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.
The applicability of a synthetic zeolite (type 4A, Union Carbide Corp., Linde Div., New York, N.Y.) as a disruptive agent in a procedure for the preparation of pure bacterial cell wall fractions from a variety of phenol-killed gram-negative, gram-positive, and acid-fast bacteria was demonstrated. The disruptive effect was found to be limited with formaldehyde-killed gram-positive cells and most gram-positive cocci killed either by phenol or formaldehyde.  相似文献   

2.
《Experimental mycology》1994,18(2):159-167
Iranzo, M., Marcilla, A., Elorza, M. V., Mormeneo, S., and Sentandreu, R. 1994. Preparation of anti-protein and anti-mannan antisera against fungal cell wall by affinity chromatography. Experimental Mycology 18, 159-167. A novel and easy chromatographic method has been developed for the isolation of anti-protein and anti-mannan antisera from a population of polyclonal antibodies obtained against Candida albicans and Yarrowia lipolytica cell wall mannoproteins. The technique is based on the immobilization of mannan (to be used as immunoadsorbent) by Affi-Prep Hz resin after the oxidation of neighboring hydroxyl groups of the polysaccharide with sodium periodate. For Y. lipolytica polyclonal antiserum, a single chromatographic step using the homologous mannan was sufficient to obtain an antiprotein antibody preparation free of antimannan antibodies. For C. albicans, three chromatographic processes using homologous and heterologous mannan were needed to obtain a satisfactory antiprotein antiserum. The potential application of the anti-protein antiserum obtained has been demonstrated by indirect immunofluorescence assays of whole cells and electrophoretic analysis of wall proteins in C. albicans and Saccharomyces cerevisiae .  相似文献   

3.
The ecological and biogeochemical relevance of hydrolytic enzymes associated with the fungal cell wall has been poorly studied in ectomycorrhizal (ECM) fungi. We used a modified sequential extraction procedure to investigate the activity of various hydrolytic enzymes (β-glucosidase, acid-phosphatase, leucine-aminopeptidase, chitinase, xylanase and glucuronidase) and their association with the cell wall of three ECM fungi (Rhizopogon roseolus, Paxillus involutus and Piloderma croceum). Fungi were grown on C-rich solid medium under three different P concentrations (3.7, 0.37 and 0.037 mM). The sequential extraction procedure classifies enzymes as: (a) cytosolic, (b) loosely bound, (c) hydrophobically bound, (d) ionically bound and (e) covalently bound. Results showed that for the same fungus absolute enzymatic activity was affected by P concentration, whilst enzymatic compartmentalization among the cytosol and the cell wall fractions was not. The association of enzymes with the cell wall was fungus- and enzyme-specific. Our data indicate also that enzymes best known for being either extracellular or cytosolic or both, do act in muro as well. The ecological implications of cell wall-bound enzymes and the potential applications and limitations of sequential extractions are further discussed.  相似文献   

4.
The occurrence of filamentous, branching, catalase-negative bacteria as a numerically predominant microflora of various soils was demonstrated by using a dilution frequency isolation procedure. The major characteristics of these organisms were those of the order Actinomycetales. However, they could not be placed in any of the present genera of this order and, therefore, a new genus, Agromyces, was proposed for these organisms. This genus includes catalase-negative, nutritionally-fastidious microorganisms whose cells produce a true branching mycelium that fragments into coccoid and diphtheroid forms. Also, they have an oxidative metabolism, are microaerophilic to aerobic, and contain neither diaminopimelic acid nor lysine as major constituents of the cell wall glycopeptide. The type species would be Agromyces ramosus, gen. n., sp. n. The possible importance of these organisms in clarifying certain phylogenetic relationships of the Actinomycetales is discussed.  相似文献   

5.
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   

6.
