al., 2011; Mendu et al., 2011; Sullivan et al., 2011). A simple hypothesis for the role of CESA5 in mucilage adherence is that it synthesizes cellulose, which interacts with the mucilage pectin to mediate adherence. Loss of CESA5 function results in a reduction of mucilage cellulose biosynthesis and a less adherent mucilage cell wall matrix (Mendu et al., 2011; Sullivan et al., 2011). The role of SOS5 in mucilage adherence is more difficult to explain. SOS5 null mutations cause a loss-of-adherence phenotype similar to cesa5-1 seeds, suggesting that SOS5 may regulate mucilage adherence by influencing CESA5 function (Harpaz-Saad et al., 2011). However, the mechanism through which SOS5 could influence CESA5 and/or cellulose biosynthesis is not clear.To better understand the role of SOS5 in mucilage adherence and its relationship to CESA5, we thoroughly investigated the seed coat epidermal cell phenotypes of the cesa5-1 and sos5-2 single mutants as well as those of the cesa5-1 sos5-2 double mutant. We also investigated how cellulose, SOS5, and pectin interact to mediate mucilage adherence by constructing double mutants with either cesa5-1 or sos5-2 together with either mum2-1 or fly1. Our results suggest that SOS5 mediates mucilage adherence independently of CESA5. Furthermore, compared with CESA5, SOS5 has a greater influence on mucilage pectin structure, suggesting that SOS5 mediates mucilage adherence through pectins, not cellulose.  相似文献   

15.
Contractile cell walls are found in various plant organs and tissues such as tendrils, contractile roots, and tension wood. The tension-generating mechanism is not known but is thought to involve special cell wall architecture. We previously postulated that tension could result from the entrapment of certain matrix polymers within cellulose microfibrils. As reported here, this hypothesis was corroborated by sequential extraction and analysis of cell wall polymers that are retained by cellulose microfibrils in tension wood and normal wood of hybrid aspen (Populus tremula × Populus tremuloides). β-(1→4)-Galactan and type II arabinogalactan were the main large matrix polymers retained by cellulose microfibrils that were specifically found in tension wood. Xyloglucan was detected mostly in oligomeric form in the alkali-labile fraction and was enriched in tension wood. β-(1→4)-Galactan and rhamnogalacturonan I backbone epitopes were localized in the gelatinous cell wall layer. Type II arabinogalactans retained by cellulose microfibrils had a higher content of (methyl)glucuronic acid and galactose in tension wood than in normal wood. Thus, β-(1→4)-galactan and a specialized form of type II arabinogalactan are trapped by cellulose microfibrils specifically in tension wood and, thus, are the main candidate polymers for the generation of tensional stresses by the entrapment mechanism. We also found high β-galactosidase activity accompanying tension wood differentiation and propose a testable hypothesis that such activity might regulate galactan entrapment and, thus, mechanical properties of cell walls in tension wood.Contractile cell walls found in plant organs and tissues such as tendrils, contractile roots, and tension wood (TW) have remarkable functions and properties. Their traits have been most intensely studied in TW of hardwoods, where they provide negative gravitropic response capacities to stems with secondary growth, as recently reviewed by Mellerowicz and Gorshkova (2012). These properties are conferred by TW fibers, which in many species contain a so-called gelatinous cell wall layer (G-layer; Norberg and Meier, 1966; Clair et al., 2008). G-layers are formed following the deposition of xylan-type secondary cell wall layer(s) and, thus, can be considered tertiary layers (Wardrop and Dadswell, 1948). They are almost or completely devoid of xylan and lignin and have very high cellulose contents (up to 85%). However, several other polymers appear to be present in TW G-layers, according to recent chemical analyses of isolated G-layers (Nishikubo et al., 2007; Kaku et al., 2009) and immunohistochemical labeling of TW sections (Arend, 2008; Bowling and Vaughn, 2008). Notably, xyloglucan (XG) has been found in G-layers of poplar (Populus spp.) TW (Nishikubo et al., 2007) and at the boundary between secondary cell wall layers (S-layers) and G-layers (Baba et al., 2009; Sandquist et al., 2010). It is also important for tension creation (Baba et al., 2009). However, it is not detectable in mature G-layers by monoclonal antibodies or XG-binding modules (Nishikubo et al., 2007; Baba et al., 2009; Sandquist et al., 2010).Structurally similar G-layers have been also identified in phloem fibers in many fibrous crops, such as flax (Linum usitatissimum), hemp (Cannabis