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 145 毫秒
1.
2.
3.
Self-incompatibility (SI) is an important genetically controlled mechanism to prevent inbreeding in higher plants. SI involves highly specific interactions during pollination, resulting in the rejection of incompatible (self) pollen. Programmed cell death (PCD) is an important mechanism for destroying cells in a precisely regulated manner. SI in field poppy (Papaver rhoeas) triggers PCD in incompatible pollen. During SI-induced PCD, we previously observed a major acidification of the pollen cytosol. Here, we present measurements of temporal alterations in cytosolic pH ([pH]cyt); they were surprisingly rapid, reaching pH 6.4 within 10 min of SI induction and stabilizing by 60 min at pH 5.5. By manipulating the [pH]cyt of the pollen tubes in vivo, we show that [pH]cyt acidification is an integral and essential event for SI-induced PCD. Here, we provide evidence showing the physiological relevance of the cytosolic acidification and identify key targets of this major physiological alteration. A small drop in [pH]cyt inhibits the activity of a soluble inorganic pyrophosphatase required for pollen tube growth. We also show that [pH]cyt acidification is necessary and sufficient for triggering several key hallmark features of the SI PCD signaling pathway, notably activation of a DEVDase/caspase-3-like activity and formation of SI-induced punctate actin foci. Importantly, the actin binding proteins Cyclase-Associated Protein and Actin-Depolymerizing Factor are identified as key downstream targets. Thus, we have shown the biological relevance of an extreme but physiologically relevant alteration in [pH]cyt and its effect on several components in the context of SI-induced events and PCD.Programmed cell death (PCD) in plants is relatively well documented and characterized (Jones and Dangl, 1996; van Doorn, 2011; van Doorn et al., 2011). There is considerable biochemical evidence for the involvement of caspase-like activities in plant PCD (van Doorn and Woltering, 2005). For example, the vacuolar processing enzyme has YVADase (caspase-1-like) activity (Hatsugai et al., 2004; Rojo et al., 2004; Hara-Nishimura et al., 2005), DEVDase (caspase-3-like) and YVADases are associated with PCD in several plant systems (del Pozo and Lam, 1998; Korthout et al., 2000; Danon et al., 2004), and VEIDase (caspase-6-like) is the main caspase-like activity involved in embryonic pattern formation (Bozhkov et al., 2004). However, because plants have no caspase gene homologs (Sanmartín et al., 2005), the nature of their caspase-like enzymes is the subject of considerable debate. Vacuolar cell death is one of two major classes of PCD in plants (van Doorn et al., 2011). It is thought that collapse of the vacuole is a key irreversible step in several plant PCD systems, including during tissue and organ formation, such as the classic differentiation of tracheary elements (Hara-Nishimura and Hatsugai, 2011). Exactly how this is achieved and what processes are involved remain unknown. Until very recently, it was generally thought that the rupturing vacuole releases proteases, hydrolases, and nucleases, allowing cellular disassembly by an autophagy-like process. Some PCD systems cannot be assigned to either class; these include PCD triggered by the hypersensitive response to biotrophic pathogens, PCD in cereal endosperm, and self-incompatibility (SI)-induced PCD (van Doorn et al., 2011).SI is a genetically controlled pollen-pistil cell-cell recognition system. Self-pollen is recognized by the stigma as being genetically identical, resulting in inhibition of pollen tube growth. Most SI systems use tightly linked polymorphic genes: the pollen (male) and pistil (female) S-determinants. In field poppy (Papaver rhoeas), the S-determinants are a 14-kD signaling ligand field poppy stigma S (PrsS) and a unique transmembrane protein field poppy pollen S (PrpS; Foote et al., 1994; Wheeler et al., 2010). These interact in an S-specific manner, and increases in cytosolic free calcium ([Ca2+]cyt) are triggered in incompatible pollen tubes (Franklin-Tong et al., 1993), resulting in phosphorylation of soluble inorganic pyrophosphatases (sPPases; Rudd et al., 1996; de Graaf et al., 2006), activation of a Mitogen-Activated Protein Kinase (MAPK; Rudd et al., 2003), and increases in reactive oxygen species (ROS) and nitric oxide (Wilkins et al., 2011, 2014). Most of these components are integrated into a signaling network leading to PCD (Bosch et al., 2008; Wilkins et al., 2014). The actin cytoskeleton is a key target in the field poppy SI response, undergoing depolymerization (Snowman et al., 2002) followed by polymerization into highly stable F-actin foci decorated with the actin binding proteins (ABPs) Actin-Depolymerizing Factor (ADF) and Cyclase-Associated Protein (CAP; Poulter et al., 2010, 2011), with both processes being involved in mediating PCD (Thomas et al., 2006). A major player in SI-mediated PCD is a caspase-3-like/DEVDase-like activity (Thomas and Franklin-Tong, 2004; Bosch and Franklin-Tong, 2007). The SI-induced caspase-3-like/DEVDase exhibits maximum substrate cleavage in vitro at pH 5, with peak activity 5 h after SI induction in vivo (Bosch and Franklin-Tong, 2007). The low pH optimum for this caspase-3-like/DEVDase activity is unusual, because most of the cytosolic plant caspase-like activities identified to date have optimal activity close to normal physiological pH (approximate pH, 6.5–7.0; Korthout et al., 2000; Bozhkov et al., 2004; Coffeen and Wolpert, 2004). Because the SI-induced cytosolic-located DEVDase requires a low pH for activity, this suggested that, during SI, the pollen tube cytosol undergoes dramatic acidification. In vivo pH measurements of the cytosol at 1 to 4 h after SI induction confirmed this, when cytosolic pH ([pH]cyt) had dropped from pH 6.9 to pH 5.5 (Bosch and Franklin-Tong, 2007). This fits the in vitro pH optimum of the caspase-3-like/DEVDase almost exactly, implicating pollen cytosolic acidification as playing a vital role in creating optimal conditions for the activation of the caspase-3-like/DEVDase-like activity and progression of PCD.Under normal cellular conditions, [pH]cyt is between approximately 6.9 and 7.5 (Kurkdjian and Guern, 1989; Felle, 2001). Pollen tubes, like other tip-growing cells, have [pH]cyt gradients (Gibbon and Kropf, 1994; Feijó et al., 1999). The [pH]cyt of the pollen tube shank is an approximate pH of 6.9 to 7.11 (Fricker et al., 1997; Messerli and Robinson, 1998). There has been much debate about the [pH]cyt gradient, comprising an apical domain with an approximate pH of 6.8 and a subapical alkaline band with an approximate pH of 7.2 to 7.8 in Lilium longiflorum and Lilium formosanum pollen tubes (Fricker et al., 1997; Messerli and Robinson, 1998; Feijó et al., 2001; Lovy-Wheeler et al., 2006). Oscillations of [pH]cyt between approximate pH values of 6.9 and 7.3 have been linked to tip growth in L. formosanum pollen tubes (Lovy-Wheeler et al., 2006). The vacuole and the apoplast have a highly acidic pH between pH 5 and pH 6 (Katsuhara et al., 1989; Feijó et al., 1999). The majority of studies of pH changes in plant cells reports modest, transient changes in [pH]cyt of approximately 0.4 and 0.7 pH units during development, gravitropic responses, decreases in light intensity, and addition of elicitors, hormones, and other treatments. For example, during root hair development in Arabidopsis (Arabidopsis thaliana), root [pH]cyt was elevated from an approximate pH of 7.3 to 7.7 (Bibikova et al., 1998). Root gravitropic responses stimulate small transient [pH]cyt alterations (Scott and Allen, 1999; Fasano et al., 2001; Johannes et al., 2001). More recently, it has been shown that the [pH]cyt drops during PCD controlling root cap development; however, exactly how many units the [pH]cyt decreased was not measured (Fendrych et al., 2014). Other studies investigating [pH]cyt in response to physiologically relevant signals also report small transient alterations. Light-adapted cells respond to a decrease in light intensity with a rapid transient cytosolic acidification by approximately 0.3 pH units (Felle et al., 1986). Addition of nodulation factors resulted in an increase of 0.2 pH units in root hairs (Felle et al., 1998), and abscisic acid increased the [pH]cyt of guard cells by 0.3 pH units (Blatt and Armstrong, 1993). Changes in [pH]cyt are thought to activate stress responses (Felle, 2001). Elicitor treatments resulted in a [pH]cyt drop of between 0.4 and 0.7 pH units in suspension cells (Mathieu et al., 1996; Kuchitsu et al., 1997), a drop of 0.2 pH units in Nitellopsis obtusa cells treated with salt (Katsuhara et al., 1989), and a drop of 0.3 to 0.7 pH units in Eschscholzia californica (Roos et al., 1998).Here, we investigate SI-induced acidification of the cytosol, providing measurements of physiologically relevant temporal alterations in [pH]cyt, and identify key targets of this, providing mechanistic insights into these events. The SI-induced acidification plays a pivotal role in the activation of a caspase-3-like/DEVDase activity, the formation of punctate F-actin foci, and ABP localization during SI PCD. We investigate the vacuole as a potential contributor to SI-induced [pH]cyt acidification.  相似文献   

4.
