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 perception and response of pollen tubes to the female guidance signals are crucial for directional pollen tube growth inside female tissues, which leads to successful reproduction. In pursuing the mechanisms underlying this biological process, we identified the Arabidopsis (Arabidopsis thaliana) abnormal pollen tube guidance1 (aptg1) mutant, whose pollen tubes showed compromised micropylar guidance. In addition to its male defect, the aptg1 mutant showed embryo lethality. APTG1 encodes a putative mannosyltransferase homolog to human PHOSPHATIDYLINOSITOL GLYCAN ANCHOR BIOSYNTHESIS B and yeast (Saccharomyces cerevisiae) GLYCOSYLPHOSPHATIDYLINOSITOL10 (GPI10), both of which are involved in the biosynthesis of GPI anchors. We found that APTG1 was expressed in most plant tissues, including mature pollen, pollen tubes, mature embryo sacs, and developing embryos. By fluorescence colabeling, we showed that APTG1 was localized in the endoplasmic reticulum, where GPI anchors are synthesized. Disruption of APTG1 affected the localization of COBRA-LIKE10, a GPI-anchored protein important for pollen tube growth and guidance. The results shown here demonstrate that APTG1 is involved in both vegetative and reproductive development in Arabidopsis, likely through processing and proper targeting of GPI-anchored proteins.Double fertilization is the biological basis for seed propagation and plant reproduction in angiosperms. Pollen tubes grow through maternal tissue to deliver the immobile sperm cells into the female gametophyte (embryo sac). During this process, pollen tube guidance into the micropyle is a critical step and is precisely regulated (Dresselhaus and Franklin-Tong, 2013). Female guidance signals are generated by both sporophytic and gametophytic tissues and operate at different stages during pollen tube growth. The sporophytic signal directs the growth of pollen tubes in the stigma, style, and transmitting tract. The signal that induces pollen tubes to turn to the funiculus and grow into the micropyle is termed gametophytic guidance (Shimizu and Okada, 2000; Higashiyama et al., 2003). Extensive cellular and genetic studies have demonstrated that female gametophytes play key roles in the micropylar guidance of pollen tubes (Kasahara et al., 2005; Márton et al., 2005; Chen et al., 2007; Alandete-Saez et al., 2008; Okuda et al., 2009; Kessler and Grossniklaus, 2011; Takeuchi and Higashiyama, 2011). The molecular natures of such guidance signals have been gradually revealed in recent years (i.e. small peptides secreted by the female gametophyte, egg apparatus, or synergid cells; Márton et al., 2005; Jones-Rhoades et al., 2007; Okuda et al., 2009).Pollen tubes need to perceive the female guidance signals at the cell surface to initiate intracellular responses for directional growth. However, the mechanisms of pollen tube perception are still obscure. A few male factors involved in signal perception during pollen tube growth into ovules have been identified. For example, the Arabidopsis (Arabidopsis thaliana) sperm cell-specific protein HAPLESS2/GENERATIVE CELL-SPECIFIC1 was necessary for pollen tubes to target the micropyle (von Besser et al., 2006). Arabidopsis CATION/PROTON EXCHANGER21 (CHX21) and CHX23 encode K+ transporters in growing pollen tubes. Pollen grains of the chx21 chx23 double mutant germinated and extended a normal tube in the transmitting tract, but their targeting of the funiculus failed (Lu et al., 2011). Arabidopsis POLLEN DEFECTIVE IN GUIDANCE1 (POD1) was expressed in pollen grains, pollen tubes, and synergid cells. The pod1 pollen tubes showed defective micropylar guidance (Li et al., 2011). The tip of the pollen tube has been hypothesized to be the site of cue perception for micropyle-directed growth. The Arabidopsis Rab GTPase RABA4D was localized at the tips of growing pollen tubes. Pollen tubes with defective RABA4D had severely reduced growth rates and ovule targeting (Szumlanski and Nielsen, 2009). Recently, two receptor-like kinases at the apical plasma membrane (PM) of growing pollen tubes, LOST IN POLLEN TUBE GUIDANCE1 (LIP1) and LIP2, were demonstrated to guide pollen tubes to the micropyle by perceiving the AtLURE1 signal from synergid cells (Liu et al., 2013).Glycosylphosphatidylinositol (GPI) anchoring provides a strategy for targeting proteins to the outer layer of the PMs in eukaryotic cells. GPI anchors are synthesized inside the endoplasmic reticulum (ER) and are attached to proteins by posttranslational modifications in the ER. After processing, GPI-anchored proteins (GPI-APs) are transported to the cell surface following an unknown trafficking route and anchored at the cell surface (Maeda and Kinoshita, 2011). GPI-APs play very important roles in plant reproductive development (Gillmor et al., 2005; Ching et al., 2006; DeBono et al., 2009). An Arabidopsis putative GPI-AP, LORELEI, functioned in pollen tube reception of female signals, double fertilization, and early seed development (Capron et al., 2008; Tsukamoto et al., 2010). Arabidopsis COBRA-LIKE10 (COBL10), a GPI-AP, regulates the polar deposition of wall components in pollen tubes growing inside female tissues and is critical for micropylar guidance (Li et al., 2013). The conserved backbone of GPI anchors in eukaryotes is ethanolamine phosphate-6-Man-α-1,2-Man-α-1,6-Man-α-1,4-glucosamine-α-1,6-myoinositol phospholipid. During the biosynthesis of GPI anchors, monosaccharides, fatty acids, and phosphoethanolamines are sequentially added onto phosphatidylinositol. This process involves at least 16 enzymes and cofactors in mammals, including PHOSPHATIDYLINOSITOL GLYCAN ANCHOR BIOSYNTHESIS (PIG) A, B, C, F, G, H, L, M, N, O, P, Q, V, W, X, and Y (Maeda and Kinoshita, 2011). The core structure of the GPI anchor contains three Man residues donated by the substrate dolichol-phosphate-Man. GPI mannosyltransferases were required for adding the three Man residues of the GPI anchor in the ER lumen (Maeda and Kinoshita, 2011). Arabidopsis PEANUT1 (PNT1) is a homolog of the mammalian GPI mannosyltransferase PIG-M, involved in the addition of the first Man during the biosynthesis of the GPI anchor. The pnt1 mutant showed the defect of pollen viability and embryo development (Gillmor et al., 2005). PIG-B of human and GPI10 of yeast (Saccharomyces cerevisiae) encode GLYCOSYLPHOSPHATIDYLINOSITOL MANNOSYLTRANSFERASE3, involved in the addition of the third Man during the biosynthesis of the GPI anchor (Takahashi et al., 1996; Sütterlin et al., 1998). Mutation of PIG-B and GPI10 resulted in the accumulation of the GPI intermediate Man2-glucosamine-(acyl) phosphatidylinositol and led to cell death in yeast.In this study, we identified the ER-localized ABNORMAL POLLEN TUBE GUIDANCE1 (APTG1), an Arabidopsis homolog of PIG-B and GPI10. Pollen tubes of the aptg1 mutant showed compromised directional growth to the micropyle and lost the apical PM localization of COBL10. Besides the male defect, the mutant showed embryo lethality. In addition, reducing the expression of APTG1 resulted in defective seedling growth, indicating that APTG1 plays important roles in both reproductive and vegetative development.  相似文献   

2.