Chemical composition of variants of aerobic actinomycetes   总被引:3,自引:0,他引:3       下载免费PDF全文
It has been shown previously that aerobic actinomycetes can be separated into four main groups on the basis of their cell wall composition. Six representatives of aerobic actinomycetes (Nocardia asteroides and Micropolyspora brevicatena, cell wall type IV; N. madurae, Microbispora rosea, cell wall type III; Actinoplanes sp., cell wall type II; Streptomyces griseus, cell wall type I) were subjected to selecting agents which permitted the isolation of stable variants morphologically different from the parent strain. Whole cell analyses of 134 substrains from the six parents revealed no significant change in the isomeric form of diaminopimelic acid or in sugar constituents. Analyses of cell wall preparations from 52 of these did not reveal any change in the diagnostic constituents of their murein or polysaccharides.  相似文献   

7.
The Reduced Wall Acetylation (RWA) proteins are involved in cell wall acetylation in plants. Previously, we described a single mutant, rwa2, which has about 20% lower level of O-acetylation in leaf cell walls and no obvious growth or developmental phenotype. In this study, we generated double, triple, and quadruple loss-of-function mutants of all four members of the RWA family in Arabidopsis (Arabidopsis thaliana). In contrast to rwa2, the triple and quadruple rwa mutants display severe growth phenotypes revealing the importance of wall acetylation for plant growth and development. The quadruple rwa mutant can be completely complemented with the RWA2 protein expressed under 35S promoter, indicating the functional redundancy of the RWA proteins. Nevertheless, the degree of acetylation of xylan, (gluco)mannan, and xyloglucan as well as overall cell wall acetylation is affected differently in different combinations of triple mutants, suggesting their diversity in substrate preference. The overall degree of wall acetylation in the rwa quadruple mutant was reduced by 63% compared with the wild type, and histochemical analysis of the rwa quadruple mutant stem indicates defects in cell differentiation of cell types with secondary cell walls.Plant cell walls are multifunctional viscoelastic networks mainly composed of polysaccharides. Many of these polysaccharides, including xylans, (gluco)mannans, xyloglucans (XyGs), and pectins, have various degrees and patterns of acetyl esterification (Gille and Pauly, 2012; Pawar et al., 2013). The biological role of cell wall acetylation is not well understood, but it is believed to be important for pathogen resistance and plant development, and the acetylation of pectin also impacts upon the mechanical properties of cell walls (Manabe et al., 2011; Orfila et al., 2012; Pogorelko et al., 2013). In vitro, acetyl groups influence susceptibility to enzymatic degradation of pectin and xylan (Selig et al., 2009; Chen et al., 2012; Gou et al., 2012; Orfila et al., 2012; Pogorelko et al., 2013), and therefore acetylation may constitute a barrier to cell wall deconstruction. Alkali treatment of wall materials, which hydrolyzes the ester bonds, is broadly used to make polysaccharides more extractable. The treatment does not only facilitate the degradation of xylan and pectins, but also improves the deconstruction of cellulose, as the depolymerization of noncellulosic polymers results in a better accessibility to cellulose by degrading enzymes (Selig et al., 2009). Low levels of acetylated polysaccharides in plant feedstocks would be desirable for downstream processing in biorefineries, firstly, because the cell wall material of plant feedstocks with low level of acetylation is expected to be more easily extracted and, secondly, because less acetate, which is highly toxic to microorganisms such as yeast (Saccharomyces cerevisiae), would be released during extraction (Manabe et al., 2011; Gille and Pauly, 2012; Pawar et al., 2013). However, although reducing the O-acetylation level of xylan by approximately 60%, as observed in the walls of the Arabidopsis (Arabidopsis thaliana) eskimo1 mutant, enhances enzymatic degradation of isolated xylan (Yuan et al., 2013), enzymatic hydrolysis yields of whole wall materials have been reported to actually be decreased (Xiong et al., 2013). This presumably results from a tighter