In plants, the copy number of the mitochondrial DNA (mtDNA) can be much lower than the number of mitochondria. The biological significance and regulatory mechanisms of this phenomenon remain poorly understood. Here, using the pollen vegetative cell, we examined the role of the Arabidopsis (Arabidopsis thaliana) mtDNA-binding protein WHIRLY2 (AtWHY2). AtWHY2 decreases during pollen development, in parallel with the rapid degradation of mtDNA; to examine the importance of this decrease, we used the pollen vegetative cell-specific promoter Lat52 to express AtWHY2. The transgenic plants (LWHY2) had very high mtDNA levels in pollen, more than 10 times more than in the wild type (ecotype Columbia-0). LWHY2 plants were fertile, morphologically normal, and set seeds; however, reciprocal crosses with heterozygous plants showed reduced transmission of LWHY2-1 through the male and slower growth of LWHY2-1 pollen tubes. We found that LWHY2-1 pollen had significantly more reactive oxygen species and less ATP compared with the wild type, indicating an effect on mitochondrial respiration. These findings reveal that AtWHY2 affects mtDNA copy number in pollen and suggest that low mtDNA copy numbers might be the normal means by which plant cells maintain mitochondrial genetic information.Reflecting their endosymbiotic origin, mitochondria contain DNA genomes (mtDNA) encoding several key proteins for oxidative phosphorylation. As most genes identified in the mitochondrial genome are indispensable for mitochondrial function, it is generally believed that each mitochondrion must possess at least one full copy of the genome. Indeed, this seems to be the case in animals. For example, although the number of mitochondria per cell varies in human, mouse, rabbit, and rat cell lines, the mtDNA copy number per mitochondrion remains constant at 2.6 ± 0.3 (Robin and Wong, 1988). Also, in mouse egg cells, each mitochondrion contains an estimated one to two copies of the mtDNA (Pikó and Matsumoto, 1976).Plant cells, however, have very few copies of the mtDNA compared with the number of mitochondria. For example, in the Cucurbitaceae, cells containing 110 to 140 copies of the mtDNA have 360 to 1,100 mitochondria (Bendich and Gauriloff, 1984). In Arabidopsis (Arabidopsis thaliana), leaf cells each contain approximately 670 mitochondria (Sheahan et al., 2005) and approximately 50 copies of the mtDNA (Draper and Hays, 2000). Thus, in plant cells, each mitochondrion does not possess one complete copy of the mtDNA, a phenomenon that occurs commonly in somatic cells of plants (Preuten et al., 2010). In addition, work in Arabidopsis, barley (Hordeum vulgare), and tobacco (Nicotiana tabacum) showed that cells in leaves, stems, and roots contain few copies of the mtDNA (40–160), whereas cells in root tips contain more copies (300–450; Preuten et al., 2010). This is consistent with the mitochondrial nucleoid diminishment previously observed in developing root and shoot tips (Fujie et al., 1993, 1994), which suggests that the low copy numbers in plant cells result from a decrease in the mtDNA copy number in nondividing cells during development.One question raised by these findings is whether some mitochondria have complete mtDNAs while others have no mtDNA or whether mitochondria have partial mtDNAs. Using techniques for the direct visualization of small amounts of DNA, our group revealed that up to two-thirds of mitochondria in Arabidopsis mesophyll cells totally lack mtDNA and the remaining one-third of mitochondria possess mtDNA of about 100 kb on average (Wang et al., 2010). This agrees well with a previously reported value for mtDNA copy number (about 50 copies per cell; Draper and Hays, 2000) and is consistent with the idea that plant mitochondrial genomes exist as submolecules smaller than the total genomic sizes (Satoh et al., 1993; Kubo and Newton, 2008). Among plant cells possessing low mtDNA copy numbers, the vegetative cell in the pollen grains is an extreme case; a mature pollen grain of Antirrhinum majus, containing many more mitochondria than a somatic cell, possesses only 16 copies of the mtDNA (Wang et al., 2010). Similar to the changes observed in somatic cells, this extremely low level of mtDNA in pollen vegetative cells results from a rapid decrease in mtDNA copy number during pollen development (Sodmergen et al., 1991; Nagata et al., 1999). In A. majus, the vegetative cell in its initial developmental stage has 482.7 copies of the mtDNA per cell, indicating a 30-fold decrease (482.7/16) during development (Wang et al., 2010). These results from both somatic and reproductive cells led to the intriguing idea that the mtDNA copy number in plants decreases in parallel with cell differentiation, to a very low value, and thus that several mitochondria must share the genetic information carried on a single copy of the mtDNA. Plant cell mitochondria undergo frequent and coupled fusions and divisions, which may explain how mitochondria share this information (Arimura et al., 2004). However, the biological significance of why plant cells lose their mtDNA, and how this benefits these cells, remains unknown. Given that pollen germination, pollen tube elongation, and sperm cell delivery all require energy conversion, the extremely low mtDNA copy numbers, such as in pollen vegetative cells, must not compromise mitochondrial function.The mtDNA copy numbers remain constant in various tissues, however, indicating that cellular mechanisms accurately regulate the levels of mtDNA in relation to cell type (Robin and Wong, 1988; Preuten et al., 2010). In yeast and animals, this regulation involves the core enzymes of mtDNA replication, such as DNA polymerase-γ (Sharief et al., 1999), RNA polymerase (Wanrooij et al., 2008), and mitochondrial helicase (Liu et al., 2009), as well as a group of DNA-binding proteins such as ARS-binding factor2 protein in yeast (Saccharomyces cerevisiae; Newman et al., 1996), MITOCHONDRIAL TRANSCRIPTION FACTOR A (TFAM) in human (Alam et al., 2003), and mitochondrial single-stranded DNA binding protein in Drosophila spp. (Maier et al., 2001). Overexpression of TFAM causes an increase in the mtDNA copy number, and RNA interference of TFAM decreases the mtDNA copy number (Ekstrand et al., 2004; Kanki et al., 2004). Also, the homozygous knockout of TFAM in mouse results in embryos that lack mtDNA and thus fail to survive (Larsson et al., 1998). Clearly, protein factors within mitochondrial nucleoids play a crucial role in the regulation of mtDNA copy number.Recent investigation in Arabidopsis revealed that, similar to the case in yeast and animal cells, DNA polymerase, the core enzyme of mtDNA replication, functions to maintain mtDNA levels in plants. Mutation of Arabidopsis PolIA or PolIB (homologs of bacterial DNA polymerase I) causes a reduction in mtDNA copy number, and double mutation of these proteins is lethal (Parent et al., 2011). Also, an Mg2+-dependent exonuclease, DEFECTIVE IN POLLEN ORGANELLE DNA DEGRADATION1 (DPD1), degrades organelle DNA, helping to produce the proper amounts of mtDNA in pollen cells (Matsushima et al., 2011; Tang et al., 2012). These results provide insights into the molecular control of mtDNA levels in plants, via both mtDNA replication and mtDNA degradation. Except for these enzymes, however, other protein factors (such as TFAM in animals) have not been identified in plants. The DNA-binding proteins, such as MutS Homolog1 (MSH1), Organellar Single-Strand DNA Binding Protein1 (OSB1), Recombinase A1 (RecA1), RecA3, and WHIRLY2 (WHY2), identified so far in plant mitochondria likely participate in genomic maintenance by affecting substoichiometric shifting (Abdelnoor et al., 2003), stoichiometric transmission (Zaegel et al., 2006), genomic stability (Shedge et al., 2007; Odahara et al., 2009), and DNA repair (Cappadocia et al., 2010). None of these plant nucleoid factors (DNA-binding proteins) has been implicated in the control of mtDNA copy number; thus, the mechanisms by which nonenzyme protein factors regulate mtDNA copy number in plants remain obscure.To test whether nucleoid DNA-binding proteins can affect mtDNA copy number, we examined the effect of producing Arabidopsis WHY2, a single-stranded DNA-binding protein (Cappadocia et al., 2010), in the pollen vegetative cell, which generally does not express WHY2 (Honys and Twell, 2004). We found that expression of WHY2 resulted in a 10-fold increase in mtDNA copy number in the pollen vegetative cell. This increase affected mitochondrial respiration, mitochondrial size, and pollen tube growth. Thus, our results uncover a novel function for WHY2, a member of the plant Whirly protein family, in regulating mtDNA amounts and indicate that, in plants, low mtDNA copy number does not compromise mitochondrial function but rather promotes proper mitochondrial function.  相似文献   

