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.
In the primary walls of growing plant cells, the glucose polymer cellulose is assembled into long microfibrils a few nanometers in diameter. The rigidity and orientation of these microfibrils control cell expansion; therefore, cellulose synthesis is a key factor in the growth and morphogenesis of plants. Celery (Apium graveolens) collenchyma is a useful model system for the study of primary wall microfibril structure because its microfibrils are oriented with unusual uniformity, facilitating spectroscopic and diffraction experiments. Using a combination of x-ray and neutron scattering methods with vibrational and nuclear magnetic resonance spectroscopy, we show that celery collenchyma microfibrils were 2.9 to 3.0 nm in mean diameter, with a most probable structure containing 24 chains in cross section, arranged in eight hydrogen-bonded sheets of three chains, with extensive disorder in lateral packing, conformation, and hydrogen bonding. A similar 18-chain structure, and 24-chain structures of different shape, fitted the data less well. Conformational disorder was largely restricted to the surface chains, but disorder in chain packing was not. That is, in position and orientation, the surface chains conformed to the disordered lattice constituting the core of each microfibril. There was evidence that adjacent microfibrils were noncovalently aggregated together over part of their length, suggesting that the need to disrupt these aggregates might be a constraining factor in growth and in the hydrolysis of cellulose for biofuel production.Growth and form in plants are controlled by the precisely oriented expansion of the walls of individual cells. The driving force for cell expansion is osmotic, but the rate and direction of expansion are controlled by the mechanical properties of the cell wall (Szymanski and Cosgrove, 2009). Expanding, primary cell walls are nanocomposite materials in which long microfibrils of cellulose, a few nanometers in diameter, run through a hydrated matrix of xyloglucans, pectins, and other polymers (Knox, 2008; Mohnen, 2008; Szymanski and Cosgrove, 2009; Scheller and Ulvskov, 2010). Native cellulose microfibrils are partially crystalline (Nishiyama, 2009; Fernandes et al., 2011). Formerly, primary wall cellulose was thought to have a unique crystal structure called cellulose IV1 (Dinand et al., 1996), but NMR evidence suggests the presence of forms similar to the better characterized cellulose Iα and Iβ crystalline forms together with large quantities of less ordered cellulose (Wickholm et al., 1998; Sturcová et al., 2004; Wada et al., 2004). Nevertheless, cellulose is much more ordered than any other component of the primary cell wall (Bootten et al., 2004), in keeping with its key role of providing strength and controlling growth.The stiffness of the cell wall is greatest in the direction of the cellulose microfibrils, where growth is directional and the predominant microfibril orientation is usually transverse to the growth direction (Green, 1999; MacKinnon et al., 2006; Szymanski and Cosgrove, 2009). Expansion of the cell wall then requires either widening of the spacing between microfibrils (Marga et al., 2005) or slippage between them (Cosgrove, 2005), or both, and the microfibrils reorient toward the direction of growth (Anderson et al., 2010). Polymer cross bridges between microfibrils (McCann et al., 1990) are thought to resist these deformations of the cell wall nanostructure and, thus, to control the rate of growth. Until recently, most attention was focused on bridging xyloglucans, hydrogen bonded to microfibril surfaces (Scheller and Ulvskov, 2010). However, there is evidence that not all xyloglucans are appropriately positioned (Fujino et al., 2000; Park and Cosgrove, 2012a) and that other bridging polymers may be involved (Zykwinska et al., 2007). It has also been suggested that bundles of aggregated microfibrils, not single microfibrils, might be the key structural units in primary cell walls (Anderson et al., 2010), as in wood (Fahlén and Salmén, 2005; Fernandes et al., 2011). If so, single microfibrils could bridge between microfibril bundles. In summary, the growth of plant cells is not well understood, and we need more information on how cellulose orientation is controlled and on the nature of the bridging polymers, the cellulose surfaces to which these polymers bind, and the cohesion between microfibril surfaces that might mediate aggregation.Cellulose microfibrils are synthesized at the cell surface by large enzyme complexes having hexagonal symmetry, sometimes called “rosettes” (Somerville, 2006). Each complex contains multiple cellulose synthases that differ between primary cell walls and wood, although the appearance of the complexes is similar (Somerville, 2006; Atanassov et al., 2009). The simultaneous synthesis, from the same end, of all the chains in a native cellulose microfibril is why they are parallel (Nishiyama et al., 2002, 2003), in contrast to the entropically favored antiparallel structure found in man-made celluloses like rayon (Langan et al., 2001). The number of chains in a microfibril and the number of cellulose synthases in the synthetic complex are evidently related. It is commonly assumed that the number of chains is divisible by six, matching the hexagonal rosette symmetry, and 36-chain models (Himmel et al., 2007) bounded by the hydrophilic [110] and [1-10] crystal faces, as in algal celluloses (Bergenstråhle et al., 2008), have been widely adopted. The assembly and orientation of cellulose are connected, as several cellulose synthase mutants have phenotypes defective in cellulose orientation and plant form as well as depleted in cellulose content (Paredez et al., 2008). In certain other mutant lines, the crystallinity of the microfibrils appears to be affected (Fujita et al., 2011; Harris et al., 2012; Sánchez-Rodríguez et al., 2012).Therefore, a detailed understanding of the structure of primary wall cellulose microfibrils would help us to understand cellulose synthesis as well as the growth and structural mechanics of living plants (Burgert and Fratzl, 2009). Primary cell walls and their cellulose skeletons also affect food quality characteristics like the crispness of salad vegetables and apples (Malus domestica; Jarvis, 2011). When biofuels are produced from lignocellulosic biomass, lignification leads to recalcitrance (Himmel et al., 2007), but some of the cell types in Miscanthus spp., switchgrass (Panicum virgatum), and arable crop residues have only primary walls with no lignin, and recalcitrance then depends on the nature of the cellulose microfibrils (Beckham et al., 2011).A relatively detailed structure has recently been proposed for the microfibrils of spruce (Picea spp.) wood (Fernandes et al., 2011), which are 3.0 nm in diameter, allowing space for only about 24 cellulose chains. Evidence from x-ray diffraction supported a “rectangular” shape (Matthews et al., 2006) bounded by the [010] and [200] faces. There was considerable disorder increasing toward the surface, and the microfibrils were aggregated into bundles about 15 to 20 nm across, with some, but not all, of the lateral interfaces being resistant to water (Fernandes et al., 2011). Disordered domains are a feature of other strong biological materials such as spider silk (van Beek et al., 2002).Therefore, it is of interest whether any of these features of wood cellulose might also be found in the cellulose microfibrils of primary (growing) cell walls. It would be particularly useful to characterize the disorder known to be present in primary wall microfibrils, that is, to define how cellulose that is not measured as “crystalline” differs from crystalline cellulose. Many of the experiments leading toward a structure for wood cellulose were dependent on exceptionally uniform orientation of the cellulose microfibrils (Sturcová et al., 2004; Fernandes et al., 2011). However, in growing cell walls, the microfibrils are not uniformly oriented. When microfibrils are first laid down at the inner face of the primary cell wall, their orientation is normally transverse to the direction of growth, but as the cell wall expands, the microfibrils reorient so that the orientation distribution, integrated across the thickness of the expanded cell wall, becomes progressively closer to random (Cosgrove, 2005; MacKinnon et al., 2006).This technical problem does not apply to the cell walls of celery (Apium graveolens) collenchyma, which are similar in composition to other primary cell walls but have their microfibrils oriented relatively uniformly along the cell axis (Sturcová et al., 2004; Kennedy et al., 2007a, 2007b). Some structural information on celery collenchyma cellulose has already been derived from spectroscopic and scattering experiments (Sturcová et al., 2004; Kennedy et al., 2007a, 2007b), confirming the disorder expected in a primary wall cellulose. Some of these experiments were analogous to what has been done on spruce cellulose (Fernandes et al., 2011), but insufficient data are available to specify the number of chains in each primary wall microfibril, the nature and location of the disorder, and the presence or absence of direct contact between microfibrils. Here, we report x-ray and neutron scattering and spectroscopic experiments addressing these questions and leading to a proposed structure for primary wall cellulose microfibrils. Characterizing a structure containing so much disorder presented unusual challenges, but disorder appears to be central to the enigmatic capacity of primary wall cellulose to provide high strength and yet to permit and control growth.  相似文献   

2.
3.