Plasmodesmata (Pd) are membranous channels that serve as a major conduit for cell-to-cell communication in plants. The Pd-associated β-1,3-glucanase (BG_pap) and CALLOSE BINDING PROTEIN1 (PDCB1) were identified as key regulators of Pd conductivity. Both are predicted glycosylphosphatidylinositol-anchored proteins (GPI-APs) carrying a conserved GPI modification signal. However, the subcellular targeting mechanism of these proteins is unknown, particularly in the context of other GPI-APs not associated with Pd. Here, we conducted a comparative analysis of the subcellular targeting of the two Pd-resident and two unrelated non-Pd GPI-APs in Arabidopsis (Arabidopsis thaliana). We show that GPI modification is necessary and sufficient for delivering both BG_pap and PDCB1 to Pd. Moreover, the GPI modification signal from both Pd- and non-Pd GPI-APs is able to target a reporter protein to Pd, likely to plasma membrane microdomains enriched at Pd. As such, the GPI modification serves as a primary Pd sorting signal in plant cells. Interestingly, the ectodomain, a region that carries the functional domain in GPI-APs, in Pd-resident proteins further enhances Pd accumulation. However, in non-Pd GPI-APs, the ectodomain overrides the Pd targeting function of the GPI signal and determines a specific GPI-dependent non-Pd localization of these proteins at the plasma membrane and cell wall. Domain-swap analysis showed that the non-Pd localization is also dominant over the Pd-enhancing function mediated by a Pd ectodomain. In conclusion, our results indicate that segregation between Pd- and non-Pd GPI-APs occurs prior to Pd targeting, providing, to our knowledge, the first evidence of the mechanism of GPI-AP sorting in plants.Plant cells are interconnected with cross-wall membranous channels called plasmodesmata (Pd). Recent studies have shown that the region of the plasma membrane (PM) lining the Pd channel is a specialized membrane microdomain whose lipid and protein composition differs from the rest of the PM (Tilsner et al., 2011, 2016; Bayer et al., 2014; González-Solís et al., 2014; Grison et al., 2015). In a similar manner, the cell wall domain surrounding the Pd channel is specialized and, unlike the rest of the cell wall, is devoid of cellulose, rich in pectin, and contains callose (an insoluble β-1,3-glucan; Zavaliev et al., 2011; Knox and Benitez-Alfonso, 2014). In response to physiological signals, callose can be transiently deposited and degraded at Pd, which provides a mechanism for controlling the Pd aperture in diverse developmental and stress-related processes (Zavaliev et al., 2011). Control of Pd functioning is mediated by proteins that are specifically targeted to Pd. Plasmodesmal proteins localized to the PM domain of Pd can be integral transmembrane proteins, such as Pd-localized proteins (Thomas et al., 2008), the receptor kinase ARABIDOPSIS CRINKLY4 (Stahl et al., 2013), and callose synthases (Vatén et al., 2011). Alternatively, Pd proteins can associate with the membrane through a lipid modification like myristoylation (e.g. remorins; Raffaele et al., 2009) or be attached by a glycosylphosphatidylinositol (GPI) anchor (e.g. Pd-associated β-1,3-glucanases [BG_pap]; Levy et al., 2007; Rinne et al., 2011; Benitez-Alfonso et al., 2013), Pd-associated callose-binding proteins (PDCBs; Simpson et al., 2009), and LYSIN MOTIF DOMAIN-CONTAINING PROTEIN2 (LYM2; Faulkner et al., 2013).Among the known Pd proteins involved in Pd-specific callose degradation is BG_pap, a cell wall enzyme carrying a glycosyl hydrolase family 17 (GH17) module as its functional domain (Levy et al., 2007). Another group of proteins controlling callose dynamics at Pd are PDCBs that harbor a callose-binding domain termed carbohydrate-binding module 43 (CBM43), implicated in stabilizing callose at Pd (Simpson et al., 2009). Some β-1,3-glucanases may combine the two callose-modifying activities by harboring both GH17 and CBM43 functional domains, and several such proteins were shown to localize to Pd (Rinne et al., 2011; Benitez-Alfonso et al., 2013; Gaudioso-Pedraza and Benitez-Alfonso, 2014). A distinct feature of BG_pap and PDCBs is that both are predicted glycosylphosphatidylinositol-anchored proteins (GPI-APs). The GPI anchor is a form of posttranslational modification common to many cell surface proteins in all eukaryotes. GPI-APs are covalently attached to the outer leaflet of the PM through the GPI anchor. The basic structure of the anchor consists of ethanolamine phosphate, followed by a glycan chain of three Man residues and glucosamine, followed by phosphatidylinositol lipid moiety (EtNP-6Manα1-2Manα1-6Manα1-4GlcNα1-6myoinositol-1-P-lipid; Ferguson et al., 2009). All predicted GPI-APs carry an N-terminal secretion signal peptide (SP) similar to other secreted proteins. Distinctly, GPI-APs also carry a structurally conserved 25- to 30-residue C-terminal GPI attachment signal, which typically begins with a small amino acid (e.g. Ala, Asn, Asp, Cys, Gly, or Ser) termed omega, followed by a spacer region of five to 10 polar residues, and ending with a transmembrane segment of 15 to 20 hydrophobic residues (Ferguson et al., 2009). The entire region between the N-terminal and the C-terminal signals of a GPI-AP is termed the ectodomain and carries the protein’s functional domain(s). The GPI modification process takes place in the lumenal face of the endoplasmic reticulum (ER) in a cotranslational manner. Upon translocation into the ER, a GPI-AP is stabilized in the ER membrane by its C-terminal signal, which is concurrently cleaved after the omega amino acid, and a preassembled GPI anchor is covalently attached to the C terminus of the omega amino acid. After attachment to a protein, the GPI anchor undergoes a series of modifications (remodeling), both at the glucan chain and at the lipid moiety. Such remodeling is crucial for the sorting of GPI-APs in the secretory pathway and the subsequent lateral heterogeneity at the PM (Kinoshita, 2015). In particular, the addition of saturated fatty acid chains to the lipid moiety of the anchor leads to the enriched accumulation of GPI-APs in the PM microdomains, also termed lipid rafts (Muñiz and Zurzolo, 2014). In Arabidopsis (Arabidopsis thaliana), GPI modification has been predicted for 210 proteins of diverse functions at the PM or the cell wall or both (Borner et al., 2002). Despite extensive research on the GPI modification pathway and the function of GPI-APs in mammalian and yeast cells, such knowledge in plant systems is scarce. In particular, despite an emerging role of GPI-APs in the regulation of the cell wall domain of Pd, their subcellular targeting and compartmentalization mechanism have not been studied. In addition, it is not known how the targeting mechanism of Pd-resident GPI-APs is different from that of other classes of GPI-APs, which are not localized to Pd.In this study, we investigated the subcellular targeting mechanism of Pd-associated callose-modifying GPI-APs, BG_pap and PDCB1, and compared it with that of two unrelated non-Pd GPI-APs, ARABINOGALACTAN PROTEIN4 (AGP4) and LIPID TRANSFER PROTEIN1 (LTPG1). Using sequential fluorescent labeling of protein domains, we found that the C-terminal GPI modification signal present in both Pd- and non-Pd GPI-APs can function as a primary signal in targeting proteins to the Pd-enriched PM domain. Moreover, we show that while the GPI signal is sufficient for Pd targeting, the ectodomains in BG_pap and PDCB1 further enhance their accumulation at Pd. In contrast, the ectodomains in non-Pd GPI-APs mediate exclusion of the proteins from the Pd-enriched targeting pathway. The Pd exclusion effect was found to be dominant over the Pd-targeting function of the GPI signal and the Pd-enhancing function of the Pd ectodomain, and it possibly occurs prior to PM localization. Our findings thus uncover a novel Pd-targeting signal and provide, to our knowledge, the first evidence of the cellular mechanism that regulates the sorting of GPI-APs in plants.  相似文献   

3.
The P6 protein of Cauliflower mosaic virus (CaMV) is responsible for the formation of inclusion bodies (IBs), which are the sites for viral gene expression, replication, and virion assembly. Moreover, recent evidence indicates that ectopically expressed P6 inclusion-like bodies (I-LBs) move in association with actin microfilaments. Because CaMV virions accumulate preferentially in P6 IBs, we hypothesized that P6 IBs have a role in delivering CaMV virions to the plasmodesmata. We have determined that the P6 protein interacts with a C2 calcium-dependent membrane-targeting protein (designated Arabidopsis [Arabidopsis thaliana] Soybean Response to Cold [AtSRC2.2]) in a yeast (Saccharomyces cerevisiae) two-hybrid screen and have confirmed this interaction through coimmunoprecipitation and colocalization assays in the CaMV host Nicotiana benthamiana. An AtSRC2.2 protein fused to red fluorescent protein (RFP) was localized to the plasma membrane and specifically associated with plasmodesmata. The AtSRC2.2-RFP fusion also colocalized with two proteins previously shown to associate with plasmodesmata: the host protein Plasmodesmata-Localized Protein1 (PDLP1) and the CaMV movement protein (MP). Because P6 I-LBs colocalized with AtSRC2.2 and the P6 protein had previously been shown to interact with CaMV MP, we investigated whether P6 I-LBs might also be associated with plasmodesmata. We examined the colocalization of P6-RFP I-LBs with PDLP1-green fluorescent protein (GFP) and aniline blue (a stain for callose normally observed at plasmodesmata) and found that P6-RFP I-LBs were associated with each of these markers. Furthermore, P6-RFP coimmunoprecipitated with PDLP1-GFP. Our evidence that a portion of P6-GFP I-LBs associate with AtSRC2.2 and PDLP1 at plasmodesmata supports a model in which P6 IBs function to transfer CaMV virions directly to MP at the plasmodesmata.Through the years, numerous studies have focused on the characterization of viral replication sites within the cell, as well as how plant virus movement proteins (MPs) modify the plasmodesmata to facilitate cell-to-cell movement (for review, see Benitez-Alfonso et al., 2010; Laliberté and Sanfaçon, 2010; Niehl and Heinlein, 2011; Ueki and Citovsky, 2011; Verchot, 2012). It is accepted that plant virus replication is associated with host membranes, and at some point, the viral genomic nucleic acid must be transferred from the site of replication in the cell to the plasmodesmata. This step could involve transport from a distant site within the cell, or alternatively, it may be that replication is coupled with transport at the entrance of the plasmodesmata (Tilsner et al., 2013). However, even with the latter model, there is ample evidence that the viral proteins necessary for replication or cell-to-cell movement utilize intracellular trafficking pathways within the cell to become positioned at the plasmodesma. These pathways may involve microfilaments, microtubules, or specific endomembranes that participate in macromolecular transport pathways, or combinations of these elements (Harries et al., 2010; Schoelz et al., 2011; Patarroyo et al., 2012; Peña and Heinlein, 2012; Tilsner and Oparka 2012; Liu and Nelson, 2013).The P6 protein of Cauliflower