sativa), and ramie (Boehmeria nivea; Gorshkova et al., 2012). These fibers occur in bundles that can be isolated for biochemical analysis. G-layers in fibers from diverse sources have a very similar structure, being largely composed of cellulose (with axial microfibril orientation, high degrees of crystallinity, and large crystallite sizes) lacking xylan and lignin (Mellerowicz et al., 2001; Pilate et al., 2004; Gorshkova et al., 2010, 2012) and having high water contents (Schreiber et al., 2010). In phloem fibers, the G-layers become very prominent, reaching thicknesses up to 15 µm and occupying over 90% of the cell wall’s total cross-sectional areas (Crônier et al., 2005). Pectic β-(1→4)-galactan with complex structures has been shown to be the major matrix polysaccharide of isolated phloem fibers in flax (Gorshkova et al., 2004; Gorshkova and Morvan, 2006; Gurjanov et al., 2007). Some of it is so strongly retained within cellulose that it cannot be extracted by concentrated alkali and can only be obtained after cellulose dissolution (Gurjanov et al., 2008). Such galactan, therefore, is a prime candidate for a polymer entrapped by cellulose microfibrils during crystallization that could substantially contribute to the contractile properties of cellulose in G-layers, according to recently formulated models (Mellerowicz et al., 2008; Mellerowicz and Gorshkova 2012). Furthermore, Roach et al. (2011) have shown that trimming of β-(1→4)-galactan by β-galactosidase is important for final cellulose crystallization, the formation of G-layer structure, and, hence, the stem’s mechanical properties.There is also immunocytochemical evidence for the presence of β-(1→4)-galactan and type II arabinogalactan (AG-II) in G-layers of TW fibers (Arend, 2008; Bowling and Vaughn, 2008). In addition, high-Mr branched galactans have been isolated from TW of Fagus sylvestris (Meier, 1962) and Fagus grandifolia (Kuo and Timell, 1969), with estimated degrees of polymerization (DP) of approximately 300 and complex structure, probably including both β-(1→4) and β-(1→6) linkages, although their exact nature remains unknown. Furthermore, Gal has been identified as one of the major sugars after Glc and Xyl in hydrolysates of isolated Populus spp. G-layers (Furuya et al., 1970; Nishikubo et al., 2007), and the Gal content of cell walls is a proposed indicator of the extent of TW development in beech (Fagus spp.; Ruel and Barnoud, 1978). However, subsequent linkage analyses identified only 2- and 3,6-linked Gal in poplar TW G-layers (Nishikubo et al., 2007), while in flax fibers, 4-linked Gal is the main component (Gorshkova et al., 1996, 2004; Gurjanov et al., 2007, 2008). Thus, the type(s) of galactans present in poplar TW remains unclear, and the galactans have not been shown previously either to have a rhamnogalacturonan-I (RG-I) backbone or to be strongly retained by cellulose microfibrils, as demonstrated for flax gelatinous fibers.To improve our understanding of cell wall properties in TW and their contraction mechanism, in the study presented here, we tested aspects of the recently proposed entrapment model (Mellerowicz et al., 2008; Mellerowicz and Gorshkova, 2012). According to this model, contraction is driven by the formation of larger cellulose structures, sometimes called macrofibrils, via interactions of cellulose microfibrils in the G-layer with each other and forming inclusions containing matrix polymers. This would induce tension within cellulose through the stretching of microfibrils required to surround the inclusions. The model is compatible with available data on the structure and action of gelatinous walls, but the main assumption, that polymers are trapped inside crystalline cellulose, such as that found in flax, has not been tested previously. Therefore, we compared matrix polymers retained by cellulose microfibrils in normal wood (NW) and TW of the model hardwood species hybrid aspen (Populus tremula × Populus tremuloides) that forms TW with gelatinous fibers. For this purpose, we used a combination of sequential cell wall extractions, similar to those used previously to characterize flax gelatinous fibers (Gurjanov et al., 2008), followed by fractionation of polymers by size-exclusion chromatography, immunological analyses, and oligosaccharide profiling by polysaccharide analysis using carbohydrate gel electrophoresis (PACE). The results reveal the main polymers of cellulose-retained fractions and key differences between NW and TW. Comparison of our results and previous findings also indicates that there are both similarities and differences in the constitution of gelatinous fibers in aspen and flax. An updated model of the contractile G-layer of TW fibers based on the data is presented.  相似文献   

16.