5.
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   

6.
7.
Necrotrophic and biotrophic pathogens are resisted by different plant defenses. While necrotrophic pathogens are sensitive to jasmonic acid (JA)-dependent resistance, biotrophic pathogens are resisted by salicylic acid (SA)- and reactive oxygen species (ROS)-dependent resistance. Although many pathogens switch from biotrophy to necrotrophy during infection, little is known about the signals triggering this transition. This study is based on the observation that the early colonization pattern and symptom development by the ascomycete pathogen Plectosphaerella cucumerina (P. cucumerina) vary between inoculation methods. Using the Arabidopsis (Arabidopsis thaliana) defense response as a proxy for infection strategy, we examined whether P. cucumerina alternates between hemibiotrophic and necrotrophic lifestyles, depending on initial spore density and distribution on the leaf surface. Untargeted metabolome analysis revealed profound differences in metabolic defense signatures upon different inoculation methods. Quantification of JA and SA, marker gene expression, and cell death confirmed that infection from high spore densities activates JA-dependent defenses with excessive cell death, while infection from low spore densities induces SA-dependent defenses with lower levels of cell death. Phenotyping of Arabidopsis mutants in JA, SA, and ROS signaling confirmed that P. cucumerina is differentially resisted by JA- and SA/ROS-dependent defenses, depending on initial spore density and distribution on the leaf. Furthermore, in situ staining for early callose deposition at the infection sites revealed that necrotrophy by P. cucumerina is associated with elevated host defense. We conclude that P. cucumerina adapts to early-acting plant defenses by switching from a hemibiotrophic to a necrotrophic infection program, thereby gaining an advantage of immunity-related cell death in the host.Plant pathogens are often classified as necrotrophic or biotrophic, depending on their infection strategy (Glazebrook, 2005; Nishimura and Dangl, 2010). Necrotrophic pathogens kill living host cells and use the decayed plant tissue as a substrate to colonize the plant, whereas biotrophic pathogens parasitize living plant cells by employing effector molecules that suppress the host immune system (Pel and Pieterse, 2013). Despite this binary classification, the majority of pathogenic microbes employ a hemibiotrophic infection strategy, which is characterized by an initial biotrophic phase followed by a necrotrophic infection strategy at later stages of infection (Perfect and Green, 2001). The pathogenic fungi Magnaporthe grisea, Sclerotinia sclerotiorum, and Mycosphaerella graminicola, the oomycete Phytophthora infestans, and the bacterial pathogen Pseudomonas syringae are examples of hemibiotrophic plant pathogens (Perfect and Green, 2001; Koeck et al., 2011; van Kan et al., 2014; Kabbage et al., 2015).Despite considerable progress in our understanding of plant resistance to necrotrophic and biotrophic pathogens (Glazebrook, 2005; Mengiste, 2012; Lai and Mengiste, 2013), recent debate highlights the dynamic and complex interplay between plant-pathogenic microbes and their hosts, which is raising concerns about the use of infection strategies as a static tool to classify plant pathogens. For instance, the fungal genus Botrytis is often labeled as an archetypal necrotroph, even though there is evidence that it can behave as an endophytic fungus with a biotrophic lifestyle (van Kan et al., 2014). The rice blast fungus Magnaporthe oryzae, which is often classified as a hemibiotrophic leaf pathogen (Perfect and Green, 2001; Koeck et al., 2011), can adopt a purely biotrophic lifestyle when infecting root tissues (Marcel et al., 2010). It remains unclear which signals are responsible for the switch from biotrophy to necrotrophy and whether these signals rely solely on the physiological state of the pathogen, or whether host-derived signals play a role as well (Kabbage et al., 2015).The plant hormones salicylic acid (SA) and jasmonic acid (JA) play a central role in the activation of plant defenses (Glazebrook, 2005; Pieterse et al., 2009, 2012). The first evidence that biotrophic and necrotrophic pathogens are resisted by different immune responses came from Thomma et al. (1998), who demonstrated that Arabidopsis (Arabidopsis thaliana) genotypes impaired in SA signaling show enhanced susceptibility to the biotrophic pathogen Hyaloperonospora arabidopsidis (formerly known as Peronospora parastitica), while JA-insensitive genotypes were more susceptible to the necrotrophic fungus Alternaria brassicicola. In subsequent years, the differential effectiveness of SA- and JA-dependent defense mechanisms has been confirmed in different plant-pathogen interactions, while additional plant hormones, such as ethylene, abscisic acid (ABA), auxins, and cytokinins, have emerged as regulators of SA- and JA-dependent defenses (Bari and Jones, 2009; Cao et al., 2011; Pieterse et al., 2012). Moreover, SA- and JA-dependent defense pathways have been shown to act antagonistically on each other, which allows plants to prioritize an appropriate defense response to attack by biotrophic pathogens, necrotrophic pathogens, or herbivores (Koornneef and Pieterse, 2008; Pieterse et al., 2009; Verhage et al., 2010).In addition to plant hormones, reactive oxygen species (ROS) play an important regulatory role in plant defenses (Torres et al., 2006; Lehmann et al., 2015). Within minutes after the perception of pathogen-associated molecular patterns, NADPH oxidases and apoplastic peroxidases generate early ROS bursts (Torres et al., 2002; Daudi et al., 2012; O’Brien et al., 2012), which activate downstream defense signaling cascades (Apel and Hirt, 2004; Torres et al., 2006; Miller et al., 2009; Mittler et al., 2011; Lehmann et al., 2015). ROS play an important regulatory role in the deposition of callose (Luna et al., 2011; Pastor et al., 2013) and can also stimulate SA-dependent defenses (Chaouch et al., 2010; Yun and Chen, 2011; Wang et al., 2014; Mammarella et al., 2015). However, the spread of SA-induced apoptosis during hyperstimulation of the plant immune system is contained by the ROS-generating NADPH oxidase RBOHD (Torres et al., 2005), presumably to allow for the sufficient generation of SA-dependent defense signals from living cells that are adjacent to apoptotic cells. Nitric oxide (NO) plays an additional role in the regulation of SA/ROS-dependent defense (Trapet et al., 2015). This gaseous molecule can stimulate ROS production and