mosaic virus (CaMV) is one viral protein that had not been considered to play a role in viral movement until recently. P6 is the most abundant protein component of the amorphous, electron-dense inclusion bodies (IBs) present during virus infection (Odell and Howell, 1980; Shockey et al., 1980). Ectopic expression of P6 in Nicotiana benthamiana leaves resulted in the formation of inclusion-like bodies (I-LBs) that were capable of intracellular movement along actin microfilaments. Furthermore, treatment of Nicotiana edwardsonii leaves with latrunculin B abolished the formation of CaMV local lesions, suggesting that intact microfilaments are required for CaMV infection (Harries et al., 2009a). A subsequent paper showed that P6 physically interacts with Chloroplast Unusual Positioning1 (CHUP1), a plant protein localized to the chloroplast outer membrane that contributes to movement of chloroplasts on microfilaments in response to changes in light intensity (Oikawa et al., 2003, 2008; Angel et al., 2013). The implication was that P6 might hijack CHUP1 to facilitate movement of the P6 IBs on microfilaments. Silencing of CHUP1 in N. edwardsonii, a host for CaMV, slowed the rate of local lesion formation, suggesting that CHUP1 contributes to intracellular movement of CaMV (Angel et al., 2013).In addition to its role in intracellular trafficking, the P6 protein has been shown to have at least four other distinct functions in the viral infection cycle. P6-containing IBs induced during virus infection are likely virion factories, as they are the primary site for CaMV protein synthesis, genome replication, and assembly of virions (Hohn and Fütterer, 1997). Second, P6 interacts with host ribosomes to facilitate reinitiation of translation of genes on the polycistronic 35S viral RNA, a process called translational transactivation (Bonneville et al., 1989; Park et al., 2001; Ryabova et al., 2002). The translational transactivator region of P6 (Fig. 1) defines the essential sequences required for translational transactivation (DeTapia et al., 1993). Third, P6 is an important pathogenicity determinant. P6 functions as an avirulence determinant in some solanaceous and cruciferous species (Daubert et al., 1984; Schoelz et al., 1986; Hapiak et al., 2008) and is a chlorosis symptom determinant in susceptible hosts (Daubert et al., 1984; Baughman et al., 1988; Goldberg et al., 1991; Cecchini et al., 1997). Finally, P6 has the capacity to compromise host defenses, as it is a suppressor of RNA silencing and cell death (Love et al., 2007; Haas et al., 2008), and it modulates signaling by salicylic acid, jasmonic acid, ethylene, and auxin (Geri et al., 2004; Love et al., 2012; Laird et al., 2013). Domain D1 of P6 has been shown to be necessary but not sufficient for suppression of silencing and salicylic acid-mediated defenses (Laird et al., 2013).Open in a separate windowFigure 1.CaMV and host constructs used for confocal microscopy or coimmunoprecipitation (co-IP). A, Structure of CaMV P6 and Arabidopsis (Arabidopsis thaliana) Soybean Response to Cold (AtSRC2.2) proteins. The functions of P6 domains D1 to D4 tested for interaction with AtSRC2.2 are indicated by the shaded boxes. The Mini TAV is the minimal region for the translational transactivation function. The NLSa sequence corresponds to the nuclear localization signal of influenza virus. The NLS sequence corresponds to the nuclear localization signal of human ribosomal protein L22. B, Structure of P6 (Angel et al., 2013), AtSRC2.2, PDLP (Thomas et al., 2008), and CaMV MP fusions developed for confocal microscopy and/or co-IP. aa, Amino acid.Because P6-containing IBs are the site for virion accumulation and they are capable of movement, they may be responsible for delivering virions to the CaMV MP located at the plasmodesmata (for review, see Schoelz et al., 2011). The vast majority of CaMV virions accumulate in association with P6-containing IBs. Furthermore, P6 physically interacts with the CaMV capsid and MP, as well as the two proteins necessary for aphid transmission, P2 and P3 (Himmelbach et al., 1996; Ryabova et al., 2002; Hapiak et al., 2008; Lutz et al., 2012). Recent studies have indicated that P6 IBs serve as a reservoir for virions, in which the virions may be rapidly transferred to P2 electron-lucent IBs for acquisition by aphids (Bak et al., 2013). It stands to reason that P6 IBs may also serve as a reservoir for CaMV virions to be transferred to the CaMV MP in the plasmodesmata.CaMV virions move from cell to cell through plasmodesmata modified into tubules through the function of its MP (Perbal et al., 1993; Kasteel et al., 1996). However, studies have suggested that CaMV virions do not appear to directly interact with the MP. Instead, the MP interacts with the CaMV P3 protein (also known as the virion-associated protein [VAP]), which forms a trimeric structure that is anchored into the virions (Leclerc et al., 1998; Leclerc et al., 2001). Electron microscopy studies have indicated that MP and VAP colocalize with virions only at the entrance to or within the plasmodesmata, and it has been suggested that the VAP/virion complex travels to the plasmodesmata independently from the MP (Stavolone et al., 2005). Consequently, there is a need for a second CaMV protein such as P6 to fulfill the role of delivery of virions to the plasmodesmata (Schoelz et al., 2011).Additional studies have shown that the CaMV MP is incorporated into vesicles and is trafficked on the endomembrane system to reach the plasmodesma (Carluccio et al., 2014). These authors suggest that the CaMV MP is recycled in a vesicular transport pathway between plasmodesmata and early endosome compartments. The CaMV MP interacts with µA-Adaptin (Carluccio et al., 2014) and Movement Protein-Interacting7 (Huang et al., 2001), two proteins shown to have a role in vesicular trafficking. Once the MP arrives at plasmodesmata, it interacts with the Plasmodesmata-Localized Protein (PDLP) proteins, which comprise a family of eight proteins associated with plasmodesmata (Amari et al., 2010). In addition to its interaction with CaMV MP, PDLP1 interacts with the 2B protein of Grapevine fan leaf virus (GFLV) at the base of tubules formed by the 2B protein. Furthermore, an Arabidopsis transfer DNA (T-DNA) mutant line in which three PDLP genes had been knocked out (pdlp1-pdlp2-pdlp3) responded to GFLV and CaMV inoculation with a delayed infection (Amari et al., 2010). This has led to the suggestion that the PDLPs might act as receptors for the MPs of the tubule-forming viruses such as GFLV and CaMV (Amari et al., 2010, 2011).To better understand the function of the P6 protein during CaMV intracellular movement, we have utilized a yeast (Saccharomyces cerevisiae) two-hybrid assay to identify host proteins that interact with CaMV P6. We show that P6 physically interacts with a C2-calcium-dependent protein (designated AtSRC2.2). AtSRC2.2 is a membrane-bound protein that is capable of forming punctate spots associated with plasmodesmata. The localization of AtSRC2.2 with plasmodesmata led to an analysis of interactions between P6 I-LBs, AtSRC2.2, PDLP1, and the CaMV MP and also revealed that a portion of P6 I-LBs are found adjacent to plasmodesmata. These results provide further evidence for a model in which P6 IBs are capable of delivery of virions to plasmodesmata for their transit to other host cells.  相似文献   

4.
The predominant structure of the hemicellulose xyloglucan (XyG) found in the cell walls of dicots is a fucogalactoXyG with an XXXG core motif, whereas in the Poaceae (grasses and cereals), the structure of XyG is less xylosylated (XXGGn core motif) and lacks fucosyl residues. However, specialized tissues of rice (Oryza sativa) also contain fucogalactoXyG. Orthologous genes of the fucogalactoXyG biosynthetic machinery of Arabidopsis (Arabidopsis thaliana) are present in the rice genome. Expression of these rice genes, including fucosyl-, galactosyl-, and acetyltransferases, in the corresponding Arabidopsis mutants confirmed their activity and substrate specificity, indicating that plants in the Poaceae family have the ability to synthesize fucogalactoXyG in vivo. The data presented here provide support for a functional conservation of XyG structure in higher plants.The plant cell wall protects and structurally supports plant cells. The wall consists of a variety of polymers, including polysaccharides, the polyphenol lignin, and glycoproteins. One of the major polysaccharides present in the primary walls (i.e. walls of growing cells) in dicots is xyloglucan (XyG), which consists of a β-1,4-glucan backbone with xylosyl substituents. XyG binds noncovalently to cellulose microfibrils and thereby, is thought to act as a spacer molecule, hindering cellulose microfibrils to aggregate (Carpita and Gibeaut, 1993; Pauly et al., 1999a; Bootten et al., 2004; Cosgrove, 2005; Hayashi and Kaida, 2011; Park and Cosgrove, 2012).The side-chain substitutions on XyG can be structurally diverse depending on plant species, tissue type, and developmental stage of the tissue (Pauly et al., 2001; Hoffman et al., 2005; Peña et al., 2008; Hsieh and Harris, 2009, 2012; Lampugnani et al., 2013; Schultink et al., 2014). A one-letter code nomenclature has been established to specify the XyG side-chain substitutions (Fry et al., 1993; Tuomivaara et al., 20145). According to this nomenclature, an unsubstituted glucosyl residue is indicated by a G, whereas a glucosyl residue substituted with a xylosyl moiety is shown as an X. In most dicots, the xylosyl residue can be further substituted with a galactosyl residue (L), which in turn, can be further decorated with a fucosyl residue (F) and/or an acetyl group (F/L). In some species, the xylosyl residue can be substituted with an arabinosyl moiety (S), and the backbone glucosyl residue can be O-acetylated (G; Jia et al., 2003; Hoffman et al., 2005).Numerous genes have been identified in Arabidopsis (Arabidopsis thaliana) that are involved in fucogalactoXyG biosynthesis (Fig. 1; Pauly et al., 2013; Schultink et al., 2014). The glucan backbone is thought to be synthesized by cellulose synthase-like C (CSLC) family proteins, such as AtCSLC4, as shown by in vitro activity data (Cocuron et al., 2007). Several xylosyltransferases (XXTs) from glycosyl transferase family 34 (GT34) are thought to be responsible for XyG xylosylation. Five of these XXTs in Arabidopsis seem to have XXT activity on XyG in vitro (Faik et al., 2002; Zabotina et al., 2008; Vuttipongchaikij et al., 2012; Mansoori et al., 2015). MURUS3 (MUR3) represents a galactosyltransferase that transfers galactosyl moieties specifically to xylosyl residues adjacent to an unsubstituted glucosyl residue on an XXXG unit, converting it to XXLG, whereas Xyloglucan L-side chain galactosyl Transferase2 (XLT2) was identified as another galactosyltransferase transferring a galactosyl moiety specifically to the second xylosyl residue, resulting in XLXG (Madson et al., 2003; Jensen et al., 2012). Both MUR3 and XLT2 belong to GT47 (Li et al., 2004). MUR2/FUCOSYLTRANSFERASE1 (FUT1) from GT37 was found to harbor fucosyltransferase activity, transferring Fuc from GDP-Fuc to a galactosyl residue adjacent to the unsubstituted glucosyl residue (i.e. onto XXLG but