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 359 毫秒
1.
2.
3.
In all eukaryotic cells, the endoplasmic reticulum (ER) forms a tubular network whose generation requires the fusion of ER membranes. In Arabidopsis (Arabidopsis thaliana), the membrane-bound GTPase ROOT HAIR DEFECTIVE3 (RHD3) is a potential candidate to mediate ER fusion. In addition, Arabidopsis has two tissue-specific isoforms of RHD3, namely RHD3-like (RL) proteins, and their function is not clear. Here, we show that a null allele of RHD3, rhd3-8, causes growth defects and shortened root hairs. A point mutant, rhd3-1, exhibits a more severe growth phenotype than the null mutant, likely because it exerts a dominant-negative effect on the RL proteins. Genetic analysis reveals that the double deletion of RHD3 and RL1 is lethal and that the rhd3 rl2 plants produce no viable pollen, suggesting that the RL proteins are redundant to RHD3. RHD3 family proteins can replace Sey1p, the homolog of RHD3 in yeast (Saccharomyces cerevisiae), in the maintenance of ER morphology, and they are able to fuse membranes both in vivo and in vitro. Our results suggest that RHD3 proteins mediate ER fusion and are essential for plant development and that the formation of the tubular ER network is of general physiological significance.In all eukaryotic cells, the endoplasmic reticulum (ER) comprises a continuous membrane system of sheets and tubules (Baumann and Walz, 2001; Shibata et al., 2006). ER tubules frequently connect through homotypic membrane fusion to form a reticular network (Lee and Chen, 1988; Prinz et al., 2000; Du et al., 2004). ER fusion in metazoans is mediated by the atlastins (ATLs), a class of dynamin-like, membrane-bound GTPases (Hu et al., 2009; Orso et al., 2009). ATL possesses a cytoplasmic N-terminal GTPase domain, followed by a helical domain, two closely spaced transmembrane domains, and a C-terminal cytosolic tail. ATL proteins localize mostly to ER tubules and they interact with the tubule-shaping proteins, reticulons and DP1 (Hu et al., 2009). A role for the ATLs in ER fusion is suggested by the fact that depletion of ATLs leads to long, nonbranched ER tubules in cultured cells (Hu et al., 2009) and to ER fragmentation in Drosophila melanogaster (Orso et al., 2009), possibly due to insufficient fusion between the tubules. Nonbranched ER tubules are also observed upon the expression of dominant-negative ATL mutants (Hu et al., 2009). In addition, antibodies to ATL inhibit ER network formation in Xenopus laevis egg extracts (Hu et al., 2009). Moreover, proteoliposomes containing purified D. melanogaster ATL undergo GTP-dependent fusion in vitro (Orso et al., 2009; Bian et al., 2011). The physiological significance of ER fusion is supported by the observation that mutations in human ATL1, the dominant isoform in the brain, cause hereditary spastic paraplegia (Zhao et al., 2001), a neurodegenerative disease characterized by axon shortening in corticospinal motor neurons and progressive spasticity and weakness of the lower limbs (Salinas et al., 2008).Many organisms lack ATL homologs. In yeast (Saccharomyces cerevisiae), another dynamin-like GTPase, Sey1p, has been found to share the same signature motifs and membrane topology as ATL (Hu et al., 2009). Recent work suggests that Sey1p mediates ER membrane fusion both in vivo and in vitro (Anwar et al., 2012). Cells lacking Sey1p grow normally (Hu et al., 2009), but additional mutation of an ER SNARE Ufe1p, which probably represents an alternative ER fusion mechanism in yeast, causes severe growth defects (Anwar et al., 2012). In Arabidopsis (Arabidopsis thaliana), the potential functional ortholog of ATL appears to be ROOT HAIR DEFECTIVE3 (RHD3; Hu et al., 2009), which was initially discovered by a genetic screen of root hair-defective mutants (Schiefelbein and Somerville, 1990). It is sequence related to Sey1p over the entire length (Wang et al., 1997; Brands and Ho, 2002). Mutations of RHD3 cause short and wavy root hairs (Schiefelbein and Somerville, 1990; Wang et al., 1997; Stefano et al., 2012) and defects in cell expansion (Wang et al., 2002).Despite the sequence homology between Sey1p and RHD3, it was reported that Sey1p could not replace RHD3 in plants and vice versa (Chen et al., 2011). Therefore, it is not clear whether RHD3 can mediate ER fusion. Another complication in plants is that the Arabidopsis RHD3 family also contains two RHD3-like (RL) proteins (Hu et al., 2003): RL1 is expressed only in pollen, whereas RL2 is expressed ubiquitously, but both are present at very low levels. Deletion of either RL protein causes no detectable defects in root hair development or overall growth (Chen et al., 2011). Whether RL proteins support the role of RHD3 in a tissue-specific manner remains to be investigated.Here, we have analyzed the function of RHD3 and RL proteins in Arabidopsis. We show that RHD3 and the two RL proteins play redundant roles but function during different stages of Arabidopsis development. In addition, we show that RHD3 proteins can functionally replace Sey1p in yeast and mediate ER membrane fusion.  相似文献   

4.
Xylans play an important role in plant cell wall integrity and have many industrial applications. Characterization of xylan synthase (XS) complexes responsible for the synthesis of these polymers is currently lacking. We recently purified XS activity from etiolated wheat (Triticum aestivum) seedlings. To further characterize this purified activity, we analyzed its protein composition and assembly. Proteomic analysis identified six main proteins: two glycosyltransferases (GTs) TaGT43-4 and TaGT47-13; two putative mutases (TaGT75-3 and TaGT75-4) and two non-GTs; a germin-like protein (TaGLP); and a vernalization related protein (TaVER2). Coexpression of TaGT43-4, TaGT47-13, TaGT75-3, and TaGT75-4 in Pichia pastoris confirmed that these proteins form a complex. Confocal microscopy showed that all these proteins interact in the endoplasmic reticulum (ER) but the complexes accumulate in Golgi, and TaGT43-4 acts as a scaffold protein that holds the other proteins. Furthermore, ER export of the complexes is dependent of the interaction between TaGT43-4 and TaGT47-13. Immunogold electron microscopy data support the conclusion that complex assembly occurs at specific areas of the ER before export to the Golgi. A di-Arg motif and a long sequence motif within the transmembrane domains were found conserved at the NH2-terminal ends of TaGT43-4 and homologous proteins from diverse taxa. These conserved motifs may control the forward trafficking of the complexes and their accumulation in the Golgi. Our findings indicate that xylan synthesis in grasses may involve a new regulatory mechanism linking complex assembly with forward trafficking and provide new insights that advance our understanding of xylan biosynthesis and regulation in plants.It is believed that Golgi-localized, multiprotein complexes synthesize plant hemicellulosic polysaccharides, including xylans. Such complexes are not well characterized in plants (Zeng et al., 2010; Atmodjo et al., 2011; Chou et al., 2012), which is in sharp contrast with mammalian and yeast cells (Jungmann and Munro, 1998; McCormick et al., 2000; Giraudo et al., 2001). Xylans are the most abundant plant hemicellulosic polysaccharides on Earth and play an important role in the integrity of cell walls, which is a key factor in plant growth. Any mutations affecting xylan backbone biosynthesis seem to result in abnormal growth of plants due mostly to thinning and weakening of secondary xylem walls, described as the irregular xylem (irx) phenotype. Thus, characterizing the xylan synthase complex (XSC) would have an impact on plant improvement, as well as many industrial applications related to food, feed, and biofuel production (Yang and Wyman, 2004; Faik, 2010). Although the Arabidopsis (Arabidopsis thaliana) irx mutants have revealed the involvement of several glycosyltransferase (GT) gene families in xylan biosynthesis (Brown et al., 2007, 2009; Lee et al., 2007, 2010; Wu et al., 2009, 2010), no XSCs have been purified/isolated from Arabidopsis tissues, and we still do not know whether some of the identified Arabidopsis GTs can assemble into functional XSCs. Furthermore, if GTs do assemble into XSCs, we don’t know the mechanisms by which plant cells control their assembly and cellular trafficking. In contrast to dicots, xylan synthase activity was recently immunopurified from etiolated wheat (Triticum aestivum) microsomes (Zeng et al., 2010). This purified wheat XS activity was shown to catalyze three activities, xylan-glucuronosyltransferase (XGlcAT), xylan-xylosyltransferase (XXylT), and xylan-arabinofuranosyltranferase (XAT), which work synergistically to synthesize xylan-type polymers in vitro (Zeng et al., 2008, 2010). This work focuses on describing protein composition, assembly, and trafficking of this purified wheat XS activity.In all eukaryotes, proteins of the secretory pathway (including GTs) are synthesized in the endoplasmic reticulum (ER) and modified as they go through the Golgi cisternae. Most proteins exit the ER from ER export sites (ERESs; Hanton et al., 2009) and use a signal-based sorting mechanism that allows them to be selectively recruited into vesicles coated by coat protein II complexes (Barlowe, 2003; Beck et al., 2008). For many Golgi-resident type II membrane proteins, di-Arg motifs, such as RR, RXR, and RRR located in their cytosolic NH2-terminal ends, have been shown to be required for their ER export (Giraudo et al., 2003; Czlapinski and Bertozzi, 2006; Schoberer et al., 2009; Tu and Banfield, 2010). Interestingly, di-Arg motifs located ∼40 amino acids from the membrane on the cytosolic side can also be used to retrieve some type II ER-resident proteins from cis-Golgi (Schutze et al., 1994; Hardt et al., 2003; Boulaflous et al., 2009). In contrast to the signal-based sorting mechanism involved in trafficking between the ER and Golgi, the steady-state localization/retention of proteins (including GTs) in the Golgi is thought to occur through vesicular cycling. Cycling is influenced by