association between these now lowly substituted xylan polymers and cellulose (Xiong et al., 2013).Recently, we reported REDUCED WALL ACETYLATION2 (RWA2), the first protein to be involved in cell wall acetylation in planta (Manabe et al., 2011). RWA2 is a member of a small family consisting of four proteins in Arabidopsis, and its loss-of-function mutants display 20% reduction of acetylation in a range of polysaccharides that include XyG and pectins. We have hypothesized, based on phylogenetic analysis, expression pattern, moderate reduction in acetylation, and the absence of morphological phenotype, that RWA proteins have redundant functions in a biochemical reaction that occurs prior to the actual acetylation of specific polysaccharides. Independently to our research, a quadruple mutant of RWA has been reported to display reduction in xylan acetylation, secondary cell wall thickness, and mechanical strength of the stem (Lee et al., 2011). Meanwhile, Gille et al. (2011) have discovered a new family of proteins involved in the acetylation of specific polysaccharides: the plant-specific DOMAIN OF UNKNOWN FUNCTION (DUF) 231 family (also known as TRICHOME BIREFRINGENCE-LIKE [TBL] family). The loss-of-function mutants altered xyloglucan4 (axy4)/tbl27 and axy4L/tbl22 lack O-acetylation specifically of XyG in certain tissues, while eskimo1/tbl29 mutants contain reduced O-acetylation of xylan (Xiong et al., 2013; Yuan et al., 2013). The TBL/DUF231 family proteins and the RWA proteins have sequence similarity to the N-terminal and C-terminal regions of the fungal protein Cas1p, respectively (Anantharaman and Aravind, 2010). This could suggest that the TBL and RWA proteins function in protein complexes where the determinants of substrate specificity reside in the TBL partner (Manabe et al., 2011). However, because there are many more TBL proteins than RWA proteins (e.g. 46 TBL proteins versus four RWA proteins in the genome of Arabidopsis), it is likely that they do not form discrete and invariable complexes. Crossing of rwa2-3 and a leaky allele of axy4, axy4-1, resulted in a double mutant with partially additive phenotype (Gille et al., 2011). Its XyG acetylation is lower compared with either single mutant. From this analysis, RWA2 and AXY4 have been hypothesized to work in synergy, although the function of RWA2 might be substituted by other RWAs (Gille et al., 2011). Here, we have generated all the combinations of double, triple, and quadruple mutants of all four members of RWA family to further investigate the functional diversity and redundancy and to explore the function of cell wall acetylation and the role of RWAs in the network of acetylation-related enzymes. The triple and quadruple mutants we have obtained displayed severe and distinct phenotypes such as extreme dwarfism. This contrasts with the very mild phenotypes reported by Lee et al. (2011). Taken together, RWAs have partially redundant functions in the process of cell wall acetylation and show distinct impacts upon different cell wall polysaccharides.  相似文献   

8.
A new HPLC method was developed to separate linear from β(1–6)-branched β(1–3)-glucooligosaccharides. This methodology has permitted the isolation of the first fungal β(1–6)/β(1–3)-glucan branching transglycosidase using a cell wall autolysate of Aspergillus fumigatus (Af). The encoding gene, AfBGT2 is an ortholog of AfBGT1, another transglycosidase of A. fumigatus previously analyzed (Mouyna, I., Hartland, R. P., Fontaine, T., Diaquin, M., Simenel, C., Delepierre, M., Henrissat, B., and Latgé, J. P. (1998) Microbiology 144, 3171–3180). Both enzymes release laminaribiose from the reducing end of a β(1–3)-linked oligosaccharide and transfer the remaining chain to another molecule of the original substrate. The AfBgt1p transfer occurs at C-6 of the non-reducing end group of the acceptor, creating a kinked β(1–3;1–6) linear molecule. The AfBgt2p transfer takes place at the C-6 of an internal group of the acceptor, resulting in a β(1–3)-linked product with a β(1–6)-linked side branch. The single Afbgt2 mutant and the double Afbgt1/Afbgt2 mutant in A. fumigatus did not display any cell wall phenotype showing that these activities were not responsible for the construction of the branched β(1–3)-glucans of the cell wall.  相似文献   

9.