not onto XLXG; Perrin et al., 1999; Vanzin et al., 2002). O-acetylation of the galactosyl residue is mediated by Altered Xyloglucan4 (AXY4) and AXY4L, both of which belong to the Trichome Birefringence-Like (TBL) protein family (Bischoff et al., 2010; Gille et al., 2011; Gille and Pauly, 2012).Open in a separate windowFigure 1.Schematic structures of two types of XyGs and known biosynthetic proteins in Arabidopsis (Hsieh and Harris, 2009; Pauly et al., 2013). The corresponding one-letter code for XyG is shown below the pictograms (Fry et al., 1993; Tuomivaara et al., 2015).XyG found throughout land plants exhibits structural diversity with respect to side-chain substitution patterns (Schultink et al., 2014). Most dicots, such as Arabidopsis, and the noncommelinoid monocots possess a fucogalactoXyG of the XXXG-type XyG structure as shown in Figure 1. However, plant species in the Solanaceae and Poaceae as well as the moss Physcomitrella patens contain a different XyG structure with a reduced level of xylosylation, resulting in an XXGGn core motif (York et al., 1996; Kato et al., 2004; Gibeaut et al., 2005; Jia et al., 2005; Peña et al., 2008; Hsieh and Harris, 2009). In addition, the glucan backbone can be O-acetylated in plants of Solanaceae and Poaceae families (Gibeaut et al., 2005; Jia et al., 2005). XyG from Solanaceae with an XXGG core motif can be further arabinosylated and/or galactosylated (Jia et al., 2005). No XyGs with an XXGGn motif backbone have been reported to be fucosylated.The function of structural diversity of XyG substitutions, such as fucosylation and/or altered xylosylation pattern, remains enigmatic. Removing the terminal fucosyl or acetyl moieties in the corresponding Arabidopsis mutants does not lead to any change in plant growth and development (Vanzin et al., 2002; Gille et al., 2011). However, removing galactosyl residues as well as fucosyl and acetyl moieties in the Arabidopsis xlt2 mur3.1 double mutant results in a dwarfed plant (Jensen et al., 2012; Kong et al., 2015). Replacing the galactosyl moiety with an arabinofuranosyl residue by, for example, expressing a tomato (Solanum lycopersicum) arabinosyltransferase in the Arabidopsis xlt2 mur3.1 mutant rescues the growth phenotype and restores wall biomechanics, indicating that galactosylation and arabinosylation in XyG have an equivalent function (Schultink et al., 2013). Recently, fucosylated XyG structures were found in the pollen tubes of tobacco (Nicotiana alata) and tomato, indicating that fucogalactoXyG is likely also present in other Solanaceae plants, albeit restricted to specific tissues (Lampugnani et al., 2013; Dardelle et al., 2015). Although there is circumstantial evidence that fucogalactoXyG is present in cell suspension cultures of rice (Oryza sativa) and cell suspension cultures of fescue (Festuca arundinaceae; McDougall and Fry, 1994; Peña et al., 2008), fucogalactoXyG has not been found in any physiologically relevant plant tissues of members of the Poaceae (Kato et al., 1982; Watanabe et al., 1984; Gibeaut et al., 2005; Hsieh and Harris, 2009; Brennan and Harris, 2011). Here, we provide chemical and genetic evidence that fucogalactoXyG is, indeed, present in plant tissues of a grass (rice) and prove that the rice genome harbors the genes that could be part of the synthetic machinery necessary to produce fucogalactoXyG.  相似文献   

5.
Protein amino (N) termini are prone to modifications and are major determinants of protein stability in bacteria, eukaryotes, and perhaps also in chloroplasts. Most chloroplast proteins undergo N-terminal maturation, but this is poorly understood due to insufficient experimental information. Consequently, N termini of mature chloroplast proteins cannot be accurately predicted. This motivated an extensive characterization of chloroplast protein N termini in Arabidopsis (Arabidopsis thaliana) using terminal amine isotopic labeling of substrates and mass spectrometry, generating nearly 14,000 tandem mass spectrometry spectra matching to protein N termini. Many nucleus-encoded plastid proteins accumulated with two or three different N termini; we evaluated the significance of these different proteoforms. Alanine, valine, threonine (often in N-α-acetylated form), and serine were by far the most observed N-terminal residues, even after normalization for their frequency in the plastid proteome, while other residues were absent or highly underrepresented. Plastid-encoded proteins showed a comparable distribution of N-terminal residues, but with a higher frequency of methionine. Infrequent residues (e.g. isoleucine, arginine, cysteine, proline, aspartate, and glutamate) were observed for several abundant proteins (e.g. heat shock proteins 70 and 90, Rubisco large subunit, and ferredoxin-glutamate synthase), likely reflecting functional regulation through their N termini. In contrast, the thylakoid lumenal proteome showed a wide diversity of N-terminal residues, including those typically associated with instability (aspartate, glutamate, leucine, and phenylalanine). We propose that, after cleavage of the chloroplast transit peptide by stromal processing peptidase, additional processing by unidentified peptidases occurs to avoid unstable or otherwise unfavorable N-terminal residues. The possibility of a chloroplast N-end rule is discussed.Following synthesis, most proteins undergo various N-terminal (Nt) protein modifications, including removal of the Nt Met and signal peptide, N-terminal α-acetylation (NAA), ubiquitination, and acylations. These Nt modifications play an important role in the regulation of cellular functions. The N terminus of proteins has also been shown to be a major determinant of protein stability in bacteria (Varshavsky, 2011), eukaryotes (Graciet et al., 2009), mitochondria, and perhaps in plastids/chloroplasts (Apel et al., 2010; Nishimura et al., 2013; van Wijk, 2015). The role of the N terminus in protein stability is conceptualized in the N-end rule, which states that certain amino acids, when exposed at the N terminus of a protein, act as triggers for degradation (Bachmair et al., 1986; Dougan et al., 2012; Tasaki et al., 2012; Gibbs et al., 2014).Most of the approximately 3,000 plastid proteins are nucleus encoded (n-encoded) and are targeted to the plastid through an Nt chloroplast transit peptide (cTP). After import, the cTP is cleaved by the stromal processing peptidase (SPP; Richter and Lamppa, 1998; Trösch and Jarvis, 2011). The consensus site of cTP cleavage by SPP is only loosely defined, and the rules, mechanisms, and enzymes for possible subsequent processing, stabilization, and other posttranslational modifications (PTMs) are not well characterized (for discussion, see van Wijk, 2015). The exact N terminus is unknown for many chloroplast proteins and cannot be accurately predicted, because SPP specificity is not sufficiently understood (Emanuelsson et al., 2000; Zybailov et al., 2008) and probably also because additional Nt processing occurs for many chloroplast proteins (Fig. 1A). The approximately 85 plastid-encoded (p-encoded) proteins typically first undergo cotranslational Nt deformylation, followed by N-terminal Met excision (NME; Giglione et al., 2009; Fig. 1B); both these PTMs are required for normal plastid/chloroplast development and protein stability (Dirk et al., 2001, 2002; Giglione et al., 2003; Meinnel et al., 2006). Both n-encoded and p-encoded proteins can undergo NAA inside the plastid (Zybailov et al., 2008; Fig. 1). Postulated functions of NAA in eukaryotes include the mediation of protein location, assembly, and stability (Jones and O’Connor, 2011; Starheim et al., 2012; Hoshiyasu et al., 2013; Xu et al., 2015), thereby affecting a variety of processes, including drought tolerance in Arabidopsis (Arabidopsis thaliana; Linster et al., 2015).Open in a separate windowFigure 1.Conceptual illustration of Nt maturation of n-encoded and p-encoded proteins. Ac, Acetylated; MAP, Met amino peptidase; NAT, N-acetyltransferase; N-term, N-terminal; PDF, peptide deformylase. A, Nt maturation of n-encoded plastid proteins including removal of cTP by SPP and potential subsequent Nt modifications. B, Nt maturation of p-encoded proteins. *, The removal depends on the penultimate residue, generally following the N-terminal Met Excision (NME) rule; **, N-terminal acetylation typically occurs only for selected residues; “Results”).Typical proteomics work flows generally yield only partial coverage of protein sequences, and it is often difficult to know which peptides represent the true N termini (Nti) or C termini. Systematic identification of Nti or C termini requires specific labeling and enrichment strategies, such as combined fractional diagonal chromatography, developed by Gevaert and colleagues (Staes et al., 2011), and terminal amine isotopic labeling of substrates (TAILS), developed by the group of Overall (Kleifeld et al., 2011; Lange and Overall, 2013). These strategies allow the identification of different Nt proteoforms and were recently also applied to plants (Tsiatsiani et al., 2013; Carrie et al., 2015; Kohler et al., 2015; Zhang et al., 2015) and diatoms (Huesgen et al., 2013). We previously reported on Nti of chloroplast proteins based on tandem mass spectrometry (MS/MS) analysis, but because no Nt enrichment/labeling technique was used, only those that underwent NAA could be considered bona fide Nti (Zybailov et al., 2008). Nt Edman degradation sequencing was systematically carried out for thylakoid lumen proteins (Peltier et al., 2000, 2002) but not for stromal proteins or chloroplast membrane proteins with their Nti exposed to the stroma. The Nti of thylakoid lumen proteins are mostly generated by lumenal peptidases (Hsu et al., 2011; Midorikawa et al., 2014), and the thylakoid lumen contains a different set of peptidases than the stroma; hence, rules for Nt maturation and stability are likely different than those for stroma-exposed proteins.The objective of this study was to systematically determine the Nti of stroma-exposed chloroplast proteins of Arabidopsis (the N-terminome) and to provide a baseline for understanding Nt protein maturation and protein stability in the chloroplast stroma. To that end, we applied the TAILS technique and determined the Nti of approximately 250 chloroplast proteins by mass spectrometry (MS). We observed that many n-encoded plastid proteins accumulated with two or even three different Nt residues, in many cases both with and without NAA. The extent of accumulation of different Nt proteoforms is surprising and will be discussed. The p-encoded proteins generally showed very similar Nt residues as compared with the n-encoded proteins, with the exception of Met. Our data show that small, apolar, or hydroxylated residues (Ala, Val, Ser, and Thr) are the most frequent Nt residues of stromal proteins, whereas other residues are strictly avoided or are only present for very specific proteins likely to aid in their function. Chloroplast protein degradation products were also detected, with enrichment for peptides generated by cleavage between Arg and Thr residues. We present testable hypotheses for understanding Nt processing and maturation, stability, and a possible N-end rule in chloroplast stroma.  相似文献   

6.