cell death in the absence of SA while preventing excessive ROS production at high cellular SA levels via S-nitrosylation of RBOHD (Yun et al., 2011). Recently, it was shown that pathogen-induced accumulation of NO and ROS promotes the production of azelaic acid, a lipid derivative that primes distal plants for SA-dependent defenses (Wang et al., 2014). Hence, NO, ROS, and SA are intertwined in a complex regulatory network to mount local and systemic resistance against biotrophic pathogens. Interestingly, pathogens with a necrotrophic lifestyle can benefit from ROS/SA-dependent defenses and associated cell death (Govrin and Levine, 2000). For instance, Kabbage et al. (2013) demonstrated that S. sclerotiorum utilizes oxalic acid to repress oxidative defense signaling during initial biotrophic colonization, but it stimulates apoptosis at later stages to advance necrotrophic colonization. Moreover, SA-induced repression of JA-dependent resistance not only benefits necrotrophic pathogens but also hemibiotrophic pathogens after having switched from biotrophy to necrotrophy (Glazebrook, 2005; Pieterse et al., 2009, 2012).Plectosphaerella cucumerina ((P. cucumerina, anamorph Plectosporum tabacinum) anamorph Plectosporum tabacinum) is a filamentous ascomycete fungus that can survive saprophytically in soil by decomposing plant material (Palm et al., 1995). The fungus can cause sudden death and blight disease in a variety of crops (Chen et al., 1999; Harrington et al., 2000). Because P. cucumerina can infect Arabidopsis leaves, the P. cucumerina-Arabidopsis interaction has emerged as a popular model system in which to study plant defense reactions to necrotrophic fungi (Berrocal-Lobo et al., 2002; Ton and Mauch-Mani, 2004; Carlucci et al., 2012; Ramos et al., 2013). Various studies have shown that Arabidopsis deploys a wide range of inducible defense strategies against P. cucumerina, including JA-, SA-, ABA-, and auxin-dependent defenses, glucosinolates (Tierens et al., 2001; Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014), callose deposition (García-Andrade et al., 2011; Gamir et al., 2012, 2014; Sánchez-Vallet et al., 2012), and ROS (Tierens et al., 2002; Sánchez-Vallet et al., 2010; Barna et al., 2012; Gamir et al., 2012, 2014; Pastor et al., 2014). Recent metabolomics studies have revealed large-scale metabolic changes in P. cucumerina-infected Arabidopsis, presumably to mobilize chemical defenses (Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014). Furthermore, various chemical agents have been reported to induce resistance against P. cucumerina. These chemicals include β-amino-butyric acid, which primes callose deposition and SA-dependent defenses, benzothiadiazole (BTH or Bion; Görlach et al., 1996; Ton and Mauch-Mani, 2004), which activates SA-related defenses (Lawton et al., 1996; Ton and Mauch-Mani, 2004; Gamir et al., 2014; Luna et al., 2014), JA (Ton and Mauch-Mani, 2004), and ABA, which primes ROS and callose deposition (Ton and Mauch-Mani, 2004; Pastor et al., 2013). However, among all these studies, there is increasing controversy about the exact signaling pathways and defense responses contributing to plant resistance against P. cucumerina. While it is clear that JA and ethylene contribute to basal resistance against the fungus, the exact roles of SA, ABA, and ROS in P. cucumerina resistance vary between studies (Thomma et al., 1998; Ton and Mauch-Mani, 2004; Sánchez-Vallet et al., 2012; Gamir et al., 2014).This study is based on the observation that the disease phenotype during P. cucumerina infection differs according to the inoculation method used. We provide evidence that the fungus follows a hemibiotrophic infection strategy when infecting from relatively low spore densities on the leaf surface. By contrast, when challenged by localized host defense to relatively high spore densities, the fungus switches to a necrotrophic infection program. Our study has uncovered a novel strategy by which plant-pathogenic fungi can take advantage of the early immune response in the host plant.  相似文献   

8.
9.
10.
Polarized exocytosis is critical for pollen tube growth, but its localization and function are still under debate. The exocyst vesicle-tethering complex functions in polarized exocytosis. Here, we show that a sec3a exocyst subunit null mutant cannot be transmitted through the male gametophyte due to a defect in pollen tube growth. The green fluorescent protein (GFP)-SEC3a fusion protein is functional and accumulates at or proximal to the pollen tube tip plasma membrane. Partial complementation of sec3a resulted in the development of pollen with multiple tips, indicating that SEC3 is required to determine the site of pollen germination pore formation. Time-lapse imaging demonstrated that SEC3a and SEC8 were highly dynamic and that SEC3a localization on the apical plasma membrane predicts the direction of growth. At the tip, polar SEC3a domains coincided with cell wall deposition. Labeling of GFP-SEC3a-expressing pollen with the endocytic marker FM4-64 revealed the presence of subdomains on the apical membrane characterized by extensive exocytosis. In steady-state growing tobacco (Nicotiana tabacum) pollen tubes, SEC3a displayed amino-terminal Pleckstrin homology-like domain (SEC3a-N)-dependent subapical membrane localization. In agreement, SEC3a-N interacted with phosphoinositides in vitro and colocalized with a phosphatidylinositol 4,5-bisphosphate (PIP2) marker in pollen tubes. Correspondingly, molecular dynamics simulations indicated that SEC3a-N associates with the membrane by interacting with PIP2. However, the interaction with PIP2 is not required for polar localization and the function of SEC3a in Arabidopsis (Arabidopsis thaliana). Taken together, our findings indicate that SEC3a is a critical determinant of polar exocytosis during tip growth and suggest differential regulation of the exocytotic machinery depending on pollen tube growth modes.Pollen tube growth provides a unique model system for studying the role of exocytosis in cell morphogenesis. Pollen tubes are characterized by a highly rapid polarized unidirectional tip growth. Given the relative simplicity of their structure, fast growth rates, haploid genome content, and ability to grow under in vitro culture conditions, pollen tubes provide an extremely attractive system for studying cell morphogenesis. Furthermore, the growth characteristics of pollen tubes resemble those of root hairs, moss protonema, and fungal hyphae and to some extent can be paralleled to neurite growth (Chebli and Geitmann, 2007; Cheung and Wu, 2008; Guan et al., 2013; Hepler and Winship, 2015).It is well established that oscillating polarized exocytosis is fundamental for pollen tube development and determines growth rate (Bove et al., 2008; McKenna et al., 2009; Chebli et al., 2013). Exocytosis is required for the delivery of membrane and cell wall components to the growing tip. Yet, the exact location where exocytosis takes place is under debate. Ultrastructural studies showing the accumulation of vesicles at the tip suggested that exocytosis takes place at the tip (Lancelle et al., 1987; Lancelle and Hepler, 1992; Derksen et al., 1995), which was further supported by studies on the dynamics of cell wall thickness (Rojas et al., 2011), secretion of pectin methyl esterase (PME) and PME inhibitor, and staining of pectin by propidium iodide (PI; Röckel et al., 2008; Rounds et al., 2014). Conversely, based on colabeling with FM1-43 and FM4-64, it was concluded that exocytosis takes place in a subapical collar located in the transition zone between the tip and the shank, as well as at the shank, but not at the tip (Bove et al., 2008; Zonia and Munnik, 2008). In agreement, the