《Experimental mycology》1995,19(3):178-185
Molloy, C., Shepherd, M. G., and Sullivan, P. A. 1995. Differential extraction of N-acetylglucosaminidase and trehalase from the cell envelope of Candida albicans. Experimental Mycology 19, 178-185. Dithiothreitol (DTT) extraction of N-acetylglucosaminidase and trehalase from intact Candida albicans ATCC 10261 cells was monitored as an index of cell envelope porosity during N-acetylglucosamine-induced morphogenesis. Trehalase, which is secreted into the cell envelope during starvation and bud-formation, displayed similar extraction kinetics in starved, germ tube-forming, and bud-forming cells, indicating that the mother cell wall remains largely unchanged during morphogenic outgrowth and that the porosity of bud and mother cell walls is similar. N-acetylglucosaminidase, which is secreted specifically during morphogenesis, was released eightfold more rapidly from germ tube-forming than bud-forming cells, reflecting major differences in porosity between bud and germ tube. In addition, by assaying DTT extracts and extracted cell residues, it was found that the total extracellular N -acetylglucosaminidase activity increased 2- to 2.5-fold during DTT treatment. Thus, DTT unmasks a cryptic form of N-acetylglucosaminidase. The cryptic activity was associated with the cell wall fraction.  相似文献   

10.
The cell wall consists of cellulose microfibrils embedded within a matrix of hemicellulose and pectin. Cellulose microfibrils are synthesized at the plasma membrane, whereas matrix polysaccharides are synthesized in the Golgi apparatus and secreted. The trafficking of vesicles containing cell wall components is thought to depend on actin-myosin. Here, we implicate microtubules in this process through studies of the kinesin-4 family member, Fragile Fiber1 (FRA1). In an fra1-5 knockout mutant, the expansion rate of the inflorescence stem is halved compared with the wild type along with the thickness of both primary and secondary cell walls. Nevertheless, cell walls in fra1-5 have an essentially unaltered composition and ultrastructure. A functional triple green fluorescent protein-tagged FRA1 fusion protein moves processively along cortical microtubules, and its abundance and motile density correlate with growth rate. Motility of FRA1 and cellulose synthase complexes is independent, indicating that FRA1 is not directly involved in cellulose biosynthesis; however, the secretion rate of fucose-alkyne-labeled pectin is greatly decreased in fra1-5, and the mutant has Golgi bodies with fewer cisternae and enlarged vesicles. Based on our results, we propose that FRA1 contributes to cell wall production by transporting Golgi-derived vesicles along cortical microtubules for secretion.The cell wall plays a vital role in the life of a plant. In growing cells, the tough but extensible primary wall determines the rate and direction of expansion and overall plant form. In differentiated cells, such as interfascicular fibers and xylem cells, the thick secondary cell wall provides strength to withstand gravity and large negative pressures. Other than mechanics, cell walls feature in essential processes, such as pathogen resistance, signal transduction, and cell-to-cell communication. In addition, cell wall biomass has potential as a feedstock for biofuel production. Therefore, understanding cell wall biogenesis is important fundamentally and practically.Cell walls consist primarily of cellulose, hemicellulose, and pectin along with small amounts of protein. Typical of eukaryotic secretory products, hemicellulose and pectin are synthesized in the Golgi and then delivered to the extracellular space through the secretory system. Atypically, cellulose microfibrils are synthesized de novo at the plasma membrane by cellulose synthase (CESA) complexes, although the CESA complexes themselves are thought to be assembled in the Golgi and trafficked to the plasma membrane (McFarlane et al., 2014). How the various cell wall components are delivered to the plasma membrane and extracellular space to form a functional cell wall remains poorly understood.Among cell wall components, cellulose has perhaps the best understood delivery process, which is influenced by cortical microtubules (Baskin, 2001; Lloyd, 2011). The cortical microtubule array organizes cellulose deposition spatially by targeting the secretion of CESA complexes (Crowell et al., 2009; Gutierrez et al., 2009) and orienting their movement (Gardiner et al., 2003; Paredez et al., 2006). What controls the delivery of other