7.
8.
9.
10.
11.
Xyloglucan (XyG) is the dominant hemicellulose present in the primary cell walls of dicotyledonous plants. Unlike Arabidopsis (Arabidopsis thaliana) XyG, which contains galactosyl and fucosyl substituents, tomato (Solanum lycopersicum) XyG contains arabinofuranosyl residues. To investigate the biological function of these differing substituents, we used a functional complementation approach. Candidate glycosyltransferases were identified from tomato by using comparative genomics with known XyG galactosyltransferase genes from Arabidopsis. These candidate genes were expressed in an Arabidopsis mutant lacking XyG galactosylation, and two of them resulted in the production of arabinosylated XyG, a structure not previously found in this plant species. These genes may therefore encode XyG arabinofuranosyltransferases. Moreover, the addition of arabinofuranosyl residues to the XyG of this Arabidopsis mutant rescued a growth and cell wall biomechanics phenotype, demonstrating that the function of XyG in plant growth, development, and mechanics has considerable flexibility in terms of the specific residues in the side chains. These experiments also highlight the potential of reengineering the sugar substituents on plant wall polysaccharides without compromising growth or viability.The cell wall of higher plants represents a composite material consisting of various polymers including cellulose, hemicellulose, lignin, pectin, and glycoproteins (Somerville et al., 2004). The quantity and fine structure of each of these components varies based on the tissue type and plant species (Pauly and Keegstra, 2008). One of the major components of the dicot primary wall (the wall of growing cells) is the hemicellulose xyloglucan (XyG), whose structure and biosynthesis are relatively well described (Zabotina, 2012). The glycan backbone of XyG consists of β-1,4-linked glucosyl residues, which are substituted with a regular pattern of xylosyl residues that can be further decorated with a diverse array of carbohydrate and noncarbohydrate substituents. A one-letter code nomenclature has been developed to specify the substituents of a particular backbone glucosyl residue (Fry et al., 1993). An unsubstituted Glc residue is depicted as G while a Glc substituted with a xylosyl residue is depicted as X. Further substitution of the Xyl with a β-galactosyl or α-arabinofuranosyl-residue is abbreviated as L or S, respectively (Fig. 1). In addition, an L side chain may contain an α-fucosyl moiety on the Gal (abbreviated F) or an acetyl group (underlined L). More than 10 additional side chain structures have been identified in various plant species (Hantus et al., 1997; Jia et al., 2003; Ray et al., 2004; Peña et al., 2008, 2012).Open in a separate windowFigure 1.Xyloglucan motifs present in the walls of Arabidopsis (XXXG, XXLG, and XLXG) and tomato (XSGG) with nomenclature indicated below the structure. AtMUR3 and AtXLT2 are galactosyltransferases required to produce XXLG and XLXG, respectively. Glc, d-Glucopyranose; Xyl, d-xylopyranose; Gal, d-galactopyranose; Ara, l-arabinofuranose.The analysis of XyG structure is facilitated by the availability of a XyG-specific endoglucanase (XEG) that can release XyG oligosaccharides from plant cell wall preparations (Pauly et al., 1999). When XyG is enzymatically released from the walls of the plant model species Arabidopsis (Arabidopsis thaliana), the oligosaccharides XXXG, XXLG, XXLG, XXFG, XXFG, XLFG, and XLFG are observed (Scheller and Ulvskov, 2010), the structures of several of which are shown in Figure 1. Arabinosylated side chains (S) have not been observed in Arabidopsis walls but are abundant in Solanaceous species such as tomato (Solanum lycopersicum). Tomato XyG consists primarily of the subunits LSGG, XSGG, LXGG, LLGG, and XXGG (Jia et al., 2003, 2005). Unlike the “XXXG” type motif in Arabidopsis, XyG in tomato is less xylosylated, with a repeating “XXGG” type motif. In addition, the glucan backbone of tomato XyG is O-acetylated, a modification that in Arabidopsis is only observed on the side chain galactosyl moiety. The functional significance and genetic basis of these structural differences are not understood.Numerous genes, mainly identified from Arabidopsis, are known to be involved in the biosynthesis of XyG (Pauly et al., 2013). These include a glucan synthase, a member of the Cellulose Synthase-Like C gene family (Cocuron et al., 2007), several XyG xylosyltransferases (XXTs; Faik et al., 2002; Cavalier and Keegstra, 2006), the galactosyltransferases MURUS3 (MUR3; Madson et al., 2003) and XyG Galactose Transferase at Position 2 (XLT2; Jensen et al., 2012), a XyG-specific galacturonosyltransferase (XUT1; Peña et al., 2012), the fucosyltransferase MUR2 (Perrin et al., 1999; Vanzin et al., 2002), and the XyG O-acetyltransferases Altered XyG4 (AXY4) and AXY4-Like (AXY4L; Gille et al., 2011). The Cellulose Synthase-Like C from nasturtium (Tropaeolum majus; Cocuron et al., 2007), where XyG is produced as a seed storage polymer, and MUR3 from eucalyptus (Eucalyptus grandis; Lopes et al., 2010) have both been investigated and appear to have specificities similar to their Arabidopsis orthologs. Glycosyltransferases (GTs) with novel specificities required for the diversity of XyG substitution found in various non-Arabidopsis species, including a XyG arabinosyltransferase, have not been identified to date.XyG figures prominently in many models of the plant cell wall, where it is thought to cross link cellulose microfibrils and have a mechanistic role in cell elongation (Somerville et al., 2004; Hayashi and Kaida, 2011). However, the recent discovery that an Arabidopsis xxt1 xxt2 double mutant lacks detectable XyG but only has relatively minor growth phenotypes questions the structural significance of this polysaccharide (Cavalier et al., 2008; Park and Cosgrove, 2012b). XyG substitution has been shown to influence polymer solubility and binding affinity in vitro, with the enzymatic removal of side chains leading to decreased polymer solubility (Sims et al., 1998; Lima et al., 2004). However, mutants deficient for MUR2, XLT2, or AXY4 show minor, if any, growth phenotypes under laboratory conditions (Vanzin et al., 2002; Gille et al., 2011; Jensen et al., 2012). A point mutant in MUR3 (mur3.1) shows minor growth phenotypes (Madson et al., 2003), while a transfer DNA mutant in MUR3 shows impaired growth and altered Golgi structure (Tamura et al., 2005). The difference between these alleles was attributed to a role of the MUR3 protein interacting with actin to help organize Golgi structure independent from its function as a GT (Tamura et al., 2005). XyG oligosaccharides, which can be liberated by endogenous or exogenous glycosyl hydrolases, have been suggested to have a role in signaling (Aldington et al., 1991), though the biological significance of this is unclear and a specific pathway has not been identified. A complimentary approach to using mutant lines lacking certain XyG substitutions to investigate the function of XyG substitution would be to introduce exogenous side chain structures by the expression of XyG biosynthetic genes from other species. This functional complementation approach requires the identification of the genes responsible for exogenous substitution patterns.The MUR3, XLT2, and XUT1 genes are in the same subclade of the inverting GT family 47 (Li et al., 2004). These three transferases all add β-glycosyl groups to the O2-position of a xylosyl group on XyG but differ in donor specificity, with MUR3 and XLT2 utilizing UDP-Gal and XUT1 utilizing UDP-GalA. The diversity of donor substrate specificity present in this subclade of GTs suggests that similar enzymes may represent good candidate genes for unidentified XyG GTs responsible for XyG side chain diversity in other species.Here, we report the identification of several tomato genes involved in XyG biosynthesis. Candidate genes were constitutively expressed in the Arabidopsis mur3.1 xlt2 double mutant, which contains mostly nonsubstituted XyG, for functional characterization. Two putative XyG arabinofuranosyltransferases were identified, and the evolutionary history of these genes was investigated using phylogenetics. The expression of these genes rescued the growth and petiole extensibility phenotypes of the mutant, demonstrating partial functional redundancy of XyG galactosylation and arabinosylation.  相似文献   

12.
13.
14.
15.