pollen tube-specific syntaxin GFP-SYP124 was observed in the inverted cone, 10 to 25 μm away from the tip (Silva et al., 2010), and fluorescence recovery after photobleaching experiments with FM dyes also have indicated that exocytosis takes place at the subapical region (Bove et al., 2008; Moscatelli et al., 2012; Idilli et al., 2013). Yet, based on pollen tube reorientation experiments in a microfluidics device, it was concluded that growth takes place at the tip rather than at a subapical collar located in the transition zone between the apex and the shank (Sanati Nezhad et al., 2014). The tip-based growth is in agreement with exocytosis taking place at the tip. Presumably, part of the disagreement regarding the site of exocytosis resulted from the lack of intracellular markers for exocytosis (Cheung and Wu, 2008; Hepler and Winship, 2015), and as a result, the relationship between the FM dye-labeled inverted cone and exocytotic events during pollen tube growth is not fully understood.In many cell types, the process of secretory vesicles tethering and docking prior to fusion with the plasma membrane is initially mediated by an evolutionarily conserved tethering complex known as the exocyst. The exocyst is a heterooligomeric protein complex composed of eight subunits, SEC3, SEC5, SEC6, SEC8, SEC10, SEC15, EXO70, and EXO84 (TerBush et al., 1996; Guo et al., 1999). Studies originally based on budding yeast (Saccharomyces cerevisiae) have shown that the exocyst functions as an effector of Rab and Rho small GTPases that specifies the sites of vesicle docking and fusion at the plasma membrane in both space and time (Guo et al., 2001; Zhang et al., 2001). Support for the function of the exocyst in vesicle tethering was demonstrated recently by ectopic Sec3p-dependent vesicle recruitment to the mitochondria (Luo et al., 2014).Land plants contain all subunits of the exocyst complex, which were shown to form the functional complex (Elias et al., 2003; Cole et al., 2005; Synek et al., 2006; Hála et al., 2008). Studies in Arabidopsis (Arabidopsis thaliana) and maize (Zea mays) have implicated the exocyst in the regulation of pollen tube and root hair growth, seed coat deposition, response to pathogens, cytokinesis, and meristem and stigma function (Cole et al., 2005; Synek et al., 2006; Hála et al., 2008; Fendrych et al., 2010; Kulich et al., 2010; Pecenková et al., 2011; Safavian and Goring, 2013; Wu et al., 2013; Safavian et al., 2015; Zhang et al., 2016). The growth arrest of pollen tubes in sec8, sec6, sec15a, and sec5a/sec5b single and double mutants (Cole et al., 2005; Hála et al., 2008) or following treatment with the EXO70 inhibitor ENDOSIDIN2 (Zhang et al., 2016), and of root hairs in maize root hairless1 (rth1) SEC3 mutant (Wen et al., 2005), the inhibition of seed coat deposition in the sec8 and exo70A1 mutants (Kulich et al., 2010), and stigmatic papillae function in exo70A1 mutant plants (Safavian and Goring, 2013; Safavian et al., 2015) have implicated the exocyst in polarized exocytosis in plants. Given their function, it was likely that exocyst subunits could be used as markers for polarized exocytosis. Furthermore, it could also be hypothesized that, by studying the mechanisms that underlie the association of the exocyst complex with the plasma membrane, it should be possible to identify mechanisms underlying the regulation of polarized exocytosis (Guan et al., 2013). Moreover, since the interaction of exocytotic vesicles with the exocyst is transient and marks the site(s) of active exocytosis in the membrane, fluorescently labeled exocyst subunits could be used as markers for exocytosis while avoiding potential imaging artifacts stemming from pollen tube tips densely populated with vesicles.We have shown previously that the ROP effector ICR1 can interact with SEC3a and that ROPs can recruit SEC3a-ICR1 complexes to the plasma membrane (Lavy et al., 2007). However, ICR1 is not expressed in pollen tubes, suggesting that SEC3a membrane binding in these cells is likely dependent on other factors. In yeast, the interaction of Sec3p and Exo70p subunits with the plasma membrane is critical for exocyst function (He and Guo, 2009). It has been shown that the membrane binding of both Sec3p and Exo70p is facilitated by their interaction with phosphatidylinositol 4,5-bisphosphate (PIP2; He et al., 2007; Zhang et al., 2008). The yeast Exo70p interacts with PIP2 via a number of positively charged residues distributed along the protein, with the highest number located at the C-terminal end (Pleskot et al., 2015). It has been suggested that yeast Sec3p interacts with PIP2 through N-terminal basic residues (Zhang et al., 2008). These data were further corroborated by x-ray crystallography studies, which showed that the yeast Sec3p N-terminal region forms a Pleckstrin homology (PH) domain fold (Baek et al., 2010; Yamashita et al., 2010), a PIP2 interaction motif (Lemmon, 2008).The localization of the exocyst subunits has been addressed in several studies. In Arabidopsis root hairs and root epidermis cells, SEC3a-GFP was observed in puncta distributed throughout the cell (Zhang et al., 2013). Studies on the Arabidopsis EXO70 subunits EXO70E2, EXO70A1, and EXO70B1 revealed them to be localized in distinct compartments that were termed exocyst-positive organelles (Wang et al., 2010). The exocyst-positive organelles, visualized mostly by ectopic expression, were shown to be cytoplasmic double membrane organelles that can fuse with the plasma membrane and secrete their contents to the apoplast in an exosome-like manner. It is not yet known whether other exocyst subunits also are localized to the same organelles and what might be the biological function of this putative compartment (Wang et al., 2010; Lin et al., 2015). In differentiating xylem cells, two coiled-coil proteins termed VESICLE TETHERING1 and VESICLE TETHERING2 recruit EXO70A1-positive puncta to microtubules via the GOLGI COMPLEX2 protein (Oda et al., 2015). Importantly, the functionality of the XFP fusion proteins used for the localization studies described above was not tested, and in most cases, the fusion proteins were overexpressed. Therefore, the functional localization of the exocyst is still unclear.Here, we studied the function and subcellular localization of the Arabidopsis exocyst SEC3a subunit using a combination of genetics, cell biology, biochemistry, and structural modeling approaches. Our results show that SEC3a is essential for the determination of pollen tube tip germination site and growth. Partial complementation of sec3a resulted in the formation of pollen with multiple pollen tube tips. In Arabidopsis growing pollen tubes, SEC3a localization is dynamic, and it accumulates in domains of polarized secretion, at or close to the tip plasma membrane (PM). Labeling of GFP-SEC3-expressing pollen with FM4-64 revealed the spatial correlation between polarized exocytosis and endocytic recycling. Furthermore, the association of SEC3a with PM at the tip marks the direction of tube elongation and positively correlates with the deposition of PI-labeled pectins and specific anti-esterified pectin antibodies in the cell wall. In tobacco (Nicotiana tabacum), the mechanisms underlying SEC3a interaction with the PM and its subcellular distribution depend on pollen tube growth mode and involve the interaction with PIP2 through the N-terminal PH domain. Collectively, our results highlight the function of SEC3a as a polarity determinant that links between polarized exocytosis and cell morphogenesis. The correlation between exocyst function and distribution in pollen tubes provides an explanation for some of the current discrepancies regarding the localization of exocytosis.  相似文献   

11.