The ordered arrangement of cortical microtubules in growing plant cells is essential for anisotropic cell expansion and, hence, for plant morphogenesis. These arrays are dismantled when the microtubule cytoskeleton is rearranged during mitosis and reassembled following completion of cytokinesis. The reassembly of the cortical array has often been considered as initiating from a state of randomness, from which order arises at least partly through self-organizing mechanisms. However, some studies have shown evidence for ordering at early stages of array assembly. To investigate how cortical arrays are initiated in higher plant cells, we performed live-cell imaging studies of cortical array assembly in tobacco (Nicotiana tabacum) Bright Yellow-2 cells after cytokinesis and drug-induced disassembly. We found that cortical arrays in both cases did not initiate randomly but with a significant overrepresentation of microtubules at diagonal angles with respect to the cell axis, which coincides with the predominant orientation of the microtubules before their disappearance from the cell cortex in preprophase. In Arabidopsis (Arabidopsis thaliana) root cells, recovery from drug-induced disassembly was also nonrandom and correlated with the organization of the previous array, although no diagonal bias was observed in these cells. Surprisingly, during initiation, only about one-half of the new microtubules were nucleated from locations marked by green fluorescent protein-γ-tubulin complex protein2-tagged γ-nucleation complexes (γ-tubulin ring complex), therefore indicating that a large proportion of early polymers was initiated by a noncanonical mechanism not involving γ-tubulin ring complex. Simulation studies indicate that the high rate of noncanonical initiation of new microtubules has the potential to accelerate the rate of array repopulation.Higher plant cells feature ordered arrays of microtubules at the cell cortex (Ledbetter and Porter, 1963) that are essential for cell and tissue morphogenesis, as revealed by disruption of cortical arrays by drugs that cause microtubule depolymerization (Green, 1962) or stabilization (Weerdenburg and Seagull, 1988) and by loss-of-function mutations in a wide variety of microtubule-associated proteins (Baskin, 2001; Whittington et al., 2001; Buschmann and Lloyd, 2008; Lucas et al., 2011). The structure of these arrays is thought to control the pattern of cell growth primarily by its role in the deposition of cellulose microfibrils, the load-bearing component of the cell wall (Somerville, 2006). Functional relations between cortical microtubules and cellulose microfibrils have been proposed since the early sixties, even before cortical microtubules had been visualized (Green, 1962). Recent live-cell imaging studies have confirmed that cortical microtubules indeed guide the movement of cellulose synthase complexes that produce cellulose microfibrils (Paredez et al., 2006) and have shown further that microtubules position the insertion of most cellulose synthase complexes into the plasma membrane (Gutierrez et al., 2009). These activities of ordered cortical microtubules are proposed to facilitate the organization of cell wall structure, creating material properties that underlie cell growth anisotropy.While organization of the interphase cortical array appears to be essential for cell morphogenesis, this organization is disrupted during the cell cycle as microtubules are rearranged to create the preprophase band, spindle, and phragmoplast during mitosis and cytokinesis (for review, see Wasteneys, 2002). Upon completion of cytokinesis, an organized interphase cortical array is regenerated, but the pathway for this reassembly is not well understood.The plant interphase microtubule array is organized and maintained without centrosomes as organizing centers (for review, see Wasteneys, 2002; Bartolini and Gundersen, 2006; Ehrhardt and Shaw, 2006), and microtubule self-organization is proposed to play an important role in cortical microtubule array ordering (Dixit and Cyr, 2004). In electron micrographs, microtubules have been observed to be closely associated with the plasma membrane (Hardham and Gunning, 1978), and live-cell imaging provides evidence for attachment of microtubules to the cell cortex (Shaw et al., 2003; Vos et al., 2004). The close association to the plasma membrane restricts the cortical microtubules to a quasi two-dimensional plane where they interact through polymerization-driven collisions (Shaw et al., 2003; Dixit and Cyr, 2004). Microtubule encounters at shallow angles (<40°) have a high probability of leading to bundling, while microtubule encounters at steeper angles most likely result in induced catastrophes or microtubule crossovers (Dixit and Cyr, 2004). Several computational modeling studies have since shown that these types of interactions between surface-bound dynamical microtubules can indeed explain spontaneous coalignment of microtubules (Allard et al., 2010; Eren et al., 2010; Hawkins et al., 2010; Tindemans et al., 2010).The question of how the orientation of the cortical array is established with respect to the cell axis is less well understood. One possibility is that microtubules are selectively destabilized with respect to cellular coordinates (Ehrhardt and Shaw, 2006). Indeed, recent results from biological observations and modeling suggest that catastrophic collisions induced at the edges between cell faces or heighted catastrophe rates in cell caps could be sufficient to selectively favor microtubules in certain orientation and hence determine the final orientation of the array (Allard et al., 2010; Eren et al., 2010; Ambrose et al., 2011; Dhonukshe et al., 2012).To date, all models of cortical array assembly assume random initial conditions. However, experimental work by Wasteneys and Williamson (1989a, 1989b) in Nitella tasmanica showed that, during array reassembly after drug-induced disruption, microtubules were initially transverse. This was followed by a less ordered phase and later by the acquisition of the final transverse order. A nonrandom initial ordering was also observed in tobacco (Nicotiana tabacum) Bright Yellow-2 (BY-2) cells by Kumagai et al. (2001), who concluded that the process of transverse array establishment starts with longitudinal order but did not provide quantitative data for the process of array assembly. The initial conditions for the cortical microtubule array formation are important to consider, as they may strongly influence the speed at which order is established and could even prevent it from being established over a biologically relevant time scale.In this study, we used live-cell imaging to follow and record the whole transition from the cortical microtubule-free state to the final transverse array and used digital tracking algorithms to quantify the microtubule order. Nucleation stands out as a central process to characterize during array initiation. Lacking a central body to organize microtubule nucleations, the higher plant cell has dispersed nucleation complexes (Wasteneys and Williamson, 1989a, 1989b; Chan et al., 2003; Shaw et al., 2003; Murata et al., 2005; Pastuglia et al., 2006; Nakamura et al., 2010). Therefore, we performed high time resolution observations to quantify nucleation complex recruitment, nucleation rates, and microtubule nucleation angles. We found evidence for a highly nonrandom initial ordering state that features diagonal microtubule orientation and an atypical microtubule initiation mechanism. Simulation analysis indicates that these atypical nucleations have the potential to accelerate the recovery of cortical array density.  相似文献   

4.
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   

5.
6.