N-Glycan processing is one of the most important cellular protein modifications in plants and as such is essential for plant development and defense mechanisms. The accuracy of Golgi-located processing steps is governed by the strict intra-Golgi localization of sequentially acting glycosidases and glycosyltransferases. Their differential distribution goes hand in hand with the compartmentalization of the Golgi stack into cis-, medial-, and trans-cisternae, which separate early from late processing steps. The mechanisms that direct differential enzyme concentration are still unknown, but the formation of multienzyme complexes is considered a feasible Golgi protein localization strategy. In this study, we used two-photon excitation-Förster resonance energy transfer-fluorescence lifetime imaging microscopy to determine the interaction of N-glycan processing enzymes with differential intra-Golgi locations. Following the coexpression of fluorescent protein-tagged amino-terminal Golgi-targeting sequences (cytoplasmic-transmembrane-stem [CTS] region) of enzyme pairs in leaves of tobacco (Nicotiana spp.), we observed that all tested cis- and medial-Golgi enzymes, namely Arabidopsis (Arabidopsis thaliana) Golgi α-mannosidase I, Nicotiana tabacum β1,2-N-acetylglucosaminyltransferase I, Arabidopsis Golgi α-mannosidase II (GMII), and Arabidopsis β1,2-xylosyltransferase, form homodimers and heterodimers, whereas among the late-acting enzymes Arabidopsis β1,3-galactosyltransferase1 (GALT1), Arabidopsis α1,4-fucosyltransferase, and Rattus norvegicus α2,6-sialyltransferase (a nonplant Golgi marker), only GALT1 and medial-Golgi GMII were found to form a heterodimer. Furthermore, the efficiency of energy transfer indicating the formation of interactions decreased considerably in a cis-to-trans fashion. The comparative fluorescence lifetime imaging of several full-length cis- and medial-Golgi enzymes and their respective catalytic domain-deleted CTS clones further suggested that the formation of protein-protein interactions can occur through their amino-terminal CTS region.The Golgi apparatus is a multifaceted, multitasking organelle that is pivotal to the life of the cell. Protein and lipid modifications, sorting of molecules, as well as the biosynthesis of cell wall polysaccharides all take place in the small stacks of flattened cisternae that make up the Golgi bodies, constituting the Golgi apparatus in plant cells. Among the various posttranslational modification reactions on proteins, the biosynthesis and processing of protein-bound N-linked oligosaccharides (N-glycans) is the most common. N-Glycans play a crucial role in protein folding, endoplasmic reticulum (ER) quality control (Liu and Howell, 2010), biotic (Saijo, 2010) and abiotic (Koiwa et al., 2003; Kang et al., 2008) stress responses, and are considered essential for the physicochemical properties and biological functions of glycoproteins. Consequently, the slightest alterations during N-glycan processing can drastically affect a protein’s folding, stability, and biological activity. Golgi-mediated N-glycan processing steps are catalyzed by numerous glycosidases and glycosyltransferases that follow a nonuniform subcompartment-specific distribution pattern along the cis-to-trans axis of the Golgi stack in the order in which they function in the processing pathway (Fig. 1A; Schoberer and Strasser, 2011). The subcompartmentalization of the Golgi stack into cis-, medial-, and trans-cisternae creates a polar biochemical and morphological gradient in the cis-to-trans direction, which allows a functional specialization of the Golgi. Enzymes catalyzing early processing steps concentrate in the cis-half of the Golgi stack, whereas enzymes acting later in the pathway peak in the trans-half.Open in a separate windowFigure 1.Overview of fluorescent protein fusion constructs used for FRET-FLIM. A, Schematic representation of the N-glycan processing pathway in the plant Golgi apparatus and the enzymes involved (for enzyme names, see CTS regions and the full-length sequences (where available) were C-terminally fused to GFP and/or mRFP, respectively, by insertion into binary plant expression vectors. Details on plasmid construction can be found in “Materials and Methods.” C, Cytoplasmic tail; CD, catalytic domain; S, luminal stem region; T, transmembrane domain.All Golgi-resident plant N-glycan processing enzymes are typical so-called type II membrane proteins with an N-terminal region comprising a short cytoplasmic tail, a single transmembrane domain, and a luminal stem region, together called the cytoplasmic-transmembrane-stem (CTS) region, which orients the C-terminal catalytic domain into the Golgi lumen (Fig. 1B). The CTS region not only contains the information necessary for enzyme targeting to the Golgi but also directs the nonuniform, overlapping distribution of glycosidases and glycosyltransferases across the distinct Golgi cisternae (or subcompartments; Saint-Jore-Dupas et al., 2006; Schoberer et al., 2009, 2010; Schoberer and Strasser, 2011). The signals or mechanisms that drive the subcompartment-specific concentration of this important class of Golgi enzymes are still widely unknown. It is also intriguing how N-glycan processing enzymes remain specifically concentrated within Golgi membranes even in the presence of a continuous bidirectional flow of membrane and proteins into and out of the Golgi. Moreover, most of the processing enzymes are highly dynamic themselves, as they continuously cycle between the Golgi and the ER.One possible mechanism responsible for intra-Golgi concentration (as described for mammalian Golgi) is the self-assembly of Golgi-resident glycosylation enzymes into complexes, which are excluded from forward transport to downstream compartments as described in the “protein aggregation model” (Machamer, 1991) or the “kin recognition model” (Nilsson et al., 1993). In fact, there is compelling in vitro and in vivo evidence on the formation of oligomers by mammalian and yeast glycosyltransferases, respectively, from each major glycosylation category, namely, for proteoglycans, glycoproteins, and glycolipids (Schachter, 1986; Nilsson et al., 1994, 1996; McCormick et al., 2000; Giraudo et al., 2001; Pinhal et al., 2001; Qian et al., 2001; Stolz and Munro, 2002; Young, 2004; Hassinen et al., 2011; Ferrari et al., 2012). Earlier studies mainly examined complex formation through in vitro coimmunoprecipitation (co-IP) assays following complete disruption of the cell, whereas in recent years, the advent of fluorescent protein technology and that of laser-based microscopy techniques utilizing bimolecular fluorescence complementation or Förster resonance energy transfer (FRET) have proven invaluable for the observation of protein-protein interactions in vivo and in real time.There is emerging evidence from co-IP and bimolecular fluorescence complementation experiments for the formation of glycosyltransferase complexes involved in various aspects of plant cell wall biosynthesis, such as the biosynthesis of xylan (Zeng et al., 2010), homogalacturonan (Atmodjo et al., 2011), pectic arabinan (Harholt et al., 2012), and xyloglycan (Chou et al., 2012). To date, it is not known whether enzymes involved in N-glycan processing assemble into similar complexes. Protein-protein interaction studies on mammalian N-glycan processing enzymes have indicated the presence of homomeric and heteromeric enzyme complexes that potentially lead to their retention in Golgi membranes and enhance their activity (Nilsson et al., 1996; Rivinoja et al., 2009; Hassinen et al., 2010, 2011).To address the question of whether the same principle applies to enzymes from the plant N-glycan processing pathway, we have tested the properties of several glycosidases and glycosyltransferases with distinct intra-Golgi locations to form protein-protein interactions when expressed transiently in tobacco (Nicotiana spp.) leaves. To observe interactions in living cells and in real time, we employed time-resolved FRET-fluorescence lifetime imaging (FLIM), which is based on energy transfer from a fluorophore in an excited state (i.e. GFP serving as the donor) to another fluorophore (i.e. monomeric red fluorescent protein [mRFP] serving as the acceptor) within a 1- to 10-nm distance. In the FRET-FLIM approach, the information gained using steady-state FRET between interacting proteins is considerably improved by monitoring the excited-state lifetime of the donor, where its quenching is evidence for a direct physical interaction. Although technically demanding, FRET-FLIM is superior to other intensity-based techniques, as it is largely independent from fluorophore concentrations and is also free of interference from spectral cross talk. Two-photon excitation (2P)-FRET-FLIM analysis (Stubbs et al., 2005; Osterrieder et al., 2009; Sparkes et al., 2010) provides several advantages over the single-photon method, including reduced cellular cytotoxicity of the excitation light and reduced photobleaching of the fluorophore. Greater sensitivity of the setup is achieved through reduced sensitivity of the excitation light (greater than 900 nm) by the photomultiplier tube as a detector. The combination of the sensitive advanced imaging technique of time-correlated single-photon counting and laser scanning techniques results in the FLIM technique. The FRET-FLIM approach has successfully been used to study protein-protein interactions in a variety of animal and plant cells at the organelle level (Bhat et al., 2005; Stubbs et al., 2005; Adjobo-Hermans et al., 2006; Aker et al., 2007; Osterrieder et al., 2009; Sparkes et al., 2010; Crosby et al., 2011; Berendzen et al., 2012).Here, we present evidence that (1) N-glycan processing enzymes can form in vivo protein-protein interactions in the form of homodimers and heterodimers, (2) the N-terminal CTS regions of several dimerizing cis- and medial-Golgi enzymes participate in physical interactions, and (3) interaction mainly occurs among the CTS domains of cis- and medial-Golgi enzymes and decreases toward the trans-Golgi in a gradient-like fashion.  相似文献   

16.