12.
Calcium plays an essential role in pollen tube tip growth. However, little is known concerning the molecular basis of the signaling pathways involved. Here, we identified Arabidopsis (Arabidopsis thaliana) CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASE19 (CIPK19) as an important element to pollen tube growth through a functional survey for CIPK family members. The CIPK19 gene was specifically expressed in pollen grains and pollen tubes, and its overexpression induced severe loss of polarity in pollen tube growth. In the CIPK19 loss-of-function mutant, tube growth and polarity were significantly impaired, as demonstrated by both in vitro and in vivo pollen tube growth assays. Genetic analysis indicated that disruption of CIPK19 resulted in a male-specific transmission defect. Furthermore, loss of polarity induced by CIPK19 overexpression was associated with elevated cytosolic Ca2+ throughout the bulging tip, whereas LaCl3, a Ca2+ influx blocker, rescued CIPK19 overexpression-induced growth inhibition. Our results suggest that CIPK19 may be involved in maintaining Ca2+ homeostasis through its potential function in the modulation of Ca2+ influx.In flowering plants, fertilization is mediated by pollen tubes that extend directionally toward the ovule for sperm delivery (Krichevsky et al., 2007; Johnson, 2012). The formation of these elongated tubular structures is dependent on extreme polar growth (termed tip growth), in which cell expansion occurs exclusively in the very apical area (Yang, 2008; Rounds and Bezanilla, 2013). As this type of tip growth is amenable to genetic manipulation and cell biological analysis, the pollen tube is an excellent model system for the functional analysis of essential genes involved in polarity control and fertilization (Yang, 2008; Qin and Yang, 2011; Bloch and Yalovsky, 2013).It is well established that Ca2+ plays a critical role in pollen germination and tube growth (Konrad et al., 2011; Hepler et al., 2012). A steep tip-focused Ca2+ gradient has been detected at the tip of elongating pollen tubes (Rathore et al., 1991; Pierson et al., 1994; Hepler, 1997). In previous studies, artificial dissipation of the Ca2+ gradient seriously inhibited tip growth of pollen tubes, whereas elevation of internal Ca2+ level induced bending of the growth axis toward the zone of higher Ca2+. These studies suggest that Ca2+ not only controls pollen tube elongation but also modulates growth orientation (Miller et al., 1992; Malho et al., 1994; Malho and Trewavas, 1996; Hepler, 1997). These Ca2+ signatures are perceived and relayed to downstream responses by a complex toolkit of Ca2+-binding proteins that function as Ca2+ sensors (Yang and Poovaiah, 2003; Harper et al., 2004; Dodd et al., 2010).To date, four major Ca2+ sensor families have been identified in Arabidopsis (Arabidopsis thaliana), including calcium-dependent protein kinase, calmodulin (CaM), calmodulin-like (CML), and CALCINEURIN B-LIKE (CBL) proteins (Luan et al., 2002, 2009; Yang and Poovaiah, 2003; Harper et al., 2004). Calcium-dependent protein kinase family members comprise a kinase domain and a CaM-like domain in a single protein; thus, they act not only as a Ca2+ sensor but also as an effector, designated as sensor responders (Cheng et al., 2002). In contrast, CaM, CML, and CBL proteins do not have any enzymatic domains but transmit Ca2+ signals to downstream targets via Ca2+-dependent protein-protein interactions. Therefore, they have been designated as sensor relays (McCormack et al., 2005). While CaM and CML proteins interact with a diverse array of target proteins, it is generally accepted that CBLs interact specifically with a group of Ser/Thr protein kinases termed CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASEs (CIPKs; Luan et al., 2002; Kolukisaoglu et al., 2004).In Arabidopsis, several CBLs coupled with their target CIPKs have been demonstrated to function in the regulation of ion homeostasis and stress responses (Luan et al., 2009). Under salt stress, SALT OVERLY SENSITIVE3 (SOS3)/CBL4-SOS2/CIPK24 regulate SOS1 at the plasma membrane for Na+ exclusion, whereas CBL10-CIPK24 complexes appear to regulate Na+ sequestration at the tonoplast (Liu et al., 2000; Qiu et al., 2002; Kim et al., 2007; Quan et al., 2007). For low-K+ stress, CBL1 and CBL9, with 87% amino acid sequence identity, interact with CIPK23, which regulates a voltage-gated ion channel (ARABIDOPSIS K+ TRANSPORTER1) to mediate the uptake of K+ in root hairs (Li et al., 2006; Xu et al., 2006; Cheong et al., 2007). In addition, CBL1 integrates plant responses to cold, drought, salinity, and hyperosmotic stresses (Albrecht et al., 2003; Cheong et al., 2003), and CBL9 is involved in abscisic acid signaling and biosynthesis during seed germination (Pandey et al., 2004). Over the past decade, the functions of CBL-CIPK complexes in abiotic stress tolerance have been studied extensively, but only limited studies focus on CBL family members in pollen tube growth. For example, CBL3 overexpression caused a defective phenotype in pollen tube growth (Zhou et al., 2009). Overexpression of CBL1 or its closest homolog CBL9 inhibited pollen germination and perturbed tube growth at high external K+, whereas disruption of CBL1 and CBL9 leads to a significantly reduced growth rate of pollen tubes under low-K+ conditions (Mähs et al., 2013). The potential roles of CIPKs in pollen tubes so far appear to be completely unknown.In this study, we demonstrated that Arabidopsis CIPK19, a CIPK specifically expressed in pollen grains and pollen tubes, functions in pollen tube tip growth, providing a new insight into the function of the CBL-CIPK network in the control of growth polarity during pollen tube extension in fertilization.  相似文献   

13.