The interaction between cellulose and xylan is important for the load-bearing secondary cell wall of flowering plants. Based on the precise, evenly spaced pattern of acetyl and glucuronosyl (MeGlcA) xylan substitutions in eudicots, we recently proposed that an unsubstituted face of xylan in a 2-fold helical screw can hydrogen bond to the hydrophilic surfaces of cellulose microfibrils. In gymnosperm cell walls, any role for xylan is unclear, and glucomannan is thought to be the important cellulose-binding polysaccharide. Here, we analyzed xylan from the secondary cell walls of the four gymnosperm lineages (Conifer, Gingko, Cycad, and Gnetophyta). Conifer, Gingko, and Cycad xylan lacks acetylation but is modified by arabinose and MeGlcA. Interestingly, the arabinosyl substitutions are located two xylosyl residues from MeGlcA, which is itself placed precisely on every sixth xylosyl residue. Notably, the Gnetophyta xylan is more akin to early-branching angiosperms and eudicot xylan, lacking arabinose but possessing acetylation on alternate xylosyl residues. All these precise substitution patterns are compatible with gymnosperm xylan binding to hydrophilic surfaces of cellulose. Molecular dynamics simulations support the stable binding of 2-fold screw conifer xylan to the hydrophilic face of cellulose microfibrils. Moreover, the binding of multiple xylan chains to adjacent planes of the cellulose fibril stabilizes the interaction further. Our results show that the type of xylan substitution varies, but an even pattern of xylan substitution is maintained among vascular plants. This suggests that 2-fold screw xylan binds hydrophilic faces of cellulose in eudicots, early-branching angiosperm, and gymnosperm cell walls.The plant secondary cell wall is a complex network of various polysaccharides and phenolic compounds that act in concert to provide strength to the cell wall (Kumar et al., 2016). Cellulose, formed of strong crystalline fibrils of linear β1,4 glucan, comprises about 40% of dry plant biomass. The other secondary cell wall polysaccharides, largely xylan and glucomannan, comprise about 30% of dry plant biomass. The abundance and structure of these hemicelluloses vary with plant species and tissues, but they have in common that they are tightly associated with cellulose. It is believed that the most important biological role of hemicelluloses is their contribution to strengthening the cell wall by interaction with cellulose and, in some walls, with lignin (Scheller and Ulvskov, 2010). However, it is unclear how hemicelluloses interact with cellulose in the cell wall (Cosgrove and Jarvis, 2012). In this work, we were interested in the interaction of cellulose with xylan, one of the most abundant polysaccharides in nature.To understand polysaccharide interactions in the cell wall, we need to know not only the hemicellulose primary structure, but also the conformation of the polysaccharide chains. The glucan chains in cellulose are similar to a flat ribbon known as a 2-fold helical screw and associate through lateral hydrogen bonding into sheets. The sheets of glucan chains stack on top of each other, resulting in highly ordered crystalline cellulose. It is still unclear how many glucan chains form a microfibril and whether the microfibril has a hexagonal or rectangular cross section. However, recent studies shed some light onto these questions (Fernandes et al., 2011; Newman et al., 2013; Thomas et al., 2013; Cosgrove, 2014; Oehme et al., 2015; Thomas et al., 2015; Wang and Hong, 2016; Vandavasi et al., 2016). Whichever model is favored, hydrophobic (e.g. 100 or 200) and hydrophilic (e.g. 110 or 010) crystal faces are exposed for interaction with other molecules such as hemicelluloses (Zhao et al., 2014; Cosgrove, 2014; Li et al., 2015).The β1,4 xylan backbone is always further modified, often by acetyl (Ac), arabinosyl (Ara), and glucuronosyl (MeGlcA) side-chain substitutions. These substitutions are supposed to be necessary to maintain xylan solubility (Mikkelsen et al., 2015). Unsubstituted xylan forms crystalline fibers of chains adopting a 3-fold screw helix (Nieduszynski and Marchessault, 1971). Consequently, xylan substitutions are essential for xylan function and vascular plant viability (Mortimer et al., 2010; Xiong et al., 2013, 2015). In vitro experiments and in silico modeling suggest xylan interacts with cellulose, and it is widely accepted that this is partly through interactions on the hydrophobic faces of the cellulose fibrils (Bosmans et al., 2014; Köhnke et al., 2011; Kabel et al., 2007; Busse-Wicher et al., 2014). In contrast to the binding to the hydrophobic faces, the backbones of highly substituted hemicelluloses are thought to be unable to hydrogen bond effectively with the hydrophilic surfaces of cellulose fibrils because of steric hindrance. For example, hydrogen bonding of the xyloglucan backbone to cellulose would be blocked by steric restrictions of the side chains (Finkenstadt et al., 1995; Zhang et al., 2011). How then does the naturally occurring, highly substituted, xylan interact with cellulose? Our recent findings in the eudicot Arabidopsis (Arabidopsis thaliana) revealed that the majority of xylan bears substitutions solely on alternate xylosyl residues. Every second Xyl is acetylated (Busse-Wicher et al., 2014; Chong et al., 2014), and MeGlcA side chains reside on evenly spaced xylosyl residues, largely at 6-, 8-, 10-, or 12-residue intervals (Bromley et al., 2013). In this scenario, on a xylan backbone in the ribbon-like 2-fold helical screw conformation, all the decorations will face one side, creating an unsubstituted xylan surface. Therefore, in addition to forming stacking interactions on the hydrophobic surface, this xylan structure is compatible with hydrogen bonding to the hydrophilic surface of cellulose (Busse-Wicher et al., 2014; Busse-Wicher et al., 2016).Both xylan and glucomannan are substantial components of vascular plant secondary cell walls (Timell, 1967; Willfor et al., 2005; McKee et al., 2016). In conifers (gymnosperms, Pinales), the main hemicellulose is glucomannan, but eudicots possess relatively little glucomannan (Scheller and Ulvskov, 2010; Huang et al., 2015), and the secondary cell walls are dominated by xylan, suggesting xylan might have adopted additional functions in these flowering plants that produce hardwoods (Dammström et al., 2009). Conifers, providing softwood for the paper, pulp, and construction industries, are of major ecological and economical value. Consequently, understanding the function and architecture of the cell wall components of softwoods and hardwoods is of great importance.To investigate whether the precise arrangement of xylan decorations on evenly spaced xylosyl residues, as seen in eudicots, is a novel feature of hardwood xylan, we analyzed the pattern of xylan substitution in various gymnosperms and angiosperms. In addition to conifers, there are three further gymnosperm lineages: Cycad, Gingko and Gnetophyta (Fig. 1). There has been a debate whether Gnetophyta are the gymnosperm lineage most closely related to the angiosperms (Davis and Schaefer, 2011; Uddenberg et al., 2015). There are few studies across gymnosperm lineages to determine any divergence in the structure of xylans.Open in a separate windowFigure 1.Schematic representation of the phylogenetic relationship between gymnosperm and angiosperm species studied in this work. The distances do not correspond to phylogenetic distances.Our work shows that some gymnosperm xylans have decorations and decoration patterns that are different to those of eudicot xylans. Nevertheless, these modifications largely reside on even xylosyl residues on the backbone. Molecular dynamics simulations support the hypothesis that this highly conserved organization of substitutions allows an unsubstituted surface of xylan to bind stably to hydrophilic faces of cellulose fibrils.  相似文献   

7.
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   

8.
9.
The chemical and structural organization of the plant cell wall was examined in Zinnia elegans tracheary elements (TEs), which specialize by developing prominent secondary wall thickenings underlying the primary wall during xylogenesis in vitro. Three imaging platforms were used in conjunction with chemical extraction of wall components to investigate the composition and structure of single Zinnia TEs. Using fluorescence microscopy with a green fluorescent protein-tagged Clostridium thermocellum family 3 carbohydrate-binding module specific for crystalline cellulose, we found that cellulose accessibility and binding in TEs increased significantly following an acidified chlorite treatment. Examination of chemical composition by synchrotron radiation-based Fourier-transform infrared spectromicroscopy indicated a loss of lignin and a modest loss of other polysaccharides in treated TEs. Atomic force microscopy was used to extensively characterize the topography of cell wall surfaces in TEs, revealing an outer granular matrix covering the underlying meshwork of cellulose fibrils. The internal organization of TEs was determined using secondary wall fragments generated by sonication. Atomic force microscopy revealed that the resulting rings, spirals, and reticulate structures were composed of fibrils arranged in parallel. Based on these combined results, we generated an architectural model of Zinnia TEs composed of three layers: an outermost granular layer, a middle primary wall composed of a meshwork of cellulose fibrils, and inner secondary wall thickenings containing parallel cellulose fibrils. In addition to insights in plant biology, studies using Zinnia TEs could prove especially productive in assessing cell wall responses to enzymatic and microbial degradation, thus aiding current efforts in lignocellulosic biofuel production.The organization and molecular architecture of plant cell walls represent some of the most challenging problems in plant biology. Although much is known about general aspects of assembly and biosynthesis of the plant cell wall, the detailed three-dimensional molecular cell wall structure remains poorly understood. The highly complex and dynamic nature of the plant cell wall has perhaps limited the generation of such detailed structural models. This information is pivotal for the successful implementation of novel approaches for conversion of biomass to liquid biofuels, given that one of the critical processing steps in biomass conversion involves systematic deconstruction of cell walls. Therefore, a comprehensive understanding of the architecture and chemical composition of the plant cell wall will not only help develop molecular-scale models, but will also help improve the efficiency of biomass deconstruction.The composition and