Quantification of brassinosteroids is essential and extremely important to study the molecular mechanisms of their physiological roles in plant growth and development. Herein, we present a simple, material and cost-saving high-performance method for determining endogenous brassinosteroids (BRs) in model plants. This new method enables simultaneous enrichment of a wide range of bioactive BRs such as brassinolide, castasterone, teasterone, and typhasterol with ion exchange solid-phase extraction and high-sensitivity quantitation of these BRs based on isotope dilution combined with internal standard approach. For routine analysis, the consumption of plant materials was reduced to one-twentieth of previously reported and the overall process could be completed within 1 day compared with previous 3 to 4 days. The strategy was validated by profiling BRs in different ecotypes and mutants of rice (Oryza sativa) and Arabidopsis (Arabidopsis thaliana), and the BR distributions in different model plants tissues were determined with the new method. The method allows plant physiologists to monitor the dynamics and distributions of BRs with 1 gram fresh weight of model plant tissues, which will speed up the process for the molecular mechanism research of BRs with these model plants in future work.Brassinosteroids (BRs) have been considered as the sixth class of endogenous plant hormones with wide occurrence across the plant kingdom (Bajguz and Tretyn, 2003). BRs play a key role in a variety of physiological processes, such as cell elongation, vascular differentiation, reproductive development, photomorphogenesis, stress tolerance, and so on (Hayat, 2010). Recently, it was found that BR deficiency could increase grain yield in rice (Oryza sativa) by more than 30%, which showed a food security-enhancing potential and guided new green revolution in the future (Sakamoto et al., 2006; Wu et al., 2008). Since BRs were first isolated and identified from rape (Brassica napus) pollen in 1970s (Mitchell et al., 1970; Grove et al., 1979), the natural occurrence of more than 60 BRs in a large quantity of plant species has been reported (Hayat, 2010). To date, research on the occurrence of BRs in different plants, physiological properties, and their action modes has made much progress (Fujioka and Yokota, 2003; Symons et al., 2008; Kim and Wang, 2010; Tang et al., 2010; Tong and Chu, 2012). However, so far, only limited information was obtained to understand the molecular mechanism of the physiological role of BRs. For example, although the biosynthetic pathway of C28 BRs has been well established, the biosynthesis of C27 and C29 BRs remains unclear, and some intermediates on their biosynthetic pathways still need to be elucidated (Noguchi et al., 2000; Fujita et al., 2006). The plant physiology research of BRs is speeded up by employing BR mutants on biosynthesis and signaling pathways (Yamamuro et al., 2000; Hong et al., 2003; Kwon and Choe, 2005; Tanabe et al., 2005); however, a simple, high-sensitivity screening, detection, and quantification method for BR analysis is still a bottleneck technique for in-depth studying of the role of BRs during the life cycle of plants.In the past 20 years, most of the detection and identification processes of BRs could be described briefly as the following steps. The harvested plant materials were lyophilized and then ground to a fine powder, followed by the CH3OH/CHCl3 extraction. The concentrate was then partitioned with the CHCl3/H2O system three times. After that, the CHCl3 fraction was subjected to a silica gel column for BR enrichment, and the collected BRs-containing fraction was purified with Sephadex LH-20 column and Sep-Pak Plus C18 cartridge in sequence. At last, the collected fractions were purified with preparative HPLC and then derivatized for analysis with gas chromatography-mass spectrometry (MS) under selected ion monitoring mode (Hong et al., 2005; Nomura et al., 2005; Kim et al., 2006). So far, this protocol has been proven to be workable in most cases and provided a great quantity of valuable data for plant physiological research (Hwang et al., 2006; Lee et al., 2010). However, at least more than 20 g of plant materials were consumed for quantifying/identifying BRs in one plant sample without replicates (Hong et al., 2005; Bancos et al., 2006; Kim et al., 2006), and it is difficult to collect sufficient plant tissues for BR measurement in some rare model plant mutants. In addition, the method involved multiple tedious and labor-intensive steps, which might result in poor recovery and low sensitivity, especially for some labile BR intermediates. The traditional method took one person about 3 to 4 d to treat one batch of samples. Most of the BR measurement experiments were performed without biological replicates using traditional methods due to the disadvantages mentioned above, which discounted the reliability of the results. Therefore, a simple, rapid, and sensitive analysis method for BRs is in urgent need, along with the development of BR research.Recently, several efforts were made to improve the BR determination (Svatos et al., 2004; Huo et al., 2012). The consumption of plant material was reduced to 2 g after modifying the BRs with a new derivatization reagent for further ultra-performance liquid chromatography (UPLC)-multiple reaction monitoring (MRM)-MS detection. However, the purification process consisting of deproteinization and multiple solid-phase extraction (SPE) steps was still quite tedious and couldn’t guarantee covering the four most important bioactive BRs, including brassinolide (BL), castasterone (CS), teasterone (TE), and typhasterol (TY; Fig. 1). In our previous study (Xin et al., 2013), we reported a simple, convenient, and high-sensitivity method for detection of endogenous BRs from real plant materials based on the dual role of specific boronate affinity. Although it was the first time to measure multiple BRs in subgram plant materials and the time duration of the method decreased to one-third of that previously reported, the synthesis of boronate affinity-functionalized magnetic nanoparticles made the method difficult to follow in biological laboratories for routine analysis.Open in a separate windowFigure 1.Chemical structure of four major bioactive BRs.BRs are neutral steroid compounds with a common four-ring cholestane skeleton and hydroxyl groups on A ring and/or the side chain linked to D ring. Especially, the vicinal diol moieties on C22 and C23 sites of BL, CS, TY, and TE allow these bioactive BRs to be derivatized with ionizable reagents for MS response enhancement. Considering the unique physicochemical properties of BRs, we herein developed a simplified high-sensitivity analytical method based on mixed-mode anion exchange (MAX)-cation exchange (MCX) solid phase extraction (SPE) purification, vicinal diol derivatization combined with UPLC-MRM3-MS detection for quantification of BL, CS, TE, and TY in model plants (Fig. 2). The performance of the method was demonstrated by determination of BRs in different tissues of both wild-type and mutant Arabidopsis (Arabidopsis thaliana) and rice.Open in a separate windowFigure 2.Simplified high-sensitivity protocol for quantitative analysis of BRs. IS, Internal standards. [See online article for color version of this figure.]  相似文献   

17.
In plant cells, secretory and endocytic routes intersect at the trans-Golgi network (TGN)/early endosome (EE), where cargos are further sorted correctly and in a timely manner. Cargo sorting is essential for plant survival and therefore necessitates complex molecular machinery. Adaptor proteins (APs) play key roles in this process by recruiting coat proteins and selecting cargos for different vesicle carriers. The µ1 subunit of AP-1 in Arabidopsis (Arabidopsis thaliana) was recently identified at the TGN/EE and shown to be essential for cytokinesis. However, little was known about other cellular activities affected by mutations in AP-1 or the developmental consequences of such mutations. We report here that HAPLESS13 (HAP13), the Arabidopsis µ1 adaptin, is essential for protein sorting at the TGN/EE. Functional loss of HAP13 displayed pleiotropic developmental defects, some of which were suggestive of disrupted auxin signaling. Consistent with this, the asymmetric localization of PIN-FORMED2 (PIN2), an auxin transporter, was compromised in the mutant. In addition, cell morphogenesis was disrupted. We further demonstrate that HAP13 is critical for brefeldin A-sensitive but wortmannin-insensitive post-Golgi trafficking. Our results show that HAP13 is a key link in the sophisticated trafficking network in plant cells.Plant cells contain sophisticated endomembrane compartments, including the endoplasmic reticulum, the Golgi, the trans-Golgi network (TGN)/early endosome (EE), the prevacuolar compartments/multivesicular bodies (PVC/MVB), various types of vesicles, and the plasma membrane (PM; Ebine and Ueda, 2009; Richter et al., 2009). Intracellular protein sorting between the various locations in the endomembrane system occurs in both secretory and endocytic routes (Richter et al., 2009; De Marcos Lousa et al., 2012). Vesicles in the secretory route start at the endoplasmic reticulum, passing through the Golgi before reaching the TGN/EE, while vesicles in the endocytic route start from the PM before reaching the TGN/EE (Dhonukshe et al., 2007; Viotti et al., 2010). The TGN/EE in Arabidopsis (Arabidopsis thaliana) is an independent and highly dynamic organelle transiently associated with the Golgi (Dettmer et al., 2006; Lam et al., 2007; Viotti et al., 2010), distinct from the animal TGN. Once reaching the TGN/EE, proteins delivered by their vesicle carriers are subject to further sorting, being incorporated either into vesicles that pass through the PVC/MVB before reaching the vacuole for degradation or into vesicles that enter the secretory pathway for delivery to the PM (Ebine and Ueda, 2009; Richter et al., 2009). Therefore, the TGN/EE is a critical sorting compartment that lies at the intersection of the secretory and endocytic routes.Fine-tuned control of intracellular protein sorting at the TGN/EE is essential for plant development (Geldner et al., 2003; Dhonukshe et al., 2007, 2008; Richter et al., 2007; Kitakura et al., 2011; Wang et al., 2013). An auxin gradient is crucial for pattern formation in plants, whose dynamic maintenance requires the polar localization of auxin efflux carrier PINs through endocytic recycling (Geldner et al., 2003; Blilou et al., 2005; Paciorek et al., 2005; Abas et al., 2006; Jaillais et al., 2006; Dhonukshe et al., 2007; Kleine-Vehn et al., 2008). Receptor-like kinases (RLKs) have also been recognized as major cargos undergoing endocytic trafficking, which are either recycled back to the PM or sent for vacuolar degradation (Geldner and Robatzek, 2008; Irani and Russinova, 2009). RLKs are involved in most if not all developmental processes of plants (De Smet et al., 2009).Intracellular protein sorting relies on sorting signals within cargo proteins and on the molecular machinery that recognizes sorting signals (Boehm and Bonifacino, 2001; Robinson, 2004; Dhonukshe et al., 2007). Adaptor proteins (AP) play a key role (Boehm and Bonifacino, 2001; Robinson, 2004) in the recognition of sorting signals. APs are heterotetrameric protein complexes composed of two large subunits (β and γ/α/δ/ε), a small subunit (σ), and a medium subunit (µ) that is crucial for cargo selection (Boehm and Bonifacino, 2001). APs associate with the cytoplasmic side of secretory and endocytic vesicles, recruiting coat proteins and recognizing sorting signals within cargo proteins for their incorporation into vesicle carriers (Boehm and Bonifacino, 2001). Five APs have been identified so far, classified by their components, subcellular localization, and function (Boehm and Bonifacino, 2001; Robinson, 2004; Hirst et al., 2011). Of the five APs, AP-1 associates with the TGN or recycling endosomes (RE) in yeast and mammals (Huang et al., 2001; Robinson, 2004), mediating the sorting of cargo proteins to compartments of the endosomal-lysosomal system or to the basolateral PM of polarized epithelial cells (Gonzalez and Rodriguez-Boulan, 2009). Knockouts of AP-1 components in multicellular organisms resulted in embryonic lethality (Boehm and Bonifacino, 2001; Robinson, 2004).We show here that the recently identified Arabidopsis µ1 adaptin AP1M2 (Park et al., 2013; Teh et al., 2013) is a key component in the cellular machinery mediating intracellular protein sorting at the TGN/EE. AP1M2 was previously named HAPLESS13 (HAP13), whose mutant allele hap13 showed male gametophytic lethality (Johnson et al., 2004). In recent quests for AP-1 in plants, HAP13/AP1M2 was confirmed as the Arabidopsis µ1 adaptin based on its interaction with other components of the AP-1 complex as well as its localization at the TGN (Park et al., 2013; Teh et al., 2013). A novel mutant allele of HAP13/AP1M2, ap1m2-1, was found to be defective in the intracellular distribution of KNOLLE, leading to defective cytokinesis (Park et al., 2013; Teh et al., 2013). However, it was not clear whether HAP13/AP1M2 mediated other cellular activities and their developmental consequences. Using the same mutant allele, we found that functional loss of HAP13 (hap13-1/ap1m2-1) resulted in a full spectrum of growth defects, suggestive of compromised auxin signaling and of defective RLK signaling. Cell morphogenesis was also disturbed in hap13-1. Importantly, hap13-1 was insensitive to brefeldin A (BFA) washout, indicative of defects in guanine nucleotide exchange factors for ADP-ribosylation factor (ArfGEF)-mediated post-Golgi trafficking. Furthermore, HAP13/AP1M2 showed evolutionarily conserved function during vacuolar fusion, providing additional support to its identity as a µ1 adaptin. These results demonstrate the importance of the Arabidopsis µ1 adaptin for intracellular protein sorting centered on the TGN/EE.  相似文献   

18.