In lily (Lilium formosanum) pollen tubes, pectin, a major component of the cell wall, is delivered through regulated exocytosis. The targeted transport and secretion of the pectin-containing vesicles may be controlled by the cortical actin fringe at the pollen tube apex. Here, we address the role of the actin fringe using three different inhibitors of growth: brefeldin A, latrunculin B, and potassium cyanide. Brefeldin A blocks membrane trafficking and inhibits exocytosis in pollen tubes; it also leads to the degradation of the actin fringe and the formation of an aggregate of filamentous actin at the base of the clear zone. Latrunculin B, which depolymerizes filamentous actin, markedly slows growth but allows focused pectin deposition to continue. Of note, the locus of deposition shifts frequently and correlates with changes in the direction of growth. Finally, potassium cyanide, an electron transport chain inhibitor, briefly stops growth while causing the actin fringe to completely disappear. Pectin deposition continues but lacks focus, instead being delivered in a wide arc across the pollen tube tip. These data support a model in which the actin fringe contributes to the focused secretion of pectin to the apical cell wall and, thus, to the polarized growth of the pollen tube.Pollen tubes provide an excellent model for studying the molecular and physiological processes that lead to polarized cell growth. Because all plant cell growth results from the regulated yielding of the cell wall in response to uniform turgor pressure (Winship et al., 2010; Rojas et al., 2011), the cell wall of the pollen tube must yield only at a particular spot: the cell apex, or tip. To accomplish the extraordinary growth rates seen in many species, and to balance the thinning of the apical wall due to rapid expansion, the pollen tube delivers prodigious amounts of wall material, largely methoxylated pectins, to the tip in a coordinated manner. Recent studies suggest that the targeted exocytosis increases the extensibility of the cell wall matrix at the tip, which then yields to the existing turgor pressure, permitting the tip to extend or grow (McKenna et al., 2009; Hepler et al., 2013). There are many factors that influence exocytosis in growing pollen tubes; in this study, we investigate the role of the apical actin fringe.For many years, it has been known that an actin structure exists near the pollen tube tip, yet its exact form has been a matter of some contention (Kost et al., 1998; Lovy-Wheeler et al., 2005; Wilsen et al., 2006; Cheung et al., 2008; Vidali et al., 2009; Qu et al., 2013). The apical actin structure has been variously described as a fringe, a basket, a collar, or a mesh. Using rapid freeze fixation of lily (Lilium formosanum) pollen tubes followed by staining with anti-actin antibodies, the structure appears as a dense fringe of longitudinally oriented microfilaments, beginning 1 to 5 µm behind the apex and extending 5 to 10 µm basally. The actin filaments are positioned in the cortical cytoplasm close to the plasma membrane (Lovy-Wheeler et al., 2005). More recently, we used Lifeact-mEGFP, a probe that consistently labels this palisade of longitudinally oriented microfilaments in living cells (Vidali et al., 2009; Fig. 1A, left column). For the purposes of this study, we will refer to this apical organization of actin as a fringe.Open in a separate windowFigure 1.The actin fringe and the thickened pollen tube tip wall are stable, although dynamic, structures during pollen tube growth. A, The left column shows a pollen tube transformed with Lifeact-mEGFP imaged with a spinning-disc confocal microscope. Maximal projections from every 15 s are shown. The right column shows epifluorescence images of a pollen tube stained with PI. Again, images captured every 15 s are shown. Bars = 10 μm. B, The data from the pollen tube in A expressing Lifeact-mEGFP were subjected to kymograph analysis using an 11-pixel strip along the image’s midline. C, The first three frames from the pollen tube in A and B were assigned the colors red, blue, and green, respectively, and then overlaid. Areas with white show the overlap of all three. The fringe is stable, but most of its constituent actin is not shared between frames.Many lines of evidence demonstrate that actin is required for pollen tube growth. Latrunculin B (LatB), which blocks actin polymerization, inhibits pollen tube growth and disrupts the cortical fringe at concentrations as low as 2 nm. Higher concentrations are needed to block pollen grain germination and cytoplasmic streaming (Gibbon et al., 1999; Vidali et al., 2001). Actin-binding proteins, including actin depolymerizing factor-cofilin, formin, profilin, and villin, and signaling proteins, such as Rho-of-Plants (ROP) GTPases and their effectors (ROP interacting crib-containing proteins [RICs]), also have been shown to play critical roles in growth and actin dynamics (Fu et al., 2001; Vidali et al., 2001; Allwood et al., 2002; Chen et al., 2002; Cheung and Wu, 2004; McKenna et al., 2004; Gu et al., 2005; Ye et al., 2009; Cheung et al., 2010; Staiger et al., 2010; Zhang et al., 2010a; Qu et al., 2013; van Gisbergen and Bezanilla, 2013).Our understanding of the process of exocytosis and pollen tube elongation has been influenced by ultrastructural images of pollen tube tips, which reveal an apical zone dense with vesicles (Cresti et al., 1987; Heslop-Harrison, 1987; Lancelle et al., 1987; Steer and Steer, 1989; Lancelle and Hepler, 1992; Derksen et al., 1995). It has long been assumed that these represent exocytotic vesicles destined to deliver new cell wall material. This model of polarized secretion has been challenged in recent years in studies using FM dyes. Two groups have suggested that exocytosis occurs in a circumpolar annular zone (Bove et al., 2008; Zonia and Munnik, 2008). However, other studies, using fluorescent beads attached to the cell surface, indicate that the maximal rate of expansion, and of necessity the greatest deposition of cell wall material, occurs at the apex along the polar axis of the tube (Dumais et al., 2006; Rojas et al., 2011). Similarly, our experiments with propidium iodide (PI; McKenna et al., 2009; Rounds et al., 2011a) and pectin methyl esterase fused to GFP (McKenna et al., 2009) show that the wall is thickest at the very tip and suggest that wall materials are deposited at the polar axis, consistent with the initial model of exocytosis (Lancelle and Hepler, 1992). Experiments using tobacco (Nicotiana tabacum) pollen and a receptor-like kinase fused to GFP also indicate that exocytosis occurs largely at the apical polar axis (Lee et al., 2008).Many researchers argue that apical actin is critical for exocytosis (Lee et al., 2008; Cheung et al., 2010; Qin and Yang, 2011; Yan and Yang, 2012). More specifically, recent work suggests that the fringe participates in targeting vesicles and thereby contributes to changes in growth direction (Kroeger et al., 2009; Bou Daher and Geitmann, 2011; Dong et al., 2012). In this article, using three different inhibitors, namely brefeldin A (BFA), LatB, and potassium cyanide (KCN), we test the hypothesis that polarized pectin deposition in pollen tubes requires the actin fringe. Our data show that during normal growth, pectin deposition is focused to the apex along the polar axis of the tube. However, when growth is modulated, different end points arise, depending on the inhibitor. With BFA, exocytosis stops completely, and the fringe disappears, with the appearance of an actin aggregate at the base of the clear zone. LatB, as shown previously (Vidali et al., 2009), incompletely degrades the actin fringe and leaves a rim of F-actin around the apical dome. Here, we show that, in the presence of LatB, pectin deposition continues, with the focus of this activity shifting in position frequently as the slowly elongating pollen tube changes direction. With KCN, the actin fringe degrades completely, but exocytosis continues and becomes depolarized, with pectin deposits now occurring across a wide arc of the apical dome. This dome often swells as deposition continues, only stopping once normal growth resumes. Taken together, these results support a role for the actin fringe in controlling the polarity of growth in the lily pollen tube.  相似文献   

14.
15.