molecular organization of the cell wall is species and cell type dependent (Vorwerk et al., 2004). Thus, the development of a model plant system, which utilizes a single cell type, has enhanced our capacity to understand cell wall architecture. The ability to generate a population of single Zinnia elegans plant cells that were synchronized throughout cell wall deposition during xylogenesis was developed in the 1980s (Fukuda and Komamine, 1980). Mesophyll cells isolated from the leaves of Zinnia and cultured in the presence of phytohormones will transdifferentiate into tracheary elements (TEs), which are individual components of the xylem vascular tissue (Fukuda and Komamine, 1980). During this transdifferentiation process, TEs gradually develop patterned secondary wall thickenings, commonly achieving annular, spiral, reticulate, scalariform, and pitted patterns (Bierhorst, 1960; Falconer and Seagull, 1988; Roberts and Haigler, 1994). These secondary wall thickenings serve as structural reinforcements that add strength and rigidity to prevent the collapse of the xylem under the high pressure created by fluid transport. During the final stages of transdifferentiation, TEs accumulate lignin in their secondary walls and undergo programmed cell death, which results in the removal of all cell contents, leaving behind a “functional corpse” (Roberts and McCann, 2000; Fukuda, 2004).In broad terms, the primary cell wall of higher plants is mainly composed of three types of polysaccharides: cellulose, hemicelluloses, and pectins (Cosgrove, 2005). Cellulose is composed of unbranched β-1,4-Glc chains that are packed together into fibrils by intermolecular and intramolecular hydrogen bonding. Hemicelluloses and pectins are groups of complex polysaccharides that are primarily composed of xyloglucans/xylans and galacturonans, respectively. Hemicelluloses are involved in cross-linking and associating with cellulose microfibrils, while pectins control wall porosity and help bind neighboring cells together. The patterned deposits of secondary wall in Zinnia TEs primarily consist of cellulose microfibrils, along with hemicelluloses, and also lignin, a complex aromatic polymer that is characteristic of secondary walls and provides reinforcement (Turner et al., 2007). All the molecular components in the cell wall correspond to a multitude of different polysaccharides, phenolic compounds, and proteins that become arranged and modified in muro, yielding a structure of great strength and resistance to degradation.Currently, electron microscopy is the primary tool for structural studies of cell walls and has provided remarkable information regarding wall organization. Fast-freeze deep-etch electron microscopy in combination with chemical and enzymatic approaches have generated recent models of the architecture of the primary wall (McCann et al., 1990; Carpita and Gibeaut, 1993; Nakashima et al., 1997; Fujino et al., 2000; Somerville et al., 2004). Direct visualization of secondary wall organization has been focused toward the examination of multiple wall layers in wood cells (Fahlen and Salmen, 2005; Zimmermann et al., 2006). However, few studies have examined the secondary wall, so our knowledge regarding the higher order architecture of this type of wall is limited. Over the past few decades, atomic force microscopy (AFM) has provided new opportunities to probe biological systems with spatial resolution similar to electron microscopy techniques (Kuznetsov et al., 1997; Muller et al., 1999), with additional ease of sample preparation and the capability to probe living native structures. AFM has been successfully applied to studies of the high-resolution architecture, assembly, and structural dynamics of a wide range of biological systems (Hoh et al., 1991; Crawford et al., 2001; Malkin et al., 2003; Plomp et al., 2007), thus enabling the observation of the ultrastructure of the plant cell wall, which is of particular interest to us (Kirby et al., 1996; Morris et al., 1997; Davies and Harris, 2003; Yan et al., 2004; Ding and Himmel, 2006).To generate more detailed structural models, knowledge about the structural organization of the cell wall can be combined with spatial information about chemical composition. Instead of utilizing chromatography techniques to analyze cell wall composition by extracting material from bulk plant samples (Mellerowicz et al., 2001; Pauly and Keegstra, 2008), Fourier transform infrared (FTIR) spectromicroscopy can be used to directly probe for polysaccharide and aromatic molecules in native as well as treated plant material (Carpita et al., 2001; McCann et al., 2001). FTIR spectromicroscopy is not only able to identify chemical components in a specific system but also can determine their distribution and relative abundance. This technique also improves the sensitivity and spatial resolution of cellular components without the derivatization needed by chemical analysis using chromatography. Polysaccharide-specific probes, such as carbohydrate-binding modules (CBMs), can also be used to understand the chemical composition of the plant cell wall. CBMs are noncatalytic protein domains existing in many glycoside hydrolases. Based on their binding specificities, CBMs are generally categorized into three groups: surface-binding CBMs specific for insoluble cellulose surfaces, chain-binding CBMs specific for single chains of polysaccharides, and end-binding CBMs specific for the ends of polysaccharides or oligosaccharides. A surface-binding CBM with high affinity for the planar faces of crystalline cellulose (Tormo et al., 1996; Lehtio et al., 2003) has been fluorescently labeled and used to label crystals as well as plant tissue (Ding et al., 2006; Porter et al., 2007; Liu et al., 2009; Xu et al., 2009). The binding capacity of the CBM family has been further exploited for the detection of different polysaccharides, such as xylans and glucans, and can thus be used for the characterization of plant cell wall composition (McCartney et al., 2004, 2006).In this study, we used a combination of AFM, synchrotron radiation-based (SR)-FTIR spectromicroscopy, and fluorescence microscopy using a cellulose-specific CBM to probe the cell wall of Zinnia TEs. The Zinnia TE culture system proved ideal for observing the structure and chemical composition of the cell wall because it comprises a single homogeneous cell type, representing a simpler system compared with plant tissues, which may contain multiple cell types. Zinnia TEs were also advantageous because they were analyzed individually, and population statistics were generated based on specific conditions. Furthermore, cultured Zinnia TEs were used for the consistent production of cell wall fragments for analysis of the organization of internal secondary wall structures. In summary, we have physically and chemically dissected Zinnia TEs using a combination of imaging techniques that revealed primary and secondary wall structures and enabled the reconstruction of TE cell wall architecture.  相似文献   

10.
Tension wood is widespread in the organs of woody plants. During its formation, it generates a large tensile mechanical stress, called maturation stress. Maturation stress performs essential biomechanical functions such as optimizing the mechanical resistance of the stem, performing adaptive movements, and ensuring long-term stability of growing plants. Although various hypotheses have recently been proposed, the mechanism generating maturation stress is not yet fully understood. In order to discriminate between these hypotheses, we investigated structural changes in cellulose microfibrils along sequences of xylem cell differentiation in tension and normal wood of poplar (Populus deltoides × Populus trichocarpa ‘I45-51’). Synchrotron radiation microdiffraction was used to measure the evolution of the angle and lattice spacing of crystalline cellulose associated with the deposition of successive cell wall layers. Profiles of normal and tension wood were very similar in early development stages corresponding to the formation of the S1 and the outer part of the S2 layer. The microfibril angle in the S2 layer was found to be lower in its inner part than in its outer part, especially in tension wood. In tension wood only, this decrease occurred together with an increase in cellulose lattice spacing, and this happened before the G-layer was visible. The relative increase in lattice spacing was found close to the usual value of maturation strains, strongly suggesting that microfibrils of this layer are put into tension and contribute to the generation of maturation stress.Wood cells are produced in the cambium at the periphery of the stem. The formation of the secondary wall occurs at the end of cell elongation by the deposition of successive layers made of cellulose microfibrils bounded by an amorphous polymeric matrix. Each layer has a specific chemical composition and is characterized by a particular orientation of the microfibrils relative to the cell axis (Mellerowicz and Sundberg, 2008). Microfibrils are made of crystalline cellulose and are by far the stiffest constituent of the cell wall. The microfibril angle (MFA) in each layer is determinant for cell wall architecture and wood mechanical properties.During the formation of wood cells, a mechanical stress of a large magnitude, known as “maturation stress” or “growth stress” (Archer, 1986; Fournier et al., 1991), occurs in the cell walls. This stress fulfills essential biomechanical functions for the tree. It compensates for the comparatively low compressive strength of wood and thus improves the stem resistance against bending loads. It also provides the tree with a motor system (Moulia et al., 2006), necessary to maintain the stem at a constant angle during growth (Alméras and Fournier, 2009) or to achieve adaptive reorientations. In angiosperms, a large tensile maturation stress is generated by a specialized tissue called “tension wood.” In poplar (Populus deltoides × Populus trichocarpa), as in most temperate tree species, tension wood fibers are characterized by the presence of a specific layer, called the G-layer (Jourez et al., 2001; Fang et al., 2008), where the matrix is almost devoid of lignin (Pilate et al., 2004) and the microfibrils are oriented parallel to the fiber axis (Fujita et al., 1974). This type of reaction cell is common in plant organs whose function involves the bending or contraction of axes, such as tendrils, twining vines (Bowling and Vaughn, 2009), or roots (Fisher, 2008).The mechanism at the origin of tensile maturation stress has been the subject of a lot of controversy and is still not fully understood. However, several recent publications have greatly improved our knowledge about the ultrastructure, chemical