19.
20.
Phytyl diphosphate (PDP) is the prenyl precursor for tocopherol biosynthesis. Based on recent genetic evidence, PDP is supplied to the tocopherol biosynthetic pathway primarily by chlorophyll degradation and sequential phytol phosphorylation. Three enzymes of Arabidopsis (Arabidopsis thaliana) are known to be capable of removing the phytol chain from chlorophyll in vitro: chlorophyllase1 (CLH1), CLH2, and pheophytin pheophorbide hydrolase (PPH), which specifically hydrolyzes pheophytin. While PPH, but not chlorophyllases, is required for in vivo chlorophyll breakdown during Arabidopsis leaf senescence, little is known about the involvement of these phytol-releasing enzymes in tocopherol biosynthesis. To explore the origin of PDP for tocopherol synthesis, seed tocopherol concentrations were determined in Arabidopsis lines engineered for seed-specific overexpression of PPH and in single and multiple mutants in the three genes encoding known dephytylating enzymes. Except for modestly increasing tocopherol content observed in the PPH overexpressor, none of the remaining lines exhibited significantly reduced tocopherol concentrations, suggesting that the known chlorophyll-derived phytol-releasing enzymes do not play major roles in tocopherol biosynthesis. Tocopherol content of seeds from double mutants in NONYELLOWING1 (NYE1) and NYE2, regulators of chlorophyll degradation, had modest reduction compared with wild-type seeds, although mature seeds of the double mutant retained significantly higher chlorophyll levels. These findings suggest that NYEs may play limited roles in regulating an unknown tocopherol biosynthesis-related phytol hydrolase. Meanwhile, seeds of wild-type over-expressing NYE1 had lower tocopherol levels, suggesting that phytol derived from NYE1-dependent chlorophyll degradation probably doesn’t enter tocopherol biosynthesis. Potential routes of chlorophyll degradation are discussed in relation to tocopherol biosynthesis.Vitamin E tocochromanols are lipidic antioxidants found in photosynthetic organisms that exist as two alternate classes, tocopherols and tocotrienols, which differ in the degree of saturation of the hydrophobic C20 prenyl side chain classes. Among these two classes, four forms occur that differ in methylation of the hydrophilic tocochromanol head group (Sattler et al., 2004). The initial step of tocopherol biosynthesis is the condensation of the aromatic head group precursor homogentisate and the prenyl tail precursor phytyl diphosphate (PDP). This reaction is catalyzed by a plastid-localized enzyme, homogentisate PDP transferase (HPT; Soll et al., 1980; Collakova and DellaPenna, 2001). PDP for tocopherol biosynthesis is either provided through direct reduction of geranylgeranyl diphosphate (Keller et al., 1998) or from chlorophyll-bound phytol through chlorophyll hydrolysis and subsequent conversion of free phytol into PDP by two consecutive kinase reactions (Fig. 1; Rise et al., 1989; Goffman et al., 1999; Matile et al., 1999; Kräutler, 2002; Hörtensteiner, 2006). The first of these phosphorylation steps was shown to be catalyzed by vitamin E pathway5 (VTE5; Valentin et al., 2006).Open in a separate windowFigure 1.The substrate PDP directing toward tocopherol biosynthesis is primarily derived from chlorophyll degradation. Two phytol-releasing activities are known, i.e. CLH catalyzing release from chlorophyll and PPH dephytylating pheophytin. Phytol is then converted to PDP by sequential kinase reactions catalyzed by VTE5 and a second, unknown kinase. Condensation of PDP and homogentisate by HPT marks the initial reaction of tocopherol biosynthesis. phy, Phytyl. [See online article for color version of this figure.]Seeds of the Arabidopsis (Arabidopsis thaliana) vte5 mutant have only about 20% of wild-type concentrations of vitamin E, while containing 3-fold more free phytol compared with seeds of wild-type plants (Valentin et al., 2006). In addition, it has been shown that tocopherol accumulation in Brassica napus seeds correlates with chlorophyll breakdown during seed development (Valentin et al., 2006). Therefore, it was concluded that in Arabidopsis, the 80% of PDP that is used for VTE5-dependent tocopherol biosynthesis in seeds arises from free phytol released during chlorophyll degradation. Chlorophyll degradation is an important catabolic process that is catalyzed by a multistep pathway and occurs during leaf senescence and fruit ripening. An early reaction of the chlorophyll degradation pathway is dephytylation. The true identity of the enzyme(s) associated with phytol release has only recently been revealed. It was long believed that chlorophyllase (CLH) is responsible for phytyl hydrolysis, yielding chlorophyllide and free phytol (Heaton and Marangoni, 1996; Takamiya et al., 2000; Hörtensteiner, 2006). However, analysis of the two CLHs in Arabidopsis, AtCLH1 and AtCLH2 (Tsuchiya et al., 1999; Takamiya et al., 2000), indicated that the AtCLH isoforms are neither chloroplast localized nor essential for senescence-related chlorophyll breakdown (Schenk et al., 2007). These findings are consistent with the observation that not all molecularly identified CLHs contain a predicted chloroplast transit peptide (Jacob-Wilk et al., 1999; Tsuchiya et al., 1999). As a consequence, subcellular compartments distinct from plastids were considered to be additional sites of chlorophyll degradation (Takamiya et al., 2000). By contrast, results obtained from Citrus spp. suggested that CLH functions as a rate-limiting enzyme in chlorophyll catabolism within the chloroplast and is controlled by posttranslational regulation (Harpaz-Saad et al., 2007; Azoulay Shemer et al., 2008). Additionally, work in Arabidopsis indicated that clh2 mutants showed a slight delay in chlorophyll degradation compared with clh1 and wild-type plants (Schenk et al., 2007).More recently, a novel plastid-localized enzyme, pheophytin pheophorbide hydrolase (PPH), was shown to be essential for chlorophyll breakdown during leaf senescence in Arabidopsis. PPH catalyzes the dephytylation of pheophytin rather than chlorophyll, resulting in pheophorbide and free phytol as the products (Schelbert et al., 2009). pph mutants are unable to degrade chlorophyll during senescence and therefore exhibit a stay-green phenotype in leaves. Altogether, these data reflect the complexity of the process of chlorophyll dephytylation and raise the question whether any of these activities may be related to tocopherol biosynthesis.Recently, Gregor Mendel’s green cotyledon gene stay-green (SGR), encoding a chloroplast-localized protein, was shown to be required for the initiation of chlorophyll breakdown (Armstead et al., 2007; Sato et al., 2007). Like in many plant species (Hörtensteiner, 2009), NON-YELLOWING1 (NYE1; also named SGR1), the Arabidopsis homolog of SGR, plays an important positive regulatory role in chlorophyll degradation during senescence, because NYE1 overexpression resulted in either pale-yellow leaves or even albino seedlings, while nye1 mutants retain chlorophyll during senescence (Ren et al., 2007). In addition, the second isoform of NYE in Arabidopsis, NYE2 (also named SGR2), is a negative regulator of chlorophyll degradation in senescent leaves (Sakuraba et al., 2014). By contrast, both enzymes positively contribute to chlorophyll breakdown during seed maturation (Delmas et al., 2013). NYE1 and NYE2 were shown to interact at light-harvesting complex II (LHCII) with other chlorophyll catabolic enzymes, including PPH. This suggests that SGRs might function as scaffold proteins in the formation of a catabolic multienzyme complex regulating chlorophyll degradation (Sakuraba et al., 2012, 2014). Whether NYE1 and NYE2 may also affect CLH function remains unclear, but their role as a key regulators for chlorophyll degradation raises the question whether NYEs may also play a role in tocopherol biosynthesis.Here, by employing Arabidopsis transferred DNA (T-DNA) insertion or nonsense mutants that are defective in known chlorophyll degradation-associated genes, and by PPH or NYE1 overexpression, we provide genetic and physiological evidence that neither CLHs nor PPH plays a major role in tocopherol biosynthesis in Arabidopsis seeds.  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