Fumarylacetoacetate hydrolase (FAH) hydrolyzes fumarylacetoacetate to fumarate and acetoacetate, the final step in the tyrosine (Tyr) degradation pathway that is essential to animals. Deficiency of FAH in animals results in an inborn lethal disorder. However, the role for the Tyr degradation pathway in plants remains to be elucidated. In this study, we isolated an Arabidopsis (Arabidopsis thaliana) short-day sensitive cell death1 (sscd1) mutant that displays a spontaneous cell death phenotype under short-day conditions. The SSCD1 gene was cloned via a map-based cloning approach and found to encode an Arabidopsis putative FAH. The spontaneous cell death phenotype of the sscd1 mutant was completely eliminated by further knockout of the gene encoding the putative homogentisate dioxygenase, which catalyzes homogentisate into maleylacetoacetate (the antepenultimate step) in the Tyr degradation pathway. Furthermore, treatment of Arabidopsis wild-type seedlings with succinylacetone, an abnormal metabolite caused by loss of FAH in the Tyr degradation pathway, mimicked the sscd1 cell death phenotype. These results demonstrate that disruption of FAH leads to cell death in Arabidopsis and suggest that the Tyr degradation pathway is essential for plant survival under short-day conditions.Programmed cell death (PCD) has been defined as a sequence of genetically regulated events that lead to the elimination of specific cells, tissues, or whole organs (Lockshin and Zakeri, 2004). In plants, PCD is essential for developmental processes and defense responses (Dangl et al., 1996; Greenberg, 1996; Durrant et al., 2007). One well-characterized example of plant PCD is the hypersensitive response occurring during incompatible plant-pathogen interactions (Lam, 2004), which results in cell death to form visible lesions at the site of infection by an avirulent pathogen and consequently limits the pathogen spread (Morel and Dangl, 1997).To date, a large number of mutants that display spontaneous cell death lesions have been identified in barley (Hordeum vulgare), maize (Zea mays), rice (Oryza sativa), and Arabidopsis (Arabidopsis thaliana; Marchetti et al., 1983; Wolter et al., 1993; Dietrich et al., 1994; Gray et al., 1997). Because lesions form in the absence of pathogen infection, these mutants have been collectively termed as lesion-mimic mutants. Many genes with regulatory roles in PCD and defense responses, including LESION SIMULATING DISEASE1, ACCELERATED CELL DEATH11, and VASCULAR ASSOCIATED DEATH1, have been cloned and characterized (Dietrich et al., 1997; Brodersen et al., 2002; Lorrain et al., 2004).The appearance of spontaneous cell death lesions in some lesion-mimic mutants is dependent on photoperiod. For example, the Arabidopsis mutant lesion simulating disease1 and myoinositol-1-phosphate synthase1 show lesions under long days (LD; Dietrich et al., 1994; Meng et al., 2009), whereas the lesion simulating disease2, lesion initiation1, enhancing RPW8-mediated HR-like cell death1, and lag one homolog1 display lesions under short days (SD; Dietrich et al., 1994; Ishikawa et al., 2003; Wang et al., 2008; Ternes et al., 2011).Blockage of some metabolic pathways in plants may cause cell death and result in lesion formation. For example, the lesion-mimic phenotypes in the Arabidopsis mutants lesion initiation2 and accelerated cell death2 and the maize mutant lesion mimic22 result from an impairment of porphyrin metabolism (Hu et al., 1998; Ishikawa et al., 2001; Mach et al., 2001). Deficiency in fatty acid, sphingolipid, and myoinositol metabolism also causes cell death in Arabidopsis (Mou et al., 2000; Liang et al., 2003; Wang et al., 2008; Meng et al., 2009; Donahue et al., 2010; Berkey et al., 2012).Tyr degradation is an essential five-step pathway in animals (Lindblad et al., 1977). First, Tyr aminotransferase catalyzes the conversion of Tyr into 4-hydroxyphenylpyruvate, which is further transformed into homogentisate by 4-hydroxyphenylpyruvate dioxygenase. Through the sequential action of homogentisate dioxygenase (HGO), maleylacetoacetate isomerase (MAAI), and fumarylacetoacetate hydrolase (FAH), homogentisate is catalyzed to generate fumarate and acetoacetate (Lindblad et al., 1977). Blockage of this pathway in animals results in metabolic disorder diseases (Lindblad et al., 1977; Ruppert et al., 1992; Grompe et al., 1993). For example, human FAH deficiency causes hereditary tyrosinemia type I (HT1), an inborn lethal disease (St-Louis and Tanguay, 1997). Although the homologous genes putatively encoding these enzymes exist in plants (Dixon et al., 2000; Lopukhina et al., 2001; Dixon and Edwards, 2006), it is unclear whether this pathway is essential for plant growth and development.In this study, we report the isolation and characterization of a recessive short-day sensitive cell death1 (sscd1) mutant in Arabidopsis. Map-based cloning of the corresponding gene revealed that SSCD1 encodes the Arabidopsis putative FAH. Further knockout of the gene encoding the Arabidopsis putative HGO completely eliminated the spontaneous cell death phenotype in the sscd1 mutant. Furthermore, we found that treatment of Arabidopsis wild-type seedlings with succinylacetone, an abnormal metabolite caused by loss of FAH in the Tyr degradation pathway (Lindblad et al., 1977), is able to mimic the sscd1 cell death phenotype. These results demonstrate that disruption of FAH leads to cell death in Arabidopsis and suggest that the Tyr degradation pathway is essential for plant survival under SD.  相似文献   

16.
In plants, K transporter (KT)/high affinity K transporter (HAK)/K uptake permease (KUP) is the largest potassium (K) transporter family; however, few of the members have had their physiological functions characterized in planta. Here, we studied OsHAK5 of the KT/HAK/KUP family in rice (Oryza sativa). We determined its cellular and tissue localization and analyzed its functions in rice using both OsHAK5 knockout mutants and overexpression lines in three genetic backgrounds. A β-glucuronidase reporter driven by the OsHAK5 native promoter indicated OsHAK5 expression in various tissue organs from root to seed, abundantly in root epidermis and stele, the vascular tissues, and mesophyll cells. Net K influx rate in roots and K transport from roots to aerial parts were severely impaired by OsHAK5 knockout but increased by OsHAK5 overexpression in 0.1 and 0.3 mm K external solution. The contribution of OsHAK5 to K mobilization within the rice plant was confirmed further by the change of K concentration in the xylem sap and K distribution in the transgenic lines when K was removed completely from the external solution. Overexpression of OsHAK5 increased the K-sodium concentration ratio in the shoots and salt stress tolerance (shoot growth), while knockout of OsHAK5 decreased the K-sodium concentration ratio in the shoots, resulting in sensitivity to salt stress. Taken together, these results demonstrate that OsHAK5 plays a major role in K acquisition by roots faced with low external K and in K upward transport from roots to shoots in K-deficient rice plants.Potassium (K) is one of the three most important macronutrients and the most abundant cation in plants. As a major osmoticum in the vacuole, K drives the generation of turgor pressure, enabling cell expansion. In the vascular tissue, K is an important participant in the generation of root pressure (for review, see Wegner, 2014 [including his new hypothesis]). In the phloem, K is critical for the transport of photoassimilates from source to sink (Marschner, 1996; Deeken et al., 2002; Gajdanowicz et al., 2011). In addition, enhancing K absorption and decreasing sodium (Na) accumulation is a major strategy of glycophytes in salt stress tolerance (Maathuis and Amtmann, 1999; Munns and Tester, 2008; Shabala and Cuin, 2008).Plants acquire K through K-permeable proteins at the root surface. Since available K concentration in the soil may vary by 100-fold, plants have developed multiple K uptake systems for adapting to this variability (Epstein et al., 1963; Grabov, 2007; Maathuis, 2009). In a classic K uptake experiment in barley (Hordeum vulgare), root K absorption has been described as a high-affinity and low-affinity biphasic transport process (Epstein et al., 1963). It is generally assumed that the low-affinity transport system (LATS) in the roots mediates K uptake in the millimolar range and that the activity of this system is insensitive to external K concentration (Maathuis and Sanders, 1997; Chérel et al., 2014). In contrast, the high-affinity transport system (HATS) was rapidly up-regulated when the supply of exogenous K was halted (Glass, 1976; Glass and Dunlop, 1978).The membrane transporters for K flux identified in plants are generally classified into three channels and three transporter families based on phylogenetic analysis (Mäser et al., 2001; Véry and Sentenac, 2003; Lebaudy et al., 2007; Alemán et al., 2011). For K uptake, it was predicted that, under most circumstances, K transporters function as HATS, while K-permeable channels mediate LATS (Maathuis and Sanders, 1997). However, a root-expressed K channel in Arabidopsis (Arabidopsis thaliana), Arabidopsis K Transporter1 (AKT1), mediates K absorption over a wide range of external K concentrations (Sentenac et al., 1992; Lagarde et al., 