wall components is less clear. Sustained transport of organelles in plants is actin based (Sparkes, 2011), and vesicle trafficking is generally assumed to be independent of microtubules, at least during interphase. Nevertheless, in xylem tracheary cells, cortical microtubule bands have been linked to not only cellulose guidance but also, the targeted exocytosis of hemicellulose and other matrix components (Fukuda, 1997). In seed coat cells, vesicles containing pectin associate with cortical microtubules that line the mucilage secretion pockets (McFarlane et al., 2008). In addition, in maize (Zea mays) roots, vesicles bind cortical microtubules densely (Tian et al., 2004). These and other observations indicate that cortical microtubules might serve as roadways for trafficking secretory vesicles.If microtubules are tracks, then the engines are kinesins, because seed plants lack dyneins (Zhu and Dixit, 2012). Kinesins are molecular motors that move along microtubules and transport various cargo, including organelles, vesicles, and proteins. Kinesins have proliferated in plant lineages, and many are expressed during interphase, which is surprising given that long-distance organelle motility is thought to be actin based. Recently, a particular plant kinesin of the kinesin-4 family, called Fragile Fiber1 (FRA1), was shown to move rapidly and processively (i.e. taking multiple steps) toward microtubule plus ends in vitro (Zhu and Dixit, 2011). This makes FRA1 a candidate for sustained and active vesicle transport.An Arabidopsis (Arabidopsis thaliana) partial loss-of-function mutant, fra1-1, was reported to have altered cellulose organization in fiber cells, despite having evidently undisturbed cortical microtubule organization (Zhong et al., 2002). Those results suggested a function for FRA1 in cell wall organization rather than secretion, and FRA1 has been proposed to link motile CESA complexes in the plasma membrane to cortical microtubules (Zhong et al., 2002; Lloyd and Chan, 2004; Zhu and Dixit, 2011). Strengthening this suggestion, a null mutant of the FRA1 ortholog in rice (Oryza sativa), brittle culm12 (bc12), also was reported to have disorganized sclerenchyma cell walls (Zhang et al., 2010). However, the motility of FRA1 in vivo and its relationship to CESA complexes remain unknown.We have reexamined FRA1 function in cell wall formation. Because the originally characterized allele, fra1-1, is predicted to give rise to a nearly full-length protein, we characterized a transfer DNA (T-DNA)-induced knockout mutant, fra1-5. Using this allele as well as imaging a functional FRA1-3GFP fusion protein, we show here that FRA1 is involved in membrane trafficking that contributes to delivery of cell wall polysaccharides, such as pectin. We propose that FRA1 drives the movement of vesicles containing cell wall cargo along cortical microtubules to facilitate their secretion.  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.
14.
The structure of the cell wall has a major impact on plant growth and development, and alteration of cell wall structural components is often detrimental to biomass production. However, the molecular mechanisms responsible for these negative effects are largely unknown. Arabidopsis (Arabidopsis thaliana) plants with altered pectin composition because of either the expression of the Aspergillus niger polygalacturonase II (AnPGII; 35S:AnPGII plants) or a mutation in the QUASIMODO2 (QUA2) gene that encodes a putative pectin methyltransferase (qua2-1 plants), display severe growth defects. Here, we show that expression of Arabidopsis PEROXIDASE71 (AtPRX71), encoding a class III peroxidase, strongly increases in 35S:AnPGII and qua2-1 plants as well as in response to treatments with the cellulose synthase inhibitor isoxaben, which also impairs cell wall integrity. Analysis of atprx71 loss-of-function mutants and plants overexpressing AtPRX71 indicates that this gene negatively influences Arabidopsis growth at different stages of development, likely limiting cell expansion. The atprx71-1 mutation partially suppresses the dwarf phenotype of qua2-1, suggesting that AtPRX71 contributes to the growth defects observed in plants undergoing cell wall damage. Furthermore, AtPRX71 seems to promote the production of reactive oxygen species in qua2-1 plants as well as plants treated with isoxaben. We propose that AtPRX71 contributes to strengthen cell walls, therefore restricting cell expansion, during normal growth and in response to cell wall damage.The cell wall is a