various mechanisms, including the length and composition of the transmembrane domain (TMD) of type II GTs (Bretscher and Munro, 1993; Colley, 1997; van Vliet et al., 2003; Sousa et al., 2003; Sharpe et al., 2010), and the oligomerization/aggregation of GTs (kin hypothesis), which suggests that formation of homo- or heterooligomers of GTs in the Golgi may prevent their recruitment into clathrin-coated vesicles (Machamer, 1991; Nilsson et al., 1993; Weisz et al., 1993; Cole et al., 1996). Some Golgi-resident GTs are predicted to have a cleavable NH2-terminal secretion signal peptide (SP) and would therefore exist as soluble proteins in the Golgi lumen. To maintain their proper Golgi localization, these processed GTs are likely part of multiprotein complexes anchored to integral membrane proteins. The fact that homologs of many of the trafficking proteins from mammalian and yeast cells are found in plants indicates that trafficking machineries of the plant secretory pathway are likely conserved (d’Enfert et al., 1992; Bar-Peled and Raikhel, 1997; Batoko et al., 2000; Pimpl et al., 2000; Phillipson et al., 2001; Hawes et al., 2008).It is becoming increasingly evident that understanding the mechanisms controlling protein-protein interaction, sorting, and trafficking of polysaccharide synthases (including XSCs) will help elucidate how plants regulate cell wall synthesis and deposition during their development. To this end, we believe that the purified wheat XS activity (Zeng et al., 2010) is an excellent model for this type of study. In this work, proteomics was used to determine the protein composition of the purified XS activity. Confocal microscopy and immunogold transmission electron microscopy (TEM) were used to investigate the assembly and trafficking of the complex. Our proteomics data showed that the purified activity contains two GTs, TaGT43-4 and TaGT47-13, two putative mutases, TaGT75-3 and TaGT75-4, and two non-GT proteins: a germin-like protein (TaGLP) belonging to cupin superfamily and a protein specific to monocots annotated as wheat vernalization-related protein 2 (TaVER2). Microscopy analyses revealed that all these proteins interact in the ER, but the assembled complexes accumulate in the Golgi. Export of these complexes from the ER is controlled by the interaction between TaGT43-4 and TaGT47-13. Characterization of the wheat XSC and its trafficking furthers our understanding of xylan biosynthesis in grasses and helps elucidate how polysaccharide synthase complexes are assembled, sorted, and maintained in different compartments of the secretory pathway.  相似文献   

5.
6.
Arabidopsis (Arabidopsis thaliana) has eight glycerol-3-phosphate acyltransferase (GPAT) genes that are members of a plant-specific family with three distinct clades. Several of these GPATs are required for the synthesis of cutin or suberin. Unlike GPATs with sn-1 regiospecificity involved in membrane or storage lipid synthesis, GPAT4 and -6 are unique bifunctional enzymes with both sn-2 acyltransferase and phosphatase activity resulting in 2-monoacylglycerol products. We present enzymology, pathway organization, and evolutionary analysis of this GPAT family. Within the cutin-associated clade, GPAT8 is demonstrated as a bifunctional sn-2 acyltransferase/phosphatase. GPAT4, -6, and -8 strongly prefer C16:0 and C18:1 ω-oxidized acyl-coenzyme As (CoAs) over unmodified or longer acyl chain substrates. In contrast, suberin-associated GPAT5 can accommodate a broad chain length range of ω-oxidized and unsubstituted acyl-CoAs. These substrate specificities (1) strongly support polyester biosynthetic pathways in which acyl transfer to glycerol occurs after oxidation of the acyl group, (2) implicate GPAT specificities as one major determinant of cutin and suberin composition, and (3) argue against a role of sn-2-GPATs (Enzyme Commission 2.3.1.198) in membrane/storage lipid synthesis. Evidence is presented that GPAT7 is induced by wounding, produces suberin-like monomers when overexpressed, and likely functions in suberin biosynthesis. Within the third clade, we demonstrate that GPAT1 possesses sn-2 acyltransferase but not phosphatase activity and can utilize dicarboxylic acyl-CoA substrates. Thus, sn-2 acyltransferase activity extends to all subbranches of the Arabidopsis GPAT family. Phylogenetic analyses of this family indicate that GPAT4/6/8 arose early in land-plant evolution (bryophytes), whereas the phosphatase-minus GPAT1 to -3 and GPAT5/7 clades diverged later with the appearance of tracheophytes.sn-Glycerol-3-phosphate 1-O-acyltransferase (GPAT; Enzyme Commission [EC] 2.3.1.15) is the first enzyme in the pathway for the de novo synthesis of membrane and storage lipids. It catalyzes the transfer of an acyl group from acyl-CoA or acyl-ACP to the sn-1 position of sn-glycerol-3-phosphate (G3P). This reaction has been extensively characterized in bacteria, fungi, animals, and plants (Murata and Tasaka, 1997; Zheng and Zou, 2001; Gimeno and Cao, 2008; Zhang and Rock, 2008; Wendel et al., 2009; Chen et al., 2011a). In the Arabidopsis (Arabidopsis thaliana) genome, there are 10 genes annotated as GPATs. One of these is the soluble, plastid-localized GPAT (At1g32200) that utilizes acyl-ACP substrates and exhibits sn-1 acyl transfer regiospecificity (Nishida et al., 1993). A second enzyme is GPAT9 (At5g60620), which is localized to the endoplasmic reticulum (Gidda et al., 2009) and may be an acyl-CoA-dependent sn-1 GPAT that enables nonplastid glycerolipid synthesis. The remaining eight GPATs cluster together in a family (Zheng et al., 2003; Beisson et al., 2007; Gidda et al., 2009) that is not required for membrane or storage lipid biosynthesis; instead, several members of the family clearly affect the composition and quantity of cutin or suberin (Beisson et al., 2012).Cutin and suberin are extracellular lipid barriers deposited by certain types of plant cells. These insoluble polymers, and associated waxes, function to control water, gas, and ion fluxes and serve as physical barriers to protect plants from pathogen invasion (Kolattukudy, 2001; Schreiber, 2010; Ranathunge et al., 2011). From an evolutionary perspective, the appearance of these lipid barriers was likely a requirement for the adaptation of plants to a terrestrial environment (Rensing et al., 2008). ω-oxidized fatty acids and glycerol are usually major constituents of both polymers (Bernards, 2002; Graça and Santos, 2007; Pollard et al., 2008). The detailed structures of cutin and suberin polymers are still largely unknown (Pollard et al., 2008), but direct esterification of fatty acids to glycerol and to each other has been demonstrated in a large number of partial depolymerization studies of cutin and suberin (Graça and Santos, 2007; Graça and Lamosa, 2010). In Arabidopsis, GPAT4 and GPAT8 are required for the accumulation of C16 and C18 ω-hydroxy fatty acid (ω-OHFA) and α,ω-dicarboxylic acid (DCA) cutin monomers in stems and leaves (Li et al., 2007a). GPAT6 is required for the incorporation of the following C16 monomers: 10,16-dihydroxypalmitate (10,16-diOH C16:0-FA), hexadecane-1,16-dioic acid (C16:0-DCA), and 16-hydroxypalmitate (16-OH C16:0-FA), into flower cutin (Li-Beisson et al., 2009). (For a fatty acid, the abbreviation used is Cm:n-FA, where m is the number of carbon atoms and n is the number of double bonds. The position and number of hydroxyl groups precedes this notation. The same nomenclature is used for DCAs.) GPAT5 controls the accumulation of C22:0- and C24:0-FA, ω-OHFA, and DCA monomers in the suberin of roots and seed coats (Beisson et al., 2007). Recently, we demonstrated that GPAT4 and -6 are bifunctional enzymes that possess sn-2 acyltransferase and phosphatase activities (Yang et al., 2010) and that therefore produce sn-2 monoacylglycerols (MAGs) as the major product. GPAT5 also exhibits strong preference for sn-2 acylation but lacks phosphatase activity; thus, sn-2 lysophosphatidic acids (LPAs) are its major product (Yang et al., 2010).These observations all attest to the fact that several members of the GPAT1 to -8 family are enzymatically very distinct from the GPATs required for membrane and storage lipid biosynthesis. Indeed, they represent a new acylglycerol biosynthesis pathway that provides precursors for cutin and suberin biosynthesis. To better understand the early steps in polyester synthesis and the roles contributed by GPAT4 to -8, and to determine whether the clade of GPAT1 to -3 has distinct or similar activity, we have characterized the regiochemistry and acyl substrate specificity of GPATs representing all three clades. We show that the cutin-associated GPAT8 is a bifunctional sn-2 acyltransferase/phosphatase, while GPAT1, an isozyme with uncertain function but important for tapetum and anther development (Zheng et al., 2003; Li et al., 2012), possesses sn-2 acyltransferase activity but not phosphatase activity. An important issue in defining the pathway of cutin/suberin biosynthesis is whether to place the P450 oxidation reactions before or after the G3P acylation reactions. As discussed (Pollard et al., 2008), previous evidence has not allowed definitive determination of the alternative pathways. However, conducting a GPAT substrate specificity study, particularly with a range of ω-oxidized and unmodified acyl-CoAs can help clarify the situation. Here, we show the acyl-CoA specificities of GPAT4, -5, -6, and -8 are concordant with the compositions of their respective cutins and suberins and the resulting changes in corresponding gpat mutants and overexpression lines. Furthermore, the results provide strong evidence that acyl transfer to glycerol by GPAT occurs after ω-oxidation of acyl chains, thus increasing our limited understanding of the biochemical pathway for cutin and suberin polymer assembly.A phylogenetic analysis of Arabidopsis GPATs with vascular and nonvascular land-plant homologs provides an evolutionary view of the expansion and divergence of the gene family into three distinct clades that are associated with morphological and functional evolution and with the loss of phosphatase activity.  相似文献   

7.