The structural role of pectins in plant primary cell walls is not yet well understood because of the complex and disordered nature of the cell wall polymers. We recently introduced multidimensional solid-state nuclear magnetic resonance spectroscopy to characterize the spatial proximities of wall polysaccharides. The data showed extensive cross peaks between pectins and cellulose in the primary wall of Arabidopsis (Arabidopsis thaliana), indicating subnanometer contacts between the two polysaccharides. This result was unexpected because stable pectin-cellulose interactions are not predicted by in vitro binding assays and prevailing cell wall models. To investigate whether the spatial contacts that give rise to the cross peaks are artifacts of sample preparation, we now compare never-dried Arabidopsis primary walls with dehydrated and rehydrated samples. One-dimensional 13C spectra, two-dimensional 13C-13C correlation spectra, water-polysaccharide correlation spectra, and dynamics data all indicate that the structure, mobility, and intermolecular contacts of the polysaccharides are indistinguishable between never-dried and rehydrated walls. Moreover, a partially depectinated cell wall in which 40% of homogalacturonan is extracted retains cellulose-pectin cross peaks, indicating that the cellulose-pectin contacts are not due to molecular crowding. The cross peaks are observed both at −20°C and at ambient temperature, thus ruling out freezing as a cause of spatial contacts. These results indicate that rhamnogalacturonan I and a portion of homogalacturonan have significant interactions with cellulose microfibrils in the native primary wall. This pectin-cellulose association may be formed during wall biosynthesis and may involve pectin entrapment in or between cellulose microfibrils, which cannot be mimicked by in vitro binding assays.Despite decades of research, many aspects of the three-dimensional (3D) structure of the primary cell walls (CWs) of eudicotyledons and noncommelinid monocotyledons remain uncertain because of the insoluble, amorphous, and complex nature of CW polymers and their tight interactions in muro (Jarvis, 1992; Cosgrove, 2005). Unlike massively cross-linked bacterial CWs, noncovalent interactions between polysaccharides play major structural roles in plant CWs (Albersheim et al., 2011). Our concepts of the arrangement and interactions of cellulose, hemicellulose, and pectins in the CW have largely been shaped by electron microscopy and by chemical or enzymatic dissections, approaches that are limited in their ability to resolve specific noncovalent interactions or that result in considerable perturbation of native CW structures (McCann et al., 1990; Talbott and Ray, 1992; Carpita and Gibeaut, 1993). Recently, two-dimensional (2D) and 3D solid-state nuclear magnetic resonance (SSNMR) techniques have been applied to study the 3D architecture of plant CW polysaccharides (Dick-Pérez et al., 2011, 2012; Wang et al., 2014). The SSNMR experiments detect polysaccharide conformation, mobility, and subnanometer interatomic contacts, and the measurements were done on CW samples that were minimally perturbed from the native state. Interestingly, the 2D and 3D 13C correlation spectra showed clear cross peaks between cellulose and Rha or galacturonic acid residues of pectic polysaccharides. Together with molecular dynamics simulations (Matthews et al., 2006; Zhao et al., 2014) and enzymatic and biomechanical analyses (Park and Cosgrove, 2012a, 2012b), these data led to a revised concept of the CW architecture in which cellulose makes physical contact with both xyloglucan and pectins and in which CW extensibility is controlled at limited biomechanical hotspots where multiple microfibrils come into close contact (Cosgrove, 2014). This new model contrasts with the prevailing tethered network model, which depicts a scaffold of cellulose microfibrils tethered by xyloglucans and embedded in a separate but coextensive gel-like phase made of pectin (Carpita and