composition, molecular activity, mechanical state, and behavior of tension wood. Different models have been proposed and discussed to explain the origin of maturation stress (Boyd, 1972; Bamber, 1987, 2001; Okuyama et al., 1994, 1995; Yamamoto, 1998, 2004; Alméras et al., 2005, 2006; Bowling and Vaughn, 2008; Goswami et al., 2008; Mellerowicz et al., 2008). The specific organization of the G-layer suggests a tensile force induced in the microfibrils during the maturation process. Different hypotheses have been proposed to explain this mechanism, such as the contraction of amorphous zones within the cellulose microfibrils (Yamamoto, 2004), the action of xyloglucans during the formation of microfibril aggregates (Nishikubo et al., 2007; Mellerowicz et al., 2008), and the effect of changes in moisture content stimulated by pectin-like substances (Bowling and Vaughn, 2008). A recent work (Goswami et al., 2008) argued an alternative model, initially proposed by Münch (1938), which proposed that the maturation stress originates in the swelling of the G-layer during cell maturation and is transmitted to the adjacent secondary layers, where the larger MFAs allow an efficient conversion of lateral stress into axial tensile stress. Although the proposed mechanism is not consistent with the known hygroscopic behavior of tension wood, which shrinks when it dries and not when it takes up water (Clair and Thibaut, 2001; Fang et al., 2007; Clair et al., 2008), this hypothesis focused attention on the possible role of cell wall layers other than the G-layer. As a matter of fact, many types of wood fibers lacking a G-layer are known to produce axial tensile stress, such as normal wood of angiosperms and conifers (Archer, 1986) and the tension wood of many tropical species (Onaka, 1949; Clair et al., 2006b; Ruelle et al., 2007), so that mechanisms strictly based on an action of the G-layer cannot provide a general explanation for the origin of tensile maturation stress in wood.In order to further understanding, direct observations of the mechanical state of the different cell wall layers and their evolution during the formation of the tension wood fibers are needed. X-ray diffraction can be used to investigate the orientation of microfibrils (Cave, 1966, 1997a, 1997b; Peura et al., 2007, 2008a, 2008b) and the lattice spacing of crystalline cellulose. The axial lattice spacing d004 is the distance between successive monomers along a cellulose microfibril and reflects its state of mechanical stress (Clair et al., 2006a; Peura et al., 2007). If cellulose microfibrils indeed support a tensile stress, they should be found in an extended state of deformation. Under this assumption, the progressive development of maturation stress during the cell wall formation should be accompanied by an increase in cellulose lattice spacing. Synchrotron radiation allows a reduction in the size of the x-ray beam to some micrometers while retaining a strong signal, whereby diffraction analysis can be performed at a very local scale (Riekel, 2000). This technique has been used to study sequences of wood cell development (Hori et al., 2000; Müller et al., 2002). In this study, we report an experiment where a microbeam was used to analyze the structural changes of cellulose in the cell wall layers of tension wood and normal wood fibers along the sequence of xylem cell differentiation extending from the cambium to mature wood (Fig. 1). The experiment was designed to make this measurement in planta, in order to minimize sources of mechanical disturbance and be as close as possible to the native mechanical state (Clair et al., 2006a). The 200 and 004 diffraction patterns of cellulose were analyzed to investigate the process of maturation stress generation in tension wood.Open in a separate windowFigure 1.Schematic of the experimental setup, showing the x-ray beam passing perpendicular to the longitudinal-radial plane of wood and the contribution of the 004 and 200 crystal planes to the diffraction pattern recorded by the camera. [See online article for color version of this figure.]  相似文献   

11.
The pollen tube is a cellular protuberance formed by the pollen grain, or male gametophyte, in flowering plants. Its principal metabolic activity is the synthesis and assembly of cell wall material, which must be precisely coordinated to sustain the characteristic rapid growth rate and to ensure geometrically correct and efficient cellular morphogenesis. Unlike other model species, the cell wall of the Arabidopsis (Arabidopsis thaliana) pollen tube has not been described in detail. We used immunohistochemistry and quantitative image analysis to provide a detailed profile of the spatial distribution of the major cell wall polymers composing the Arabidopsis pollen tube cell wall. Comparison with predictions made by a mechanical model for pollen tube growth revealed the importance of pectin deesterification in determining the cell diameter. Scanning electron microscopy demonstrated that cellulose microfibrils are oriented in near longitudinal orientation in the Arabidopsis pollen tube cell wall, consistent with a linear arrangement of cellulose synthase CESA6 in the plasma membrane. The cellulose label was also found inside cytoplasmic vesicles and might originate from an early activation of cellulose synthases prior to their insertion into the plasma membrane or from recycling of short cellulose polymers by endocytosis. A series of strategic enzymatic treatments also suggests that pectins, cellulose, and callose are highly cross linked to each other.Upon contact with the stigma, the pollen grain swells through water uptake and develops a cellular protrusion, the pollen tube. During its growth in planta, the pollen tube invades the transmitting tissue of the pistil and finds its way to the ovary to deliver the male gametes for double fertilization to happen (Heslop-Harrison, 1987). Depending on the species, pollen tubes can grow extremely rapidly both in planta and in in vitro conditions. To fulfill its biological function, the pollen tube has to (1) adhere to and invade transmitting tissues (Hill and Lord, 1987; Lennon et al., 1998), (2) provide physical protection to the sperm cells, and (3) control its own shape and invasive behavior (Parre and Geitmann, 2005b; Geitmann and Steer, 2006). For all of these functions, the pollen tube cell wall plays an important regulatory and structural role. Although the pollen tube does not form a conventional secondary cell wall layer, its wall is assembled in two phases. The “primary layer” is mainly formed of pectins and other matrix components secreted at the apical end of the cell. The “secondary layer” is assembled by the deposition of callose in more distal regions of the cell (Heslop-Harrison, 1987). Depending on the species, cellulose microfibrils have been found to be associated either with the outer pectic or with the inner callosic layer. Unlike most other plant cells, cellulose is not very abundant representing only 10% of total neutral polysaccharides in Nicotiana alata pollen tubes, whereas callose accounts for more than 80% in this species (Schlüpmann et al., 1994).The biochemical composition of the pollen tube cell wall has been well characterized in many species such as Lilium longiflorum (Lancelle and Hepler, 1992; Jauh and Lord, 1996), tobacco (Nicotiana tabacum; Kroh and Knuiman, 1982; Geitmann et al., 1995; Ferguson et al., 1998; Derksen et al., 2011), Petunia hybrida (Derksen et al., 1999), Pinus sylvestris (Derksen et al., 1999), and Solanum chacoense (Parre and Geitmann, 2005a). But for Arabidopsis (Arabidopsis thaliana), the model for plant molecular biology studies (Arabidopsis Genome Initiative, 2000), there is a striking lack of quantitative information concerning the composition of the pollen tube cell wall as well as the spatial distribution of its components. This is all the more surprising because numerous mutants defective in enzymes involved in cell wall synthesis exhibit a pollen tube phenotype (for example, Jiang et al., 2005; Nishikawa et al., 2005; Wang et al., 2011). Two studies have characterized the Arabidopsis pollen germinating in vitro (Derksen et al., 2002) and in vivo (Lennon and Lord, 2000), but both are qualitative rather than quantitative. A biochemical study by Dardelle and coworkers investigated the cell wall sugar composition in a more quantitative way but does not provide any detailed spatial information (Dardelle et al., 2010; Lehner et al., 2010). This lack of information is not surprising given that until recently Arabidopsis pollen was known to be rather challenging to germinate reproducibly in vitro and more difficult to manipulate than the pollen of many other plant species (Bou Daher et al., 2009). With the publication of optimized methods for in vitro germination (Boavida and McCormick, 2007; Bou Daher et al., 2009), it has become much more feasible to germinate healthy-looking Arabidopsis pollen tubes in vitro in a highly reproducible way.The precisely controlled spatial distribution of biochemical components in the pollen tube cell wall is crucial for shape generation and maintenance of this perfectly cylindrical cell (Geitmann and Parre, 2004; Aouar et al., 2010; Fayant et al., 2010; Geitmann, 2010). The pollen tube, therefore, represents an ideal model system to study the link between intracellular signaling, biochemistry, cell mechanical properties, and morphogenesis in plant cells. Because of its typically fast growth rates, it responds quickly to any environmental triggers such as pharmacological, hormone, or enzymatic treatments. Adding Arabidopsis to the group of commonly studied pollen tube species is particularly timely, because one-third of the approximately 800 cell wall synthesis genes identified in this species are expressed in or are specific to its pollen (Pina et al., 2005). Therefore, the Arabidopsis pollen tube has become a valuable system for cell wall studies, especially with the increasing availability of cell wall mutant lines (Liepman et al., 2010).Here we describe the biochemical composition of the Arabidopsis pollen tube cell wall grown in in vitro conditions using immunocytochemical labeling coupled with epifluorescence and electron microscopic techniques. Rather than relying on imaging alone, we developed a quantitative strategy to assess the precise spatial distribution of cell wall components. This quantitative approach will provide an important tool and baseline dataset for the investigation of mutant phenotypes and for the interpretation of pharmacological studies. Furthermore, we used selective and strategically combined enzymatic digestions to determine the degree of connectivity between the individual types of cell wall polysaccharide networks.  相似文献   

12.