8.
9.
The endoplasmic reticulum (ER) is a network of tubules and sheet-like structures in eukaryotic cells. Some ER tubules dynamically change their morphology, and others form stable structures. In plants, it has been thought that the ER tubule extension is driven by the actin-myosin machinery. Here, we show that microtubules also contribute to the ER tubule extension with an almost 20-fold slower rate than the actin filament-based ER extension. Treatment with the actin-depolymerizing drug Latrunculin B made it possible to visualize the slow extension of the ER tubules in transgenic Arabidopsis (Arabidopsis thaliana) plants expressing ER-targeted green fluorescent protein. The ER tubules elongated along microtubules in both directions of microtubules, which have a distinct polarity. This feature is similar to the kinesin- or dynein-driven ER tubule extension in animal cells. In contrast to the animal case, ER tubules elongating with the growing microtubule ends were not observed in Arabidopsis. We also found the spots where microtubules are stably colocalized with the ER subdomains during long observations of 1,040 s, suggesting that cortical microtubules contribute to provide ER anchoring points. The anchoring points acted as the branching points of the ER tubules, resulting in the formation of multiway junctions. The density of the ER tubule junction positively correlated with the microtubule density in both elongating cells and mature cells of leaf epidermis, showing the requirement of microtubules for formation of the complex ER network. Taken together, our findings show that plants use microtubules for ER anchoring and ER tubule extension, which establish fine network structures of the ER within the cell.The endoplasmic reticulum (ER) is a complex network composed of tubules and sheet structures. The ER network’s morphology changes dynamically by elongation and shrinkage of tubules, sheet expansion, and sliding junctions. For example, an ER tubule elongates straight forward from a cisterna and subsequently, fuses to another cisterna, producing a linkage between two cisternae. If an elongating tubule fails to fuse to another cisterna, the tubule contracts into the original cisterna. However, the ER has stable anchoring points that associate with other cellular structures, such as the plasma membrane or cytoskeleton. When an elongating ER tubule reaches an association point, it forms a stable ER anchor (i.e. establishment of the ER anchoring points forms stable ER tubules). Hence, increasing the number of ER anchoring points produces fine ER meshwork.ER dynamics in eukaryotes depend on the cytoskeleton. In plants, major contributors for ER organization are actin filaments (Quader et al., 1989; Knebel et al., 1990; Lichtscheidl and Hepler, 1996; Sparkes et al., 2009a) and the actin-associated motor proteins (myosins; Prokhnevsky et al., 2008; Peremyslov et al., 2010; Ueda et al., 2010). However, it had generally been thought that microtubules are not involved in ER organization in plants, because microtubule-depolymerizing drugs do not induce obvious changes in the ER network (Quader et al., 1989; Knebel et al., 1990; Lichtscheidl and Hepler, 1996; Sparkes et al., 2009a). Nevertheless, involvement of microtubules in plant ER organization has been suspected from several electron microscopy observations that showed microtubules located close to the ER membrane in Vicia faba guard cells, Nicotiana alata pollen tubes, and Funaria hygrometrica caulonemata (Lancelle et al., 1987; Hepler et al., 1990; McCauley and Hepler, 1992).Foissner et al. (2009) have suggested that microtubules are involved in motility and orientation of cortical ER in Characean algae (Nitella translucens, Nitella flexilis, Nitella hyalina, and Nitella pseudoflabellata) internodal cells. Characean cortical ER is spatially separated from inner cytoplasmic streaming by the middle layer of fixed chloroplasts. The cortical ER forms a tight meshwork of predominantly transverse ER tubules that frequently coalign with microtubules, and microtubule depolymerization reduces the transverse ER tubules and increases mesh size (Foissner et al., 2009). Consistently, Hamada et al. (2012) have shown in Arabidopsis (Arabidopsis thaliana) that microtubule depolymerization increases mesh size in young elongating cells. In addition, stable ER tubule junctions are often colocalized with cortical microtubules (Hamada et al., 2012), suggesting that microtubules stabilize ER tubule junctions to form fine ER meshes. Oryzalin-induced ER nodulation (Langhans et al., 2009) was not observed in our experimental conditions.Here, we showed that ER tubules elongate along microtubules in plant cells. In addition, we revealed that the ER is stably anchored to defined points on cortical microtubules. The stable anchoring points are the basis of various ER shapes, such as three-way, two-way, or dead-end ER tubules. These microtubule-ER interactions, together with the actin-myosin system, contribute to ER network organization.  相似文献   

10.
The endoplasmic reticulum (ER) is a ubiquitous organelle that plays roles in secretory protein production, folding, quality control, and lipid biosynthesis. The cortical ER in plants is pleomorphic and structured as a tubular network capable of morphing into flat cisternae, mainly at three-way junctions, and back to tubules. Plant reticulon family proteins (RTNLB) tubulate the ER by dimerization and oligomerization, creating localized ER membrane tensions that result in membrane curvature. Some RTNLB ER-shaping proteins are present in the plasmodesmata (PD) proteome and may contribute to the formation of the desmotubule, the axial ER-derived structure that traverses primary PD. Here, we investigate the binding partners of two PD-resident reticulon proteins, RTNLB3 and RTNLB6, that are located in primary PD at cytokinesis in tobacco (Nicotiana tabacum). Coimmunoprecipitation of green fluorescent protein-tagged RTNLB3 and RTNLB6 followed by mass spectrometry detected a high percentage of known PD-localized proteins as well as plasma membrane proteins with putative membrane-anchoring roles. Förster resonance energy transfer by fluorescence lifetime imaging microscopy assays revealed a highly significant interaction of the detected PD proteins with the bait RTNLB proteins. Our data suggest that RTNLB proteins, in addition to a role in ER modeling, may play important roles in linking the cortical ER to the plasma membrane.The endoplasmic reticulum (ER) is a multifunctional organelle (Hawes et al., 2015) and is the site of secretory protein production, folding, and quality control (Brandizzi et al., 2003) and lipid biosynthesis (Wallis and Browse, 2010), but it is also involved in many other aspects of day-to-day plant life, including auxin regulation (Friml and Jones, 2010) and oil and protein body formation (Huang, 1996; Herman, 2008). The cortical ER network displays a remarkable polygonal arrangement of motile tubules that are capable of morphing into small cisternae, mainly at the three-way junctions of the ER network (Sparkes et al., 2009). The cortical ER network of plants has been shown to play multiple roles in protein trafficking (Palade, 1975; Vitale and Denecke, 1999) and pathogen responses (for review, see Pattison and Amtmann, 2009; Beck et al., 2012).In plants, the protein family of reticulons (RTNLBs) contributes significantly to tubulation of the ER (Tolley et al., 2008, 2010; Chen et al., 2012). RTNLBs are integral ER membrane proteins that feature a C-terminal reticulon homology domain (RHD) that contains two major hydrophobic regions. These regions form two V-shaped transmembrane wedges joined together via a cytosolic loop, with the C and N termini of the protein facing the cytosol. RTNLBs can dimerize or oligomerize, creating localized tensions in the ER membrane, inducing varying degrees of membrane curvature (Sparkes et al., 2010). Hence, RTNLBs are considered to be essential in maintaining the tubular ER network.The ability of RTNLBs to constrict membranes is of interest in the context of cell plate development and the formation of primary plasmodesmata (PD; Knox et al., 2015). PD formation involves extensive remodeling of the cortical ER into tightly furled tubules to form the desmotubules, axial structures that run through the PD pore (Overall and Blackman, 1996; Ehlers and Kollmann, 2001). At only 15 nm in diameter, the desmotubule is one of the most constricted membrane structures found in nature, with no animal counterparts (Tilsner et al., 2011). PD are membrane-rich structures characterized by a close association of the plasma membrane (PM) with the ER. The forces that model the ER into desmotubules, however, are poorly understood. RTNLBs are excellent candidates for this process and can constrict fluorescent protein-labeled ER membranes into extremely fine tubules (Sparkes et al., 2010). We showed recently that two of the RTNLBs present in the PD proteome, RTNLB3 and RTNLB6 (Fernandez-Calvino et al., 2011), are present in primary PD at cytokinesis (Knox et al., 2015). However, nothing is known of the proteins that interact with RTNLBs identified in the PD proteome or that may link RTNLBs to the PM. To date, the only protein shown to bind to plant RTNLBs is RHD3-LIKE2, the plant homolog of the ER tubule fusion protein ATLASTIN (Lee et al., 2013).Here, we used a dual approach to identify interacting partners of RTNLB3 and RTNLB6 (Fernandez-Calvino et al., 2011; Knox et al.., 2015). First, we used GFP immunoprecipitation assays coupled to mass spectrometry (MS) to identify proteins potentially binding to RTNLB3 and RTNLB6. Second, from the proteins we identified, we conducted a detailed Förster resonance energy transfer by fluorescence lifetime imaging microscopy (FRET-FLIM) analysis to confirm prey-bait interactions in vivo.The application of time-resolved fluorescence spectroscopy to imaging biological systems has allowed the design and implementation of fluorescence lifetime imaging microscopy (FLIM). The technique allows measuring and determining the space map of picosecond fluorescence decay at each pixel of the image through confocal single and multiphoton excitation. The general