complex, multifunctional and dynamic structure that provides mechanical support to plant cells, and it is involved in cell adhesion, defense against pathogens, regulation of metabolic functions, and cell-to-cell communication (Keegstra, 2010). Cell walls are usually composed of polysaccharides (cellulose, hemicelluloses, and pectins), phenolic compounds (e.g. ferulic acid and lignin), and proteins (Carpita and McCann, 2000). In the apoplast, cellulose microfibrils are associated to hemicelluloses, such as xyloglucans (XGs), producing a network embedded in a matrix of pectins. The latter are mainly composed of linear chains of homogalacturonan (HG) and branched chains of rhamnogalacturonans. Pectins are abundant in the middle lamella, where they ensure cell adhesion, as well as in primary and, to a lesser degree, secondary walls (Willats et al., 2001). HG is synthesized in a highly esterified form in the Golgi apparatus (Zhang and Staehelin, 1992) and then secreted in the apoplast, where pectin methylesterases remove part of the methyl groups (Pelloux et al., 2007). Free carboxylic groups allow the formation of so-called egg box structures, in which adjacent HG chains are linked by calcium-mediated ionic bridges, making the pectin matrix more rigid (Micheli, 2001). Pectins can also form other types of interactions, such as covalent cross links with other cell wall polysaccharides, phenolic compounds, and proteins (Caffall and Mohnen, 2009; Tan et al., 2013).The wall structure influences both the extent and the direction of cell expansion (Mirabet et al., 2011). Growth takes place perpendicularly to the direction of cellulose microfibrils, which are deposited along the perpendicular axis of the cell, providing resistance to turgor pressure and extensibility along the longitudinal axis (Crowell et al., 2010). Two major classes of proteins have been proposed to promote cell wall expansion: XG endotransglycosylases, which cleave XG chains and link together the newly generated reduced end to a new XG chain (Fry et al., 1992), and expansins, which promote primary cell walls relaxation by disrupting cellulose-hemicellulose noncovalent links (Cosgrove, 2000). During expansion, the cell wall is relaxed, whereas turgor forces induce its deformation; subsequently, expansin inhibition and formation of cross links between structural proteins (such as extensins), polysaccharides, and/or monolignols cause wall stiffening and, consequently, slow down expansion (Wolf et al., 2009).Regulation of apoplastic levels of reactive oxygen species (ROS) is important to determine cell expansion rate and organ size (Gapper and Dolan, 2006). ROS production in the cell wall is controlled, both under physiological conditions and in response to environmental stimuli, by several classes of enzymes, most prominently plasma membrane NADPH oxidases (Torres and Dangl, 2005) and class III peroxidases (CIII Prxs; Bolwell et al., 1999; Cosio and Dunand, 2009). NADPH oxidases, commonly known as respiratory burst oxidase homologs (Rbohs), are transmembrane proteins that oxidize cytoplasmic NADPH, translocate electrons across the plasma membrane, and generate superoxide radicals in the cell wall (Torres et al., 2002; Torres and Dangl, 2005). Superoxide radicals are then rapidly converted into hydrogen peroxide (H2O2) either spontaneously or in a reaction catalyzed by superoxide dismutases (Bolwell et al., 1999).CIII Prxs are heme-containing enzymes secreted in the extracellular space or the vacuole, where they perform two different enzymatic cycles, namely the peroxidative and hydroxylic cycles (Welinder et al., 2002; Passardi et al., 2004). During the peroxidative cycle, the enzyme uses H2O2 as an oxidant in a two-step reaction to convert different substrates, including cell wall phenolic compounds and structural proteins, into free radicals that can subsequently combine together to form covalent linkages. This activity contributes to cell wall stiffening and therefore limits growth. CIII Prxs can also cause cell wall loosening through the hydroxylic cycle, in which H2O2 and O2 are used in a Fenton-type reaction to generate hydroxyl radicals, including OH, that lead to nonenzymatic cleavage of polysaccharides (Chen and Schopfer, 1999; Dunand et al., 2003; Passardi et al., 2005). CIII Prxs can, therefore, play opposite roles in cell expansion, being able to cause both wall stiffening and loosening, depending on the growth conditions. CIII Prxs can also generate O2, which is then dismutated into H2O2, through the oxidation of NAD(P)H. These