13.
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   

14.
15.
16.
A collection of 130 new plant cell wall glycan-directed monoclonal antibodies (mAbs) was generated with the aim of facilitating in-depth analysis of cell wall glycans. An enzyme-linked immunosorbent assay-based screen against a diverse panel of 54 plant polysaccharides was used to characterize the binding patterns of these new mAbs, together with 50 other previously generated mAbs, against plant cell wall glycans. Hierarchical clustering analysis was used to group these mAbs based on the polysaccharide recognition patterns observed. The mAb groupings in the resulting cladogram were further verified by immunolocalization studies in Arabidopsis (Arabidopsis thaliana) stems. The mAbs could be resolved into 19 clades of antibodies that recognize distinct epitopes present on all major classes of plant cell wall glycans, including arabinogalactans (both protein- and polysaccharide-linked), pectins (homogalacturonan, rhamnogalacturonan I), xyloglucans, xylans, mannans, and glucans. In most cases, multiple subclades of antibodies were observed to bind to each glycan class, suggesting that the mAbs in these subgroups recognize distinct epitopes present on the cell wall glycans. The epitopes recognized by many of the mAbs in the toolkit, particularly those recognizing arabinose- and/or galactose-containing structures, are present on more than one glycan class, consistent with the known structural diversity and complexity of plant cell wall glycans. Thus, these cell wall glycan-directed mAbs should be viewed and utilized as epitope-specific, rather than polymer-specific, probes. The current world-wide toolkit of approximately 180 glycan-directed antibodies from various laboratories provides a large and diverse set of probes for studies of plant cell wall structure, function, dynamics, and biosynthesis.Cell walls play important roles in the structure, physiology, growth, and development of plants (Carpita and Gibeaut, 1993). Plant cell wall materials are also important sources of human and animal nutrition, natural textile fibers, paper and wood products, and raw materials for biofuel production (Somerville, 2007). Many genes thought to be responsible for plant wall biosynthesis and modification have been identified (Burton et al., 2005; Lerouxel et al., 2006; Mohnen et al., 2008), and 15% of the Arabidopsis (Arabidopsis thaliana) genome is likely devoted to these functions (Carpita et al., 2001). However, phenotypic analysis in plants carrying cell wall-related mutations has proven particularly difficult. First, cell wall-related genes are often expressed differentially and at low levels between cells of different tissues (Sarria et al., 2001). Also, plants have compensatory mechanisms to maintain wall function in the absence of a particular gene (Somerville et al., 2004). Thus, novel tools and approaches are needed to characterize wall structures and the genes responsible for their synthesis and modification.Monoclonal antibodies (mAbs) developed against cell wall polymers have emerged as an important tool for the study of plant cell wall structure and function (Knox, 2008). Previous studies have utilized mAbs that bind epitopes present on rhamnogalacturonan I (RG-I; Freshour et al., 1996; Jones et al., 1997; Willats et al., 1998; McCartney et al., 2000; Clausen et al., 2004; Altaner et al., 2007), homogalacturonan (Willats et al., 2001; Clausen et al., 2003), xylogalacturonan (Willats et al., 2004), xylans and arabinoxylans (McCartney et al., 2005), xyloglucan (Freshour et al., 1996, 2003; Marcus et al., 2008), arabinogalactan(protein) (Pennell et al., 1991; Puhlmann et al., 1994; Dolan et al., 1995; Smallwood et al., 1996), and extensins (Smallwood et al., 1995) to localize these epitopes in plant cells and tissues. In addition, mAbs have been used to characterize plants carrying mutations in genes thought to be associated with cell wall biosynthesis and metabolism (Orfila et al., 2001; Seifert, 2004; Persson et al., 2007; Cavalier et al., 2008; Zabotina et al., 2008). Despite their utility, the available set of mAbs against carbohydrate structures is relatively small given the structural complexity of wall polymers (Ridley et al., 2001; O''Neill and York, 2003), and knowledge of their epitope specificity is limited. Thus, additional mAbs specific to diverse epitope structures and methods for rapid epitope characterization are needed (Somerville et al., 2004).Here, we report the generation of 130 new mAbs that bind to diverse epitopes present on a broad spectrum of plant cell wall glycans. In addition, approximately 50 previously reported or generated mAbs were included in the ELISA-based screens used to group the antibodies according to their binding patterns against a diverse panel of 54 polysaccharides. The resulting ELISA data were analyzed by hierarchical clustering to illustrate the relationships between the available mAbs. Nineteen groups of mAbs were identified from the clustering analysis. Some initial information regarding possible epitopes recognized by some of these antibodies could be inferred from the clustering analysis.  相似文献   

17.
Mannans are hemicellulosic polysaccharides that are considered to have both structural and storage functions in the plant cell wall. However, it is not yet known how mannans function in Arabidopsis (Arabidopsis thaliana) seed mucilage. In this study, CELLULOSE SYNTHASE-LIKE A2 (CSLA2; At5g22740) expression was observed in several seed tissues, including the epidermal cells of developing seed coats. Disruption of CSLA2 resulted in thinner adherent mucilage halos, although the total amount of the adherent mucilage did not change compared with the wild type. This suggested that the adherent mucilage in the mutant was more compact compared with that of the wild type. In accordance with the role of CSLA2 in glucomannan synthesis, csla2-1 mucilage contained 30% less mannosyl and glucosyl content than did the wild type. No appreciable changes in the composition, structure, or macromolecular properties were observed for nonmannan polysaccharides in mutant mucilage. Biochemical analysis revealed that cellulose crystallinity was substantially reduced in csla2-1 mucilage; this was supported by the removal of most mucilage cellulose through treatment of csla2-1 seeds with endo-β-glucanase. Mutation in CSLA2 also resulted in altered spatial distribution of cellulose and an absence of birefringent cellulose microfibrils within the adherent mucilage. As with the observed changes in crystalline cellulose, the spatial distribution of pectin was also modified in csla2-1 mucilage. Taken together, our results demonstrate that glucomannans synthesized by CSLA2 are involved in modulating the structure of adherent mucilage, potentially through altering cellulose organization and crystallization.Mannan polysaccharides are a complex set of hemicellulosic cell wall polymers that are considered to have both structural and storage functions. Based on the particular chemical composition of the backbone and the side chains, mannan polysaccharides are classified into four types: pure mannan, glucomannan, galactomannan, and galactoglucomannan (Moreira and Filho, 2008; Wang et al., 2012; Pauly et al., 2013). Each of these polysaccharides is composed of a β-1,4-linked backbone containing Man or a combination of Glc and Man residues. In addition, the mannan backbone can be substituted with side chains of α-1,6-linked Gal residues. Mannan polysaccharides have been proposed to cross link with cellulose and other hemicelluloses via hydrogen bonds (Fry, 1986; Iiyama et al., 1994; Obel et al., 2007; Scheller and Ulvskov, 2010). Furthermore, it has been reported that heteromannans with different levels of substitution can interact with cellulose in diverse ways (Whitney et al., 1998). Together, these observations indicate the complexity of mannan polysaccharides in the context of cell wall architecture.CELLULOSE SYNTHASE-LIKE A (CSLA) enzymes have been shown to have mannan synthase activity in vitro. These enzymes polymerize the β-1,4-linked backbone of mannans or glucomannans, depending on the substrates (GDP-Man and/or GDP-Glc) provided (Richmond and Somerville, 2000; Liepman et al., 2005, 2007; Pauly et al., 2013). In Arabidopsis (Arabidopsis thaliana), nine CSLA genes have been identified; different CSLAs are responsible for the synthesis of different mannan types (Liepman et al., 2005, 2007). CSLA7 has mannan synthase activity in vitro (Liepman et al., 2005) and has been shown to synthesize stem glucomannan in vivo (Goubet et al., 2009). Disrupting the CSLA7 gene results