fluorescence or Förster resonance energy transfer (FRET) to determine the colocalization of two color chromophores can now be improved to determine physical interactions using FRET-FLIM and protein pairs tagged with appropriate GFP fluorophores and monomeric red fluorescent protein (mRFP). FRET-FLIM measures the reduction in the excited-state lifetime of GFP (donor) fluorescence in the presence of an acceptor fluorophore (e.g. mRFP) that is independent of the problems associated with steady-state intensity measurements. The observation of such a reduction is an indication that the two proteins are within a distance of 1 to 10 nm, thus indicating a direct physical interaction between the two protein fusions (Osterrieder et al., 2009; Sparkes et al., 2010; Schoberer and Botchway, 2014). It was shown previously that a reduction of as little as approximately 200 ps in the excited-state lifetime of the GFP-labeled protein represents quenching through a protein-protein interaction (Stubbs et al., 2005).Our interaction data identified a large percentage (40%) of ER proteins, including other RTNLB family members. However, we also found a relatively large number (25%) of proteins present in the published PD proteome (Fernandez-Calvino et al., 2011) and a surprisingly high proportion (35%) of PM proteins. Of the PD-resident proteins we identified, a significant number were shown previously to be targets of viral movement proteins (MPs) or proteins present within lipid rafts, consistent with the view that PD are lipid-rich microdomains (Bayer et al., 2014). Additional proteins identified suggested roles for RTNLBs in transport and pathogen defense. We suggest that RTNLBs may play key roles in anchoring and/or signaling between the cortical ER and PM.  相似文献   

11.
12.
The growing relevance of plants for the production of recombinant proteins makes understanding the secretory machinery, including the identification of glycosylation sites in secreted proteins, an important goal of plant proteomics. Barley (Hordeum vulgare) aleurone layers maintained in vitro respond to gibberellic acid by secreting an array of proteins and provide a unique system for the analysis of plant protein secretion. Perturbation of protein secretion in gibberellic acid-induced aleurone layers by two independent mechanisms, heat shock and tunicamycin treatment, demonstrated overlapping effects on both the intracellular and secreted proteomes. Proteins in a total of 22 and 178 two-dimensional gel spots changing in intensity in extracellular and intracellular fractions, respectively, were identified by mass spectrometry. Among these are proteins with key roles in protein processing and secretion, such as calreticulin, protein disulfide isomerase, proteasome subunits, and isopentenyl diphosphate isomerase. Sixteen heat shock proteins in 29 spots showed diverse responses to the treatments, with only a minority increasing in response to heat shock. The majority, all of which were small heat shock proteins, decreased in heat-shocked aleurone layers. Additionally, glycopeptide enrichment and N-glycosylation analysis identified 73 glycosylation sites in 65 aleurone layer proteins, with 53 of the glycoproteins found in extracellular fractions and 36 found in intracellular fractions. This represents major progress in characterization of the barley N-glycoproteome, since only four of these sites were previously described. Overall, these findings considerably advance knowledge of the plant protein secretion system in general and emphasize the versatility of the aleurone layer as a model system for studying plant protein secretion.Plant proteins that are secreted to the apoplast have important functions in signaling, defense, and cell regulation. The classical protein secretory pathway is less characterized in plants than in mammals or yeast but is of growing interest due to the potential of plant systems as hosts for the production of recombinant proteins (Erlendsson et al., 2010; De Wilde et al., 2013). Plant secretomics, therefore, is a rapidly expanding area applied to gain further insight into these processes (Agrawal et al., 2010; Alexandersson et al., 2013). Many secretory proteins contain putative N-glycosylation sites, and the identification and characterization of these sites is an important element in secretomics analysis. However, to date, only a few plant glycoproteomes have been described (Fitchette et al., 2007; Minic et al., 2007; Palmisano et al., 2010; Melo-Braga et al., 2012; Zhang et al., 2012; Thannhauser et al., 2013).The cereal aleurone layer is of major importance due to its central role in grain germination. Previous proteomic studies have been reported for aleurone layers dissected from mature (Finnie and Svensson, 2003) and germinating (Bønsager et al., 2007) barley (Hordeum vulgare) or developing (Tasleem-Tahir et al., 2011) and mature (Laubin et al., 2008; Jerkovic et al., 2010; Meziani et al., 2012) wheat (Triticum aestivum) grains. The in vitro culture of isolated aleurone layers was developed by Chrispeels and Varner (1967). Since then, this system has become an excellent tool for the study of germination signaling in response to phytohormones (Bush et al., 1986; Jones and Jacobsen, 1991; Bethke et al., 1997; Ishibashi et al., 2012). More recently, it has been adopted as a unique system for the analysis of plant secretory proteins (Hägglund et al., 2010; Finnie et al., 2011). The addition of GA3 to the isolated aleurone layers induces the synthesis and secretion of hydrolytic enzymes. In the in vitro system, these accumulate in the incubation buffer, facilitating their identification and characterization using proteomics techniques. Thus, numerous secreted proteins with roles in the hydrolysis of starch, cell wall polysaccharides, and proteins could be identified (Finnie et al., 2011). Furthermore, several of the proteins were also detected in intracellular extracts from the same aleurone layers, presumably prior to their release. Many of the proteins appeared in multiple forms on two-dimensional (2D) gels, and often with higher Mr than expected, suggesting the presence of posttranslational modifications (PTMs; Finnie et al., 2011). Bak-Jensen et al. (2007) and Finnie et al. (2011) observed a highly complex pattern of α-amylase-containing spots on 2D gels, which originated from only two α-AMYLASE2 (AMY2) and two AMY1 gene products from a total of 10 genes, probably reflecting multiple forms due to PTMs.In eukaryotic cells, proteins synthesized in the endoplasmic reticulum (ER) must be correctly folded and assembled before continuing in the secretory pathway. Perturbations of redox state, calcium regulation, Glc deprivation, and viral infection can lead to ER stress, triggered by the accumulation of unfolded and misfolded proteins in the ER lumen. This provokes a triple response from the cell, consisting of an up-regulation of chaperones and vesicle trafficking, a down-regulation of genes encoding secretory proteins, and an up-regulation of proteins involved in endoplasmic reticulum-associated protein degradation (ERAD). In plants, the molecular mechanisms underlying ER stress in plants have yet to be fully resolved (Martínez and Chrispeels, 2003; Nagashima et al., 2011; Moreno et al., 2012).Tunicamycin (TN), an inhibitor of GlcNAc phosphotransferase, which catalyzes the first step in glycoprotein synthesis, has been used to induce ER stress by causing the accumulation of unfolded proteins in the ER lumen (Noh et al., 2003; Kamauchi et al., 2005; Reis et al., 2011). If unfolded proteins are not removed, the prolonged stress will induce programmed cell death. Links between ER stress and apoptosis have been reported in response to TN treatment in mammalian cells, whereas in plants, this correlation has been suggested, but the pathways of signal transduction remain unknown (Kamauchi et al., 2005; Reis et al., 2011).In all plant tissues, heat shock (HS) induces the synthesis of a variety of heat shock proteins (HSPs), which are responsible for protein refolding under stress conditions (Craig et al., 1994) and translocation and degradation in a broad array of normal cellular processes (Bond and Schlesinger, 1986; Spiess et al., 1999). In the barley aleurone layer, HS selectively suppresses the synthesis of secretory proteins, including α-amylase, due to the selective destabilization of secretory protein mRNA (Belanger et al., 1986; Brodl and Ho, 1991). However, an acclimation effect has been described in aleurone cells after prolonged incubation at warm temperatures, resulting in a resumption of the protein secretory machinery (Shaw and Brodl, 2003). The connection between heat stress response and ER stress has been well established in mammals and yeast, but scarce information is available in plants (Denecke et al., 1995).Over the last years, the dual role of reactive oxygen species (ROS) has been established in plants: at higher concentrations, ROS act as toxic molecules damaging cellular macromolecules, eventually causing cell death, but at lower concentrations, ROS seem to be necessary for seed germination and seedling growth by controlling the cellular redox status, regulating growth and protecting against pathogens (Bailly, 2004; Bailly et al., 2008). In the barley aleurone layer, GA3 perceived at the plasma membrane induces ROS generation as a by-product from intense lipid metabolism, and the redox regulation of the GA3-induced response has been proposed (Maya-Ampudia and Bernal-Lugo, 2006). This suggests that during the secretory function of the tissue, moderate levels of ROS may be acting as cellular messengers. In aleurone cells, ROS, especially hydrogen peroxide (H2O2), are involved in the process of programmed cell death, but the molecular mechanisms remain unclear (Bethke and Jones, 2001; Ishibashi et al., 2012).Until now, none of the protein components of the ER stress pathways have been identified in barley; also, little is known about the glycosylation of barley proteins. In this work, numerous N-glycoslation sites are identified, and the effect of perturbing N-glycosylation and the secretory pathway by TN and HS treatments is analyzed in GA3-induced barley aleurone layers.  相似文献   

13.