Gibeaut, 1993; Somerville et al., 2004; Cosgrove, 2005). Additional SSNMR data demonstrated that expansins target limited sites in the CW with altered cellulose conformation and enriched xyloglucan content (Wang et al., 2013), sites that are very similar to the hypothesized biomechanical hotspots (Park and Cosgrove, 2012b).Despite these advances, the structural roles of pectins in the CW remain puzzling and unresolved. The major pectic polysaccharides include homogalacturonan (HG), rhamnogalacturonan I (RGI), and rhamnogalacturonan II (Caffall and Mohnen, 2009). Neutral pectins, including arabinans and galactans, are found as side chains of RGI and possibly as free polymers. Rhamnogalacturonan II units are cross linked by borate diesters that influence the strength and porosity of primary CWs (Ishii and Matsunaga, 1996; Fleischer et al., 1999). Some pectins may be covalently linked to arabinogalactan proteins (Tan et al., 2013) or xyloglucan (Popper and Fry, 2008; Cornuault et al., 2014), but the prevalence and structural significance of such hybrid molecules remain to be determined. Early extraction studies of a number of plant CWs showed that pectic polysaccharides remained with cellulose in the unextracted residue (Ryden and Selvendran, 1990; Goodneratne et al., 1994), and galactan side chains remain highly mobile in extracted onion (Allium cepa) CWs, which indirectly suggests that pectins may attach to cellulose via the rhamnogalacturonan backbone or HG (Foster et al., 1996). However, these results were not accounted for in traditional CW structural models (Cosgrove, 2001), which depict pectins as a relatively compliant gel-like network in which the cellulose-xyloglucan network is embedded. In this view, interactions between the two networks are limited to polymer entanglements and do not include extensive noncovalent interactions. Supporting this model are in vitro binding studies that indicate negligible affinity between pectins and cellulose, in contrast to the high affinity between xyloglucan and cellulose (Zykwinska et al., 2005, 2008). The recent 2D and 3D SSNMR spectra showed a significant number of cross peaks between cellulose and pectins (Dick-Perez et al., 2011, 2012), indicating that some of the pectic sugars come into subnanometer contact with cellulose. While NMR cross peaks report interatomic distances and not binding energies, the cross peaks nevertheless suggest that significant interactions exist between cellulose and some of the pectins. Such interactions, if true, would disagree with the in vitro binding assays but would be consistent with the early extraction results.An unexpected result such as this one naturally meets with skepticism and a search for alternative explanations. One possibility is that the preparation of the first CW samples, which included washing with ethanol and other organic solvents followed by air drying and rehydration, might have caused pectins to precipitate irreversibly onto cellulose and other wall components, thus giving rise to artificial spatial contacts between different polymers. A second possibility is that the intermolecular contacts result from molecular crowding of the wall polymers rather than specific interactions stabilized by hydrogen bonding, van der Waals interactions, or other noncovalent interactions (Cosgrove, 2014). Binding assays indicate that neutral arabinan and galactan side chains of RGI moderately bind cellulose in vitro (Zykwinska et al., 2005, 2008), leading to the proposal that pectic side chains may function as structural linkers between cellulose microfibrils (Zykwinska et al., 2007; Peaucelle et al., 2012). Although the mechanical significance of pectins was not confirmed by experiments in which wild-type Arabidopsis (Arabidopsis thaliana) CWs were subjected to pectin-loosening treatments, pectins assumed a larger mechanical role in xyloglucan-deficient CWs (Park and Cosgrove, 2012a).In this work, we investigate the nature of the pectin-cellulose contacts in greater depth by