in defective pollen growth and embryo lethality phenotypes in Arabidopsis, indicating structural or signaling functions of mannan polysaccharides during plant embryo development (Goubet et al., 2003). A mutation in CSLA9 results in the inhibition of Agrobacterium tumefaciens-mediated root transformation in the rat4 mutant (Zhu et al., 2003). CSLA2, CSLA3, and CSLA9 are proposed to play nonredundant roles in the biosynthesis of stem glucomannans, although mutations in CSLA2, CSLA3, or CSLA9 have no effect on stem development or strength (Goubet et al., 2009). All of the Arabidopsis CSLA proteins have been shown to be involved in the biosynthesis of mannan polysaccharides in the plant cell wall (Liepman et al., 2005, 2007), although the precise physiological functions of only CSLA7 and CSLA9 have been conclusively demonstrated.In Arabidopsis, when mature dry seeds are hydrated, gel-like mucilage is extruded to envelop the entire seed. Ruthenium red staining of Arabidopsis seeds reveals two different mucilage layers, termed the nonadherent and the adherent mucilage layers (Western et al., 2000; Macquet et al., 2007a). The outer, nonadherent mucilage is loosely attached and can be easily extracted by shaking seeds in water. Compositional and linkage analyses suggest that this layer is almost exclusively composed of unbranched rhamnogalacturonan I (RG-I) (>80% to 90%), with small amounts of branched RG-I, arabinoxylan, and high methylesterified homogalacturonan (HG). By contrast, the inner, adherent mucilage layer is tightly attached to the seed and can only be removed by strong acid or base treatment, or by enzymatic digestion (Macquet et al., 2007a; Huang et al., 2011; Walker et al., 2011). As with the nonadherent layer, adherent mucilage is also mainly composed of unbranched RG-I, but with small numbers of arabinan and galactan ramifications (Penfield et al., 2001; Willats et al., 2001; Dean et al., 2007; Macquet et al., 2007a, 2007b; Arsovski et al., 2009; Haughn and Western, 2012). There are also minor amounts of pectic HG in the adherent mucilage, with high methylesterified HG in the external domain compared with the internal domain of the adherent layer (Willats et al., 2001; Macquet et al., 2007a; Rautengarten et al., 2008; Sullivan et al., 2011; Saez-Aguayo et al., 2013). In addition, the adherent mucilage contains cellulose (Blake et al., 2006; Macquet et al., 2007a), which is entangled with RG-I and is thought to anchor the pectin-rich mucilage onto seeds (Macquet et al., 2007a; Harpaz-Saad et al., 2011, 2012; Mendu et al., 2011; Sullivan et al., 2011). As such, Arabidopsis seed mucilage is considered to be a useful model for investigating the biosynthesis of cell wall polysaccharides and how this process is regulated in vivo (Haughn and Western, 2012).Screening for altered seed coat mucilage has led to the identification of several genes encoding enzymes that are involved in the biosynthesis or modification of mucilage components. RHAMNOSE SYNTHASE2/MUCILAGE-MODIFIED4 (MUM4) is responsible for the synthesis of UDP-l-Rha (Usadel et al., 2004; Western et al., 2004; Oka et al., 2007). The putative GALACTURONSYLTRANSFERASE11 can potentially synthesize mucilage RG-I or HG pectin from UDP-d-GalUA (Caffall et al., 2009). GALACTURONSYLTRANSFERASE-LIKE5 appears to function in the regulation of the final size of the mucilage RG-I (Kong et al., 2011, 2013). Mutant seeds defective in these genes display reduced thickness of the extruded mucilage layer compared with wild-type Arabidopsis seeds.RG-I deposited in the apoplast of seed coat epidermal cells appears to be synthesized in a branched form that is subsequently modified by enzymes in the apoplast. MUM2 encodes a β-galactosidase that removes Gal residues from RG-I side chains (Dean et al., 2007; Macquet et al., 2007b). β-XYLOSIDASE1 encodes an α-l-arabinfuranosidase that removes Ara residues from RG-I side chains (Arsovski et al., 2009). Disruptions of these genes lead to defective hydration properties and affect the extrusion of mucilage. Furthermore, correct methylesterification of mucilage HG is also required for mucilage extrusion. HG is secreted into the wall in a high methylesterified form that can then be enzymatically demethylesterified by pectin methylesterases (PMEs; Bosch and Hepler, 2005). PECTIN METHYLESTERASE INHIBITOR6 (PMEI6) inhibits PME activities (Saez-Aguayo et al., 2013). The subtilisin-like Ser protease (SBT1.7) can activate other PME inhibitors, but not PMEI6 (Rautengarten et al., 2008; Saez-Aguayo et al., 2013). Disruption of either PMEI6 or SBT1.7 results in the delay of mucilage release.Although cellulose is present at low levels in adherent mucilage, it plays an important adhesive role for the attachment of mucilage pectin to the seed coat epidermal cells. The orientation and amount of pectin associated with the cellulose network is largely determined by cellulose conformation properties (Macquet et al., 2007a; Haughn and Western, 2012). Previous studies have demonstrated that CELLULOSE SYNTHASE A5 (CESA5) is required for the production of seed mucilage cellulose and the adherent mucilage in the cesa5 mutant can be easily extracted with water (Harpaz-Saad et al., 2011, 2012; Mendu et al., 2011; Sullivan et al., 2011).Despite all of these discoveries, large gaps remain in the current knowledge of the biosynthesis and functions of mucilage polysaccharides in seed coats. In this study, we show that CSLA2 is involved in the biosynthesis of mucilage glucomannan. Furthermore, we show that CSLA2 functions in the maintenance of the normal structure of the adherent mucilage layer through modifying the mucilage cellulose ultrastructure.  相似文献   

18.
Plants invest a lot of their resources into the production of an extracellular matrix built of polysaccharides. While the composition of the cell wall is relatively well characterized, the functions of the individual polymers and the enzymes that catalyze their biosynthesis remain poorly understood. We exploited the Arabidopsis (Arabidopsis thaliana) seed coat epidermis (SCE) to study cell wall synthesis. SCE cells produce mucilage, a specialized secondary wall that is rich in pectin, at a precise stage of development. A coexpression search for MUCILAGE-RELATED (MUCI) genes identified MUCI10 as a key determinant of mucilage properties. MUCI10 is closely related to a fenugreek (Trigonella foenumgraecum) enzyme that has in vitro galactomannan α-1,6-galactosyltransferase activity. Our detailed analysis of the muci10 mutants demonstrates that mucilage contains highly branched galactoglucomannan (GGM) rather than unbranched glucomannan. MUCI10 likely decorates glucomannan, synthesized by CELLULOSE SYNTHASE-LIKE A2, with galactose residues in vivo. The degree of galactosylation is essential for the synthesis of the GGM backbone, the structure of cellulose, mucilage density, as well as the adherence of pectin. We propose that GGM scaffolds control mucilage architecture along with cellulosic rays and show that Arabidopsis SCE cells represent an excellent model in which to study the synthesis and function of GGM. Arabidopsis natural varieties with defects similar to muci10 mutants may reveal additional genes involved in GGM synthesis. Since GGM is the most abundant hemicellulose in the secondary walls of gymnosperms, understanding its biosynthesis may facilitate improvements in the production of valuable commodities from softwoods.The plant cell wall is the key determinant of plant growth (Cosgrove, 2005) and represents the most abundant source of biopolymers on the planet (Pauly and Keegstra, 2010). Consequently, plants invest a lot of their resources into the production of this extracellular structure. Thus, it is not surprising that approximately 15% of Arabidopsis (Arabidopsis thaliana) genes are likely dedicated to the biosynthesis and modification of cell wall polymers (Carpita et al., 2001). Plant walls consist mainly of polysaccharides (cellulose, hemicellulose, and pectin) but also contain lignin and glycoproteins. While the biochemical structure of each wall component has been relatively well characterized, the molecular players involved in their biogenesis remain poorly understood (Keegstra, 2010). The functions of the individual polymers, and how they are assembled into a three-dimensional matrix, are also largely unknown (Burton et al., 2010; Burton and Fincher, 2012).Significant breakthroughs in cell wall research have been achieved through the examination of specialized plant tissues that contain elevated levels of a single polysaccharide (Pauly and Keegstra, 2010). Some species, particularly legumes, accumulate large amounts