The multifunctional movement protein (MP) of Tomato mosaic tobamovirus (ToMV) is involved in viral cell-to-cell movement, symptom development, and resistance gene recognition. However, it remains to be elucidated how ToMV MP plays such diverse roles in plants. Here, we show that ToMV MP interacts with the Rubisco small subunit (RbCS) of Nicotiana benthamiana in vitro and in vivo. In susceptible N. benthamiana plants, silencing of NbRbCS enabled ToMV to induce necrosis in inoculated leaves, thus enhancing virus local infectivity. However, the development of systemic viral symptoms was delayed. In transgenic N. benthamiana plants harboring Tobacco mosaic virus resistance-22 (Tm-22), which mediates extreme resistance to ToMV, silencing of NbRbCS compromised Tm-22-dependent resistance. ToMV was able to establish efficient local infection but was not able to move systemically. These findings suggest that NbRbCS plays a vital role in tobamovirus movement and plant antiviral defenses.Plant viruses use at least one movement protein (MP) to facilitate viral spread between plant cells via plasmodesmata (PD; Lucas and Gilbertson, 1994; Ghoshroy et al., 1997). Among viral MPs, the MP of tobamoviruses, such as Tobacco mosaic virus (TMV) and its close relative Tomato mosaic virus (ToMV), is the best characterized. TMV MP specifically accumulates in PD and modifies the plasmodesmatal size exclusion limit in mature source leaves or tissues (Wolf et al., 1989; Deom et al., 1990; Ding et al., 1992). TMV MP and viral genomic RNA form a mobile ribonucleoprotein complex that is essential for cell-to-cell movement of viral infection (Watanabe et al., 1984; Deom et al., 1987; Citovsky et al., 1990, 1992; Kiselyova et al., 2001; Kawakami et al., 2004; Waigmann et al., 2007). TMV MP also enhances intercellular RNA silencing (Vogler et al., 2008) and affects viral symptom development, host range, and host susceptibility to virus (Dardick et al., 2000; Bazzini et al., 2007). Furthermore, ToMV MP is identified as an avirulence factor that is recognized by tomato (Solanum lycopersicum) resistance proteins Tobacco mosaic virus resistance-2 (Tm-2) and Tm-22 (Meshi et al., 1989; Lanfermeijer et al., 2004). Indeed, tomato Tm-22 confers extreme resistance against TMV and ToMV in tomato plants and even in heterologous tobacco (Nicotiana tabacum) plants (Lanfermeijer et al., 2003, 2004).To date, several host factors that interact with TMV MP have been identified. These TMV MP-binding host factors include cell wall-associated proteins such as pectin methylesterase (Chen et al., 2000), calreticulin (Meshi et al., 1989), ANK1 (Ueki et al., 2010), and the cellular DnaJ-like protein MPIP1 (Shimizu et al., 2009). Many cytoskeletal components such as actin filaments (McLean et al., 1995), microtubules (Heinlein et al., 1995), and the microtubule-associated proteins MPB2C (Kragler et al., 2003) and EB1a (Brandner et al., 2008) also interact with TMV MP. Most of these factors are involved in TMV cell-to-cell movement.Rubisco catalyzes the first step of CO2 assimilation in photosynthesis and photorespiration. The Rubisco holoenzyme is a heteropolymer consisting of eight large subunits (RbCLs) and eight small subunits (RbCSs). RbCL was reported to interact with the coat protein of Potato virus Y (Feki et al., 2005). Both RbCS and RbCL were reported to interact with the P3 proteins encoded by several potyviruses, including Shallot yellow stripe virus, Onion yellow dwarf virus, Soybean mosaic virus, and Turnip mosaic virus (Lin et al., 2011). Proteomic analysis of the plant-virus interactome revealed that RbCS participates in the formation of virus complexes of Rice yellow mottle virus (Brizard et al., 2006). However, the biological function of Rubisco in viral infection remains unknown.In this study, we show that RbCS plays an essential role in virus movement, host susceptibility, and Tm-22-mediated extreme resistance in the ToMV-host plant interaction.  相似文献   

14.
Like many other viruses, Tobacco mosaic virus replicates in association with the endoplasmic reticulum (ER) and exploits this membrane network for intercellular spread through plasmodesmata (PD), a process depending on virus-encoded movement protein (MP). The movement process involves interactions of MP with the ER and the cytoskeleton as well as its targeting to PD. Later in the infection cycle, the MP further accumulates and localizes to ER-associated inclusions, the viral factories, and along microtubules before it is finally degraded. Although these patterns of MP accumulation have been described in great detail, the underlying mechanisms that control MP fate and function during infection are not known. Here, we identify CELL-DIVISION-CYCLE protein48 (CDC48), a conserved chaperone controlling protein fate in yeast (Saccharomyces cerevisiae) and animal cells by extracting protein substrates from membranes or complexes, as a cellular factor regulating MP accumulation patterns in plant cells. We demonstrate that Arabidopsis (Arabidopsis thaliana) CDC48 is induced upon infection, interacts with MP in ER inclusions dependent on the MP N terminus, and promotes degradation of the protein. We further provide evidence that CDC48 extracts MP from ER inclusions to the cytosol, where it subsequently accumulates on and stabilizes microtubules. We show that virus movement is impaired upon overexpression of CDC48, suggesting that CDC48 further functions in controlling virus movement by removal of MP from the ER transport pathway and by promoting interference of MP with microtubule dynamics. CDC48 acts also in response to other proteins expressed in the ER, thus suggesting a general role of CDC48 in ER membrane maintenance upon ER stress.Plant viruses are obligate intracellular pathogens that replicate in association with host membranes (Laliberté and Sanfaçon, 2010) and subvert host intra- and intercellular trafficking pathways to achieve cell-to-cell and systemic spread (Harries and Ding, 2011; Niehl and Heinlein, 2011). In the case of the well-studied Tobacco mosaic virus (TMV), viral replication factories form on membranes of the endoplasmic reticulum (ER; Heinlein et al., 1995, 1998). As the plant ER is continuous between cells through plasmodesmata (PD; Ding et al., 1992), this membrane network provides a direct pathway for the spread of replicated virus from the replication sites in infected cells into the ER network of noninfected cells. The spread of plant viruses depends on virus-encoded movement proteins (MPs; Deom et al., 1987; Lucas, 2006). The MP of TMV facilitates the cell-to-cell passage of the infectious particle by forming a ribonucleoprotein complex with the viral RNA (Citovsky et al., 1990) and by increasing the size exclusion limit of PD (Wolf et al., 1989).During the course of infection, as well as when ectopically expressed, the MP associates with PD, the ER/actin network, and microtubules (Heinlein et al., 1995, 1998; Reichel and Beachy, 1998; Wright et al., 2007; Sambade et al., 2008; Hofmann et al., 2009; Boutant et al., 2010; Peña and Heinlein, 2012; Supplemental Fig. S1). Shortly after infection of a new cell, the MP localizes to small, mobile, ER-associated particles proposed to play a role in PD targeting of the viral RNA (Boyko et al., 2007; Sambade et al., 2008). Similar small, mobile MP particles are observed early upon ectopic expression of the protein. These particles colocalize with RNA and undergo stop-and-go movements in association with the ER (Sambade et al., 2008). The particle movements pause at microtubule proximal sites and their detachment requires microtubule polymerization (Sambade et al., 2008). These observations suggest that the interaction with the microtubule system plays a critical role in the maturation and ER-mediated delivery of infectious viral RNA particles to PD during early infection stages. Consistently, tobacco (Nicotiana tabacum) mutants with reduced microtubule dynamics exhibit reduced TMV movement (Ouko et al., 2010). Following virus movement, the previously infected cell further accumulates MP at the ER, a process that coincides with the formation of large ER inclusions that contain viral replicase and viral RNA in addition to MP and likely function as virus factories (Heinlein et al., 1998; Más and Beachy, 1999). In mature form, these inclusions may represent the so-called viroplasms or X-bodies described in the classical literature (Bawden and Sheffield, 1939; Esau and Cronshaw, 1967; Hills et al., 1987). Their formation is associated with rearrangements of the ER membrane and likely mediated by the accumulated MP since the inclusions diminish and reconstitute a native ER structure when MP becomes degraded by the 26S proteasome (Reichel and Beachy, 1998, 2000). Transfected cells accumulate MP in similar inclusions as those formed during infection, indicating that accumulated MP is indeed necessary and sufficient to form inclusions in association with the ER (Reichel and Beachy, 1998; Supplemental Fig. S1). Following accumulation of MP in virus factories, the infected cells accumulate the MP also along microtubules (Heinlein et al., 1998). The accumulation of MP in virus factories and on microtubules in cells behind the leading front of infection is dispensable for virus movement (Heinlein et al., 1998; Boyko et al., 2000a). At these late infection stages, the virus factories may enable the