enzymes can both positively and negatively modulate apoplastic ROS levels (Passardi et al., 2004).The enzymatic characteristics of CIII Prxs allow them to participate in a wide range of physiological processes, including seed germination, plant growth, and elongation and defense against pathogens, as well as the catabolism of several extracellular molecules, including auxins (Hiraga et al., 2001). CIII Prxs are also involved in lignification through the H2O2-dependent generation of monolignol phenoxy radicals that spontaneously form lignin polymers (Marjamaa et al., 2009). Lignin monomers can also cross link cell wall polysaccharides, including pectins, through ferulate bridges or diferulate bonds formed by CIII Prxs in the presence of H2O2 (Iiyama et al., 1994). Notably, in some plant species, CIII Prxs-mediated cross links between ferulic acid and pectins arrest cell expansion (Iiyama et al., 1994). Lastly, CIII Prxs can cross link Tyr and Lys residues of extensins, contributing to the formation of a dense network within the cell wall (Schnabelrauch et al., 1996).CIII Prxs likely appeared when plants colonized land and subsequently underwent a high rate of gene duplication, with the consequent increase of functional specialization (Passardi et al., 2005). For instance, the Arabidopsis (Arabidopsis thaliana) genome encodes 73 putative CIII Prxs, with various spatiotemporal expression profiles, suggesting that different isoforms play specific roles in growth, development, and adaptation to the environment (Tognolli et al., 2002; Welinder et al., 2002). In this article, we show that the Arabidopsis CIII Prx gene AtPRX71, which is strongly expressed upon loss of cell wall integrity (CWI), negatively affects growth and cell size and positively regulates ROS levels. In addition, when AtPRX71 is lacking, both ROS accumulation and growth defects caused by cell wall alterations are significantly reduced. We propose that accumulation of ROS-generating CIII Prxs is a mechanism to cope with loss of CWI both during normal development and in response to stress, leading to wall stiffening and therefore limiting cell expansion and growth.  相似文献   

15.
16.
Host responses to Meloidodera floridensis Chitwood et al., 1956, M. charis Hooper, 1960, and M. belli Wouts, 1973 were examined on loblolly pine, peony, and sage, respectively, with light, scanning, and transmission electron microscopy. In each case the nematodes induce a single uninucleate giant cell. The giant cell is initiated in the pericycle and expands as it matures. The mature giant cell induced by M. floridensis is surrounded by vascular parenchyma, whereas that caused by M. charts and M. belli coutacts xylem and phloem. The cell wall of giant cells induced by all three Meloidodera spp. is generally thicker than that of surrounding cells, with the thickest part adjacent to the lip region of the nematode. The thinner portion of the wall includes numerous pit fields with plasmodesmata, but wall ingrowths were not detected in a thorough examination of the entire wall. The nucleus of a giant cell induced by M. goridensis is highly irregular in shape with deep invaginations, whereas those caused by M. charis and M. belli include a cluster of apparently interconnected nuclear units. Organelles, including mitochondria, endoplasmic reticulum, and plastids of giant cells caused by Meloidodera, are typical of those reported in host responses of other Heteroderidae. The formation of a single uninucleate giant cell by Meloidodera, Cryphodera, Hylonerna, and Sarisodera, but a syncytium by Atalodera and Heterodera sensu lato, might be considered in conjunction with additional characters to determine the most parsimonious pattern of phylogeny of Heteroderidae.  相似文献   

17.
18.
Three rust fungi from high mountains and pear-producing areas in Taiwan were described using morphological and molecular data based on 34 specimens. Gymnosporangium corniforme was demonstrated to produce spermogonia and aecia on Sorbus randaiensis based on molecular analyses and inoculation experiments. The pear rust pathogen G. unicorne was discovered in Taiwan for the first time. Gymnosporangium niitakayamense sp. nov. was observed on the leaves of Photinia niitakayamensis. It was distinct from other species in peridial cell wall structures, i.e., smooth outer wall, rugose side wall, and coralloid projections on the inner wall, and in having echinulate aeciospores.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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