of the hemicellulose galactomannan during secondary wall thickening of the seed (Srivastava and Kapoor, 2005). Analysis of the developing fenugreek (Trigonella foenumgraecum) endosperm led to the purification of a GALACTOMANNAN GALACTOSYLTRANSFERASE (TfGMGT), the first glycosyltransferase (GT) whose activity in plant cell wall synthesis was demonstrated in vitro (Scheller and Ulvskov, 2010). TfGMGT catalyzes the decoration of mannan chains with single α-1,6-galactosyl residues (Edwards et al., 1999). A similar approach in guar (Cyamopsis tetragonoloba) seeds revealed that the β-1,4-linked mannan backbone is synthesized by a member of the CELLULOSE SYNTHASE-LIKE A (CSLA) protein family (Dhugga et al., 2004).Galactomannan functions as a storage polymer in the endosperm of the aforementioned seeds, analogous to starch in cereal grains (Dhugga et al., 2004), but it also has important rheological properties in the cell wall that have been exploited to produce valuable stabilizers and gelling agents for human consumption (Srivastava and Kapoor, 2005). The Man-to-Gal ratio is essential for the application of galactomannan gums in the food industry (Edwards et al., 1992). This is because unsubstituted mannan chains can interact via hydrogen bonds to produce crystalline microfibrils similar to cellulose (Millane and Hendrixson, 1994). Indeed, some algae that lack cellulose employ mannan fibrils as a structural material (Preston, 1968). The addition of Gal branches to the smooth, ribbon-like mannan chains creates hairy regions that limit self-association and promote gelation (Dea et al., 1977). All mannans are likely synthesized as highly substituted polymers that are trimmed in the cell wall (Scheller and Ulvskov, 2010).Generally, polysaccharides containing backbones of β-1,4-linked Man units can be classified as heteromannan (HM). Galactoglucomannan (GGM) is the main hemicellulose in gymnosperm secondary walls and, in contrast to galactomannan, has a backbone that contains both Glc and Man units (Pauly et al., 2013). HM is detected in most Arabidopsis cell types (Handford et al., 2003) and facilitates embryogenesis (Goubet et al., 2009), germination (Rodríguez-Gacio et al., 2012), tip growth (Bernal et al., 2008), and vascular development (Benová-Kákosová et al., 2006; Yin et al., 2011). In the last 10 years, in vitro mannan synthase activity has been demonstrated for recombinant CSLA proteins from many land plants (Liepman et al., 2005, 2007; Suzuki et al., 2006; Gille et al., 2011; Wang et al., 2012). HM synthesis may also involve CELLULOSE SYNTHASE-LIKE D (CSLD) enzymes and MANNAN SYNTHESIS-RELATED (MSR) accessory proteins (Yin et al., 2011; Wang et al., 2013), but their precise roles in relation to the CSLAs have not been established. Arabidopsis CSLA2, like most other isoforms, can use both GDP-Man and GDP-Glc as substrates in vitro (Liepman et al., 2005, 2007) and is responsible for stem glucomannan synthesis in vivo along with CSLA3 and CSLA7 (Goubet et al., 2009). CSLA2 also participates in the synthesis of glucomannan present in mucilage produced by seed coat epidermal (SCE) cells (Yu et al., 2014).Arabidopsis SCE cells represent an excellent genetic model in which to study the synthesis, polar secretion, and modification of polysaccharides, since these processes dominate a precise stage of seed coat development but are not essential for seed viability in laboratory conditions (Haughn and Western, 2012; North et al., 2014; Voiniciuc et al., 2015). Hydration of mature seeds in water releases a large gelatinous capsule, rich in the pectic polymer rhamnogalacturonan I, which can be easily stained or extracted (Macquet et al., 2007). Biochemical and cytological experiments indicate that Arabidopsis seed mucilage is more than just pectin and, in addition to cellulose, is likely to contain glycoproteins and at least two hemicellulosic polymers (Voiniciuc et al., 2015). There is mounting evidence that, despite their low abundance, these components play critical functions in seed mucilage architecture. The structure of homogalacturonan (HG), the major pectin in primary cell walls but a minor mucilage component, appears to be a key determinant of gelling properties and mucilage extrusion (Rautengarten et al., 2008; Saez-Aguayo et al., 2013; Voiniciuc et al., 2013). Mucilage attachment to seeds is maintained by the SALT OVERLY SENSITIVE5 glycoprotein and cellulose synthesized by multiple CELLULOSE SYNTHASE (CESA) isoforms (Harpaz-Saad et al., 2011; Mendu et al., 2011; Sullivan et al., 2011; Griffiths et al., 2014, 2015). From more than 35 genes that are reported to affect Arabidopsis seed mucilage properties (Voiniciuc et al., 2015), only CSLA2, CESA3, CESA5, GALACTURONOSYLTRANSFERASE11 (GAUT11; Caffall et al., 2009), and GAUT-LIKE5 (GATL5; Kong et al., 2013) are predicted to encode GTs. This highlights that, despite many detailed studies about mucilage production in SCE cells, the synthesis of its components remains poorly understood.To address this issue, we conducted a reverse genetic search for MUCILAGE-RELATED (MUCI) genes that may be required for polysaccharide biosynthesis. One of these, MUCI10, encodes a member of the Carbohydrate Active Enzymes family, GT34 (Lombard et al., 2014), which includes at least two enzymatic activities and seven Arabidopsis proteins (Keegstra and Cavalier, 2010). Five of them function as XYLOGLUCAN XYLOSYLTRANSFERASES (XXT1–XXT5) in vivo and/or in vitro (Faik et al., 2002; Cavalier et al., 2008; Vuttipongchaikij et al., 2012). MUCI10/GT7 (At2g22900) and its paralog GT6 (At4g37690) do not function as XXTs (Vuttipongchaikij et al., 2012) and are more closely related to the TfGMGT enzyme (Faik et al., 2002; Keegstra and Cavalier, 2010). MUCI10, also called GALACTOSYLTRANSFERASE-LIKE6 (GTL6), served as a Golgi marker in multiple proteomic studies of Arabidopsis callus cultures (Dunkley et al., 2004, 2006; Nikolovski et al., 2012, 2014). Nevertheless, the role of TfGMGT orthologs in Arabidopsis remained unknown. We show that MUCI10 is responsible for the extensive galactosylation of glucomannan in mucilage and influences glucomannan backbone synthesis, cellulose structure, and the distribution of pectin.  相似文献   

19.
A 5.5-y-old intact male cynomolgus macaque (Macaca fasicularis) presented with inappetence and weight loss 57 d after heterotopic heart and thymus transplantation while receiving an immunosuppressant regimen consisting of tacrolimus, mycophenolate mofetil, and methylprednisolone to prevent graft rejection. A serum chemistry panel, a glycated hemoglobin test, and urinalysis performed at presentation revealed elevated blood glucose and glycated hemoglobin (HbA1c) levels (727 mg/dL and 10.1%, respectively), glucosuria, and ketonuria. Diabetes mellitus was diagnosed, and insulin therapy was initiated immediately. The macaque was weaned off the immunosuppressive therapy as his clinical condition improved and stabilized. Approximately 74 d after discontinuation of the immunosuppressants, the blood glucose normalized, and the insulin therapy was stopped. The animal''s blood glucose and HbA1c values have remained within normal limits since this time. We suspect that our macaque experienced new-onset diabetes mellitus after transplantation, a condition that is commonly observed in human transplant patients but not well described in NHP. To our knowledge, this report represents the first documented case of new-onset diabetes mellitus after transplantation in a cynomolgus macaque.Abbreviations: NODAT, new-onset diabetes mellitus after transplantationNew-onset diabetes mellitus after transplantation (NODAT, formerly known as posttransplantation diabetes mellitus) is an important consequence of solid-organ transplantation in humans.7-10,15,17,19,21,25-28,31,33,34,37,38,42 A variety of risk factors have been identified including increased age, sex (male prevalence), elevated pretransplant fasting plasma glucose levels, and immunosuppressive therapy.7-10,15,17,19,21,25-28,31,33,34,37,38,42 The relationship between calcineurin inhibitors, such as tacrolimus and cyclosporin, and the development of NODAT is widely recognized in human medicine.7-10,15,17,19,21,25-28,31,33,34,37,38,42 Cynomolgus macaques (Macaca fasicularis) are a commonly used NHP model in organ transplantation research. Cases of natural and induced diabetes of cynomolgus monkeys have been described in the literature;14,43,45 however, NODAT in a macaque model of solid-organ transplantation has not been reported previously to our knowledge.  相似文献   

20.
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   

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