virus to produce high virion titers (Laliberté and Sanfaçon, 2010; Tilsner et al., 2012), and the subsequent accumulation along microtubules may play a role in withdrawing MP from the cell-to-cell communication pathway (Curin et al., 2007) and in stockpiling MP prior to degradation (Padgett et al., 1996; Gillespie et al., 2002).The molecular mechanisms that guide the MP to the ER and subsequently to microtubules during infection are not known. The MP is a hydrophobic protein that behaves like a membrane-integral or tightly membrane-associated protein in differential fractionation experiments and contains two predicted transmembrane domains (Reichel and Beachy, 1998; Brill et al., 2000, 2004) involved in ER association (Fujiki et al., 2006). The association with microtubules depends on MP amino acids 1 to 213 required for MP function (Kahn et al., 1998; Boyko et al., 2000b,Boyko et al., 2000c, 2002; Kotlizky et al., 2001). Moreover, certain amino acid exchange mutations known to affect the function of MP in virus movement in a temperature-sensitive manner also affect the ability of MP to interact with microtubules (Boyko et al., 2007,Boyko et al., 2000b). Interestingly, these mutations cluster together in a short domain of 25 amino acids showing a structural similarity with the M-loop of tubulin involved in tubulin-tubulin interactions (Boyko et al., 2000b; Waigmann et al., 2007). Importantly, this M-loop similarity domain overlaps with the predicted transmembrane domain (Brill et al., 2000, 2004) thus suggesting that the association of MP with membranes or microtubules is an alternative event that may depend on specific posttranslational modifications or specific folds of MP. However, although the different subcellular localizations of MPs during the course of infection indicate directional transport of MP from the ER to microtubules and may indicate different folds and functions of the protein when associated with these different subcellular components, the mechanism that controls the subcellular localization and, thus, the fate and function of MP is not known.Here, we identify CELL-DIVISION-CYCLE protein48 (CDC48), named p97/VCP (Valosin-containing protein) in mammals and Cdc48p in yeast (Saccharomyces cerevisiae), as a cellular factor regulating MP subcellular accumulation patterns. CDC48 functions are well characterized in mammalian and yeast systems but remain poorly investigated in plants. Yeast and mammalian CDC48s are essential, conserved chaperones involved in diverse cellular processes by controlling protein fate through extraction of substrates from membranes or complexes (Tsai et al., 2002; Meusser et al., 2005; Römisch, 2005; Rumpf and Jentsch, 2006; Schrader et al., 2009; Eisele et al., 2010; Meyer et al., 2012; Yamanaka et al., 2012). We show that virus infection leads to the induction of Arabidopsis (Arabidopsis thaliana) CDC48 isoforms and demonstrate a function of CDC48 in ER maintenance upon ER stress conditions. We further demonstrate that CDC48 interacts with MP and that CDC48 activity is required for MP degradation. Interaction of CDC48 with MP depends on the MP N terminus, which is required for degradation of the protein, for PD localization and microtubule accumulation of MP, and for function of MP in cell-to-cell transport of the viral RNA. Overexpressed CDC48 shifts MP subcellular localization from ER inclusions to microtubules, suggesting that CDC48 extracts the MP from ER-associated inclusions, where it accumulates in midstages of infection, to the cytosol, where it accumulates along microtubules during late infection stages. Moreover, overexpression of active, but not inactive, CDC48 inhibits virus movement. Our data demonstrate that a CDC48-dependent pathway leading to the clearance of ER-associated protein inclusions exists in plants, that plant viral MPs are substrates for this pathway, and that this pathway determines viral protein fate during infection. We suggest that CDC48-mediated extraction of MP from the ER is part of a plant defense response to remove MP from the ER, the compartment the virus uses for replication and movement.  相似文献   

15.
16.
Primary plasmodesmata (PD) arise at cytokinesis when the new cell plate forms. During this process, fine strands of endoplasmic reticulum (ER) are laid down between enlarging Golgi-derived vesicles to form nascent PD, each pore containing a desmotubule, a membranous rod derived from the cortical ER. Little is known about the forces that model the ER during cell plate formation. Here, we show that members of the reticulon (RTNLB) family of ER-tubulating proteins in Arabidopsis (Arabidopsis thaliana) may play a role in the formation of the desmotubule. RTNLB3 and RTNLB6, two RTNLBs present in the PD proteome, are recruited to the cell plate at late telophase, when primary PD are formed, and remain associated with primary PD in the mature cell wall. Both RTNLBs showed significant colocalization at PD with the viral movement protein of Tobacco mosaic virus, while superresolution imaging (three-dimensional structured illumination microscopy) of primary PD revealed the central desmotubule to be labeled by RTNLB6. Fluorescence recovery after photobleaching studies showed that these RTNLBs are mobile at the edge of the developing cell plate, where new wall materials are being delivered, but significantly less mobile at its center, where PD are forming. A truncated RTNLB3, unable to constrict the ER, was not recruited to the cell plate at cytokinesis. We discuss the potential roles of RTNLBs in desmotubule formation.Plasmodesmata (PD), the small pores that connect higher plant cells, are complex structures of about 50 nm in diameter. Each PD pore is lined by the plasma membrane and contains an axial endoplasmic reticulum (ER)-derived structure known as the desmotubule (Overall and Blackman, 1996; Maule, 2008; Tilsner et al., 2011). The desmotubule is an enigmatic structure whose function has not been fully elucidated. The small spiraling space between the desmotubule and the plasma membrane, known as the cytoplasmic sleeve, is almost certainly a conduit for the movement of small molecules (Oparka et al., 1999). Some reports, however, suggest that the desmotubule may also function in cell-to-cell trafficking, providing an ER-derived pathway between cells along which macromolecules may diffuse (Cantrill et al., 1999). The desmotubule is one of the most tightly constricted membrane structures found in nature (Tilsner et al., 2011), but the forces that generate its intense curvature are not understood. In most PD, the desmotubule is a tightly furled tube of about 15 nm in diameter in which the membranes of the ER are in close contact along its length. The desmotubule may balloon out in the region of the middle lamella into a central cavity, but at the neck regions of the PD pore it is tightly constricted (Robinson-Beers and Evert, 1991; Ding et al., 1992; Glockmann and Kollmann, 1996; Overall and Blackman, 1996; Ehlers and Kollmann, 2001). Studies of PD using GFP targeted to the ER lumen (e.g. GFP-HDEL) have shown that GFP is excluded from the desmotubule due to the constriction of ER membranes in this structure (Oparka et al., 1999; Crawford and Zambryski, 2000; Martens et al., 2006; Guenoune-Gelbart et al., 2008). Therefore, lumenal GFP is unable to move between plant cells unless the membranes of the desmotubule become relaxed in some way. On the other hand, dyes and some proteins inserted into the ER membrane can apparently move through the desmotubule, either along the membrane or through the lumen, at least under some conditions (Grabski et al., 1993; Cantrill et al., 1999; Martens et al., 2006; Guenoune-Gelbart et al., 2008).Recently, a number of proteins have been described in mammalian, yeast, and plant systems that induce extreme membrane curvature. Among these are the RETICULONS (RTNs), integral membrane proteins that induce curvature of the ER to form tubules (Voeltz et al., 2006; Hu et al., 2008; Tolley et al., 2008, 2010; Sparkes et al., 2010). In animals, RTNs have been shown to be involved in a wide array of endomembrane-related processes, including intracellular transport and vesicle formation, and as RTNs can also influence axonal growth, they may have roles in neurodegenerative disorders such as Alzheimer’s disease (Yang and Strittmatter, 2007). Arabidopsis (Arabidopsis thaliana) has 21 RTN homologs, known as RTNLBs (Nziengui et al., 2007; Sparkes et al., 2010), considerably more than in yeast or mammals, but most have not been examined. RTNLBs contain two unusually long hydrophobic helices that form reentrant loops (Voeltz et al., 2006; Hu et al., 2008; Sparkes et al., 2010; Tolley et al., 2010). These are thought to induce membrane curvature by the molecular wedge principle (Hu et al., 2008; Shibata et al., 2009). When RTNLBs are overexpressed transiently in cells expressing GFP-HDEL, the ER becomes tightly constricted and GFP-HDEL is excluded from the lumen of the constricted ER tubules (Tolley et al., 2008, 2010), a situation similar to that which occurs in desmotubules (Oparka et al., 1999; Crawford and Zambryski, 2000; Martens et al., 2006). In vitro studies with isolated membranes have shown that the degree of tubulation is proportional to the number and spacing of RTNLB proteins in the membrane (Hu et al., 2008). For example, to constrict the ER membrane into a structure