carrying out SSNMR experiments on a new set of CW samples that allows us to assess the effects of sample preparation, hydration history, temperature, and polysaccharide density on the cellulose-pectin spatial contacts. We prepared Arabidopsis primary CWs that were never dried throughout the sample preparation. We show that 13C chemical shifts, intramolecular and intermolecular cross peaks, segmental order parameters, and nuclear spin relaxation times are indistinguishable between the never-dried and rehydrated walls. Difference 2D spectra (Wang et al., 2015) that suppress intramolecular cross peaks are obtained that better reveal the intermolecular cross peaks between cellulose and pectins. Variable-temperature experiments show that cross peaks persist at ambient temperature, thus ruling out artifactual interaction by freezing. Finally, a partially HG-extracted CW retains cellulose-pectin correlation peaks, indicating that molecular crowding is not the main cause of the intermolecular contacts. Taken together, these data indicate that cellulose-pectin spatial proximity is an intrinsic feature of the never-dried primary CW, and some of the pectins may be entrapped within or between cellulose microfibrils.  相似文献   

17.
18.
In Arabidopsis thaliana, responses to pathogen-associated molecular patterns (PAMPs) are mediated by cell surface pattern recognition receptors (PRRs) and include the accumulation of reactive oxygen species, callose deposition in the cell wall, and the generation of the signal molecule salicylic acid (SA). SA acts in a positive feedback loop with ACCELERATED CELL DEATH6 (ACD6), a membrane protein that contributes to immunity. This work shows that PRRs associate with and are part of the ACD6/SA feedback loop. ACD6 positively regulates the abundance of several PRRs and affects the responsiveness of plants to two PAMPs. SA accumulation also causes increased levels of PRRs and potentiates the responsiveness of plants to PAMPs. Finally, SA induces PRR- and ACD6-dependent signaling to induce callose deposition independent of the presence of PAMPs. This PAMP-independent effect of SA causes a transient reduction of PRRs and ACD6-dependent reduced responsiveness to PAMPs. Thus, SA has a dynamic effect on the regulation and function of PRRs. Within a few hours, SA signaling promotes defenses and downregulates PRRs, whereas later (within 24 to 48 h) SA signaling upregulates PRRs, and plants are rendered more responsive to PAMPs. These results implicate multiple modes of signaling for PRRs in response to PAMPs and SA.  相似文献   

19.
20.
Plasmodesmata (PD) form tubular connections that function as intercellular communication channels. They are essential for transporting nutrients and for coordinating development. During cytokinesis, simple PDs are inserted into the developing cell plate, while during wall extension, more complex (branched) forms of PD are laid down. We show that complex PDs are derived from existing simple PDs in a pattern that is accelerated when leaves undergo the sink–source transition. Complex PDs are inserted initially at the three-way junctions between epidermal cells but develop most rapidly in the anisocytic complexes around stomata. For a quantitative analysis of complex PD formation, we established a high-throughput imaging platform and constructed PDQUANT, a custom algorithm that detected cell boundaries and PD numbers in different wall faces. For anticlinal walls, the number of complex PDs increased with increasing cell size, while for periclinal walls, the number of PDs decreased. Complex PD insertion was accelerated by up to threefold in response to salicylic acid treatment and challenges with mannitol. In a single 30-min run, we could derive data for up to 11k PDs from 3k epidermal cells. This facile approach opens the door to a large-scale analysis of the endogenous and exogenous factors that influence PD formation.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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