of 15 nm, the diameter of a desmotubule, would require RTNLBs to be inserted every 2 nm or less along the desmotubule axis (Hu et al., 2008), potentially making the desmotubule an extremely protein-rich structure (Tilney et al., 1991). Interestingly, a number of RTNLB proteins appear in the recently described PD proteome (Fernandez-Calvino et al., 2011), suggesting that RTNLBs are good candidates for proteins that model the cortical ER into desmotubules.Primary PD form at cytokinesis during the assembly of the cell plate (Hawes et al., 1981; Hepler, 1982). Of the numerous studies devoted to the structure of the cell plate, very few have examined the behavior of the ER during cytokinesis. During mitosis, elements of the ER are located in the spindle apparatus, separated from the cytoplasm (Hepler, 1980). Just prior to cytokinesis, there is a relative paucity of ER in the region destined to become the cell plate (Hepler, 1980; Hawes et al., 1981). The studies of Hawes et al. (1981) and Hepler (1982), exploiting heavy-metal impregnation of the ER, showed that during the formation of the new cell plate, strands of cortical ER are inserted across the developing wall, between the Golgi-derived vesicles that deposit wall materials. These ER strands become increasingly thinner during formation of the desmotubule, eventually excluding heavy metal stains from the ER lumen (Hepler, 1982). The center of the desmotubule often appears electron opaque in transmission electron microscopy images and has been referred to as the central rod (Overall and Blackman, 1996). This structure may consist of proteins that extend from the inner ER leaflets or may correspond to head groups of the membrane lipids themselves. In the fully formed primary PD, the desmotubule remains continuous with the cortical ER that runs close to the new cell wall (Hawes et al., 1981; Hepler, 1982; Oparka et al., 1994).Here, we show that two of the RTNLBs present in the PD proteome, RTNLB3 and RTNLB6, become localized to the cell plate during the formation of primary PD. These RTNLBs remain associated with the desmotubule in fully formed PD and are immobile, as evidenced by fluorescence recovery after photobleaching (FRAP) studies. A truncated version of RTNLB3, in which the second hydrophobic region was deleted (Sparkes et al., 2010), was not recruited to the cell plate at cytokinesis. We suggest that RTNLBs play an important role in the formation of primary PD and discuss mechanisms by which these proteins may model the ER into desmotubules.  相似文献   

17.
18.
The actin cytoskeleton is a major regulator of cell morphogenesis and responses to biotic and abiotic stimuli. The organization and activities of the cytoskeleton are choreographed by hundreds of accessory proteins. Many actin-binding proteins are thought to be stimulus-response regulators that bind to signaling phospholipids and change their activity upon lipid binding. Whether these proteins associate with and/or are regulated by signaling lipids in plant cells remains poorly understood. Heterodimeric capping protein (CP) is a conserved and ubiquitous regulator of actin dynamics. It binds to the barbed end of filaments with high affinity and modulates filament assembly and disassembly reactions in vitro. Direct interaction of CP with phospholipids, including phosphatidic acid, results in uncapping of filament ends in vitro. Live-cell imaging and reverse-genetic analyses of cp mutants in Arabidopsis (Arabidopsis thaliana) recently provided compelling support for a model in which CP activity is negatively regulated by phosphatidic acid in vivo. Here, we used complementary biochemical, subcellular fractionation, and immunofluorescence microscopy approaches to elucidate CP-membrane association. We found that CP is moderately abundant in Arabidopsis tissues and present in a microsomal membrane fraction. Sucrose density gradient separation and immunoblotting with known compartment markers were used to demonstrate that CP is enriched on membrane-bound organelles such as the endoplasmic reticulum and Golgi. This association could facilitate cross talk between the actin cytoskeleton and a wide spectrum of essential cellular functions such as organelle motility and signal transduction.The cellular levels of membrane-associated lipids undergo dynamic changes in response to developmental and environmental stimuli. Different species of phospholipids target specific proteins and this often affects the activity and/or subcellular localization of these lipid-binding proteins. One such membrane lipid, phosphatidic acid (PA), serves as a second messenger and regulates multiple developmental processes in plants, including seedling development, root hair growth and pattern formation, pollen tube growth, leaf senescence, and fruit ripening. PA levels also change during various stress responses, including high salinity and dehydration, pathogen attack, and cold tolerance (Testerink and Munnik, 2005, 2011; Wang, 2005; Li et al., 2009). In mammalian cells, PA is critical for vesicle trafficking events, such as vesicle budding from the Golgi apparatus, vesicle transport, exocytosis, endocytosis, and vesicle fusion (Liscovitch et al., 2000; Freyberg et al., 2003; Jenkins and Frohman, 2005).The actin cytoskeleton and a plethora of actin-binding proteins (ABPs) are well-known targets and transducers of lipid signaling (Drøbak et al., 2004; Saarikangas et al., 2010; Pleskot et al., 2013). For example, several ABPs have the ability to bind phosphoinositide lipids, such as phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2]. The severing or actin filament depolymerizing proteins such as villin, cofilin, and profilin are inhibited when bound to PtdIns(4,5)P2. One ABP appears to be strongly regulated by another phospholipid; human gelsolin binds to lysophosphatidic acid and its filament severing and barbed-end capping activities are inhibited by this biologically active lipid (Meerschaert et al., 1998). Gelsolin is not, however, regulated by PA (Meerschaert et al., 1998), nor are profilin (Lassing and Lindberg, 1985), α-actinin (Fraley et al., 2003), or chicken CapZ (Schafer et al., 1996).The heterodimeric capping protein (CP) from Arabidopsis (Arabidopsis thaliana) also binds to and its activity is inhibited by phospholipids, including both PtdIns(4,5)P2 and PA (Huang et al., 2003, 2006). PA and phospholipase D activity have been implicated in the actin-dependent tip growth of root hairs and pollen tubes (Ohashi et al., 2003; Potocký et al., 2003; Samaj et al., 2004; Monteiro et al., 2005a; Pleskot et al., 2010). Exogenous application of PA causes an elevation of actin filament levels in suspension cells, pollen, and Arabidopsis epidermal cells (Lee et al., 2003; Potocký et al., 2003; Huang et al., 2006; Li et al., 2012; Pleskot et al., 2013). Capping protein (CP) binds to the barbed end of actin filaments with high (nanomolar) affinity, dissociates quite slowly, and prevents the addition of actin subunits at this end (Huang et al., 2003, 2006; Kim et al., 2007). In the presence of phospholipids, AtCP is not able to bind to the barbed end of actin filaments (Huang et al., 2003, 2006). Furthermore, capped filament ends are uncapped by the addition of PA, allowing actin assembly from a pool of profilin-actin (Huang et al., 2006). Collectively, these data lead to a simple model whereby CP, working in concert with profilin-actin, serves to maintain tight regulation of actin assembly at filament barbed ends (Huang et al., 2006; Blanchoin et al., 2010; Henty-Ridilla et al., 2013; Pleskot et al., 2013). Furthermore, the availability of CP for filament ends can be modulated by fluxes in signaling lipids. Genetic evidence for this model was recently obtained by analyzing the dynamic behavior of actin filament ends in living Arabidopsis epidermal cells after treatment with exogenous PA (Li et al., 2012). Specifically, changes in the architecture of cortical actin arrays and dynamics of individual actin filaments that are induced by PA treatment were found to be attenuated in cp mutant cells (Li et al., 2012; Pleskot et al., 2013).Structural characterization of chicken CapZ demonstrates that the α- and β-subunits of the heterodimer form a compact structure resembling a mushroom with pseudo-two-fold rotational symmetry (Yamashita et al., 2003). Actin- and phospholipid-binding sites are conserved on the C-terminal regions, sometimes referred to as tentacles, which comprise amphipathic α-helices (Cooper and Sept, 2008; Pleskot et al., 2012). Coarse-grained molecular dynamics (CG-MD) simulations recently revealed the mechanism of chicken and AtCP association with membranes (Pleskot et al., 2012). AtCP interacts specifically with lipid bilayers through interactions between PA and the amphipathic helix of the α-subunit tentacle. Extensive polar contacts between lipid headgroups and basic residues on CP (including K278, which is unique to plant CP), as well as partial embedding of nonpolar groups into the lipid bilayer, are observed (Pleskot et al., 2012). Moreover, a glutathione S-transferase fusion protein containing the C-terminal 38 amino acids from capping protein α subunit (CPA) is sufficient to bind PA-containing liposomes in vitro (Pleskot et al., 2012). Collectively, these findings lead us to predict that AtCP will behave like a membrane-associated protein in plant cells.Additional evidence from animal and microbial cells supports the association of CP with biological membranes. In Acanthamoeba castellanii, CP is localized primarily to the hyaline ectoplasm in a region of the cytoplasm just under the plasma membrane that contains a high concentration of actin filaments (Cooper et al., 1984). Localization of CP with regions rich in actin filaments and with membranes was supported by subcellular fractionation experiments, in which CP was associated with a crude membrane fraction that included plasma membrane (Cooper et al., 1984). Further evidence demonstrates that CP localizes to cortical actin patches at sites of new cell wall growth in budding yeast (Saccharomyces cerevisiae), including the site of bud emergence. By contrast, CP did not colocalize with actin cables in S. cerevisiae (Amatruda and Cooper, 1992). CP may localize to these sites by direct interactions with membrane lipids, through binding the ends of actin filaments, or by association with another protein different from actin. In support of this hypothesis, GFP-CP fusion proteins demonstrate that sites of actin assembling in living cells contain both CP and the actin-related protein2/3 (Arp2/3) complex, and CP is located in two types of structures: (1) motile regions of the cell periphery, which reflect movement of the edge of the lamella during extension and ruffling; and (2) dynamic spots within the lamella (Schafer et al., 1998). CP has been colocalized to the F-actin patches in fission yeast (Schizosaccharomyces pombe; Kovar et al., 2005), which promotes Arp2/3-dependent nucleation and branching and limits the extent of filament elongation (Akin and Mullins, 2008). These findings lend additional support for a model whereby CP cooperates with the Arp2/3 complex to regulate actin dynamics (Nakano and Mabuchi, 2006). Activities and localization of other plant ABPs are linked to membranes. Membrane association has been linked to the assembly status of the ARP2/3 complex, an actin filament nucleator, in Arabidopsis (Kotchoni et al., 2009). SPIKE1 (SPK1), a Rho of plants (Rop)-guanine nucleotide exchange factor (GEF) and peripheral membrane protein, maintains the homeostasis of the early secretory pathway and signal integration during morphogenesis through specialized domains in the endoplasmic reticulum (ER; Zhang et al., 2010). Furthermore, Nck-associated protein1 (NAP1), a component of the suppressor of cAMP receptor/WASP-family verprolin homology protein (SCAR/WAVE) complex, strongly associates with membranes and is particularly enriched in ER membranes (Zhang et al., 2013a). Finally, a superfamily of plant ABPs, called NETWORKED proteins, was recently discovered; these link the actin cytoskeleton to various cellular membranes (Deeks et al., 2012; Hawkins et al., 2014; Wang et al., 2014).In this work, we demonstrate that CP is a membrane-associated protein in Arabidopsis. To our knowledge, this is the first direct evidence for CP-membrane association in plants. This interaction likely targets CP to cellular compartments such as the ER and Golgi. This unique location may allow CP to remodel the actin cytoskeleton in the vicinity of endomembrane compartments and/or to respond rapidly to fluxes in signaling lipids.  相似文献   

19.
20.
This study dealt with the visualization of the sieve element (SE) cytoskeleton and its involvement in electrical responses to local cold shocks, exemplifying the role of the cytoskeleton in Ca2+-triggered signal cascades in SEs. High-affinity fluorescent phalloidin as well as immunocytochemistry using anti-actin antibodies demonstrated a fully developed parietal actin meshwork in SEs. The involvement of the cytoskeleton in electrical responses and forisome conformation changes as indicators of Ca2+ influx was investigated by the application of cold shocks in the presence of diverse actin disruptors (latrunculin A and cytochalasin D). Under control conditions, cold shocks elicited a graded initial voltage transient, ΔV1, reduced by external La3+ in keeping with the involvement of Ca2+ channels, and a second voltage transient, ΔV2. Cytochalasin D had no effect on ΔV1, while ΔV1 was significantly reduced with 500 nm latrunculin A. Forisome dispersion was triggered by cold shocks of 4°C or greater, which was indicative of an all-or-none behavior. Forisome dispersion was suppressed by incubation with latrunculin A. In conclusion, the cytoskeleton controls cold shock-induced Ca2+ influx into SEs, leading to forisome dispersion and sieve plate occlusion in fava bean (Vicia faba).It has been argued for a long time that sieve elements (SEs) are devoid of a cytoskeleton (Parthasarathy and Pesacreta, 1980; Thorsch and Esau, 1981; Evert, 1990), but more recent biochemical and cytological studies favor the opposite view. Actin as well as profilin were detected in phloem exudates of various monocot and dicot species (Schobert et al., 1998, 2000), while immunocytochemical tests showed the presence of actin and tubulin in phloem exudates of pumpkin (Cucurbita maxima; Kulikova and Puryaseva, 2002). Proteome analyses gave further credence to the occurrence of microfilaments in SEs in castor bean (Ricinus communis; profilin; Barnes et al., 2004), pumpkin (actin; Walz et al., 2004), canola (Brassica napus; actin, profilin1 and profilin2, actin-depolymerizing factor4; Giavalisco et al., 2006), and rice (Oryza sativa; actin1, actin-depolymerizing factor2, actin depolymerizing-factor3, and actin-depolymerizing factor6; Aki et al., 2008). Moreover, cytological evidence suggests residues of a cytoskeleton in SEs; fluorescent immunolabeling identified an actin/myosin system at the sieve plates (Chaffey and Barlow, 2002).Theoretical considerations also call for the presence of a cytoskeleton in SEs. Turnover and addressing of macromolecules (Fisher et al., 1992; Leineweber et al., 2000) requires a local distribution network in SEs. This function was attributed to an endoplasmic reticulum (ER) continuous to the ER strands running through pore plasmodesma units (Blackman et al., 1998) into the companion cells. Although such a mechanism is essentially conceivable, an interaction between the ER and cytoskeleton would provide a more conventional mode of intracellular distribution (Hepler et al., 1990; Boevink et al., 1998; Ueda et al., 2010; Yokota et al., 2011; Chen et al., 2012). Moreover, macromolecular trafficking through pore plasmodesma units (Lucas et al., 2001) was proposed to be executed by actin and myosin (Oparka, 2004), implying the presence of a cytoskeleton in SEs. Despite the massive circumstantial evidence, however, a complete cytoskeleton network and its spatial distribution in SEs have not been visually documented thus far.The existence of an SE cytoskeleton would raise questions regarding its task(s) in this highly specialized cell type. In other plant cells, the cytoskeleton was proposed to be engaged, among others, in ion channel operation and intracellular signaling (Trewavas and Malho, 1997; Mazars et al., 1997, and refs. therein; Thuleau et al., 1998; Örvar et al., 2000; Sangwan et al., 2001; Drøbak et al., 2004; Davies and Stankovic, 2006), as in animal cells (Janmey, 1998; Lange and Gartzke, 2006). For instance, K+ fluxes are regulated by actin dynamics (Hwang et al., 1997; Liu and Luan, 1998; Chérel, 2004), while Ca2+ influx into the cytoplasm appears to be mediated by voltage-dependent Ca2+-permeable channels associated with microtubules (Mazars et al., 1997; Thion et al., 1998) or by mechanosensitive channels possibly associated with microfilaments (Wang et al., 2004; Zhang et al., 2007).Both types of Ca2+-permeable channels probably reside in the SE plasma membrane (Knoblauch et al., 2001; Hafke et al., 2007, 2009; Furch et al., 2009), where they are likely involved in Ca2+-dependent systemic signaling (Furch et al., 2009; Hafke et al., 2009; van Bel et al., 2011; Hafke and van Bel, 2013). These channels are also putative initiators of Ca2+-induced signal transduction in SEs, leading to sieve-plate occlusion in response to local cold shocks (Thorpe et al., 2010). In fava bean (Vicia faba), Ca2+-dependent sieve tube occlusion by dispersion of special phloem-specific proteins (P-proteins) known as forisomes has been studied intensely (Knoblauch et al., 2001; Furch et al., 2007, 2009; Thorpe et al., 2010). Thus, apart from its distributive tasks, a cytoskeleton may be of major importance for intracellular signaling cascades in the highly specialized, sparsely equipped SEs.Our objective was to investigate the existence and spatial distribution of an SE cytoskeleton and its engagement in local signaling through Ca2+ influx brought about by cold shocks. This study dealt with the visualization of cytoskeletal components in intact sieve tubes using microinjection of fluorescent phalloidin and immunocytochemistry. Confocal laser-scanning micrography (CLSM) and transmission electron microscopy unequivocally showed a parietally located cylindrical actin meshwork. We demonstrated the engagement of the network in local cold shock-induced electrical responses and its association with Ca2+ influx, since we found effects of the Ca2+ channel blocker La3+ and of the cytoskeleton disruptor latrunculin A (LatA) on electrical signatures triggered by cold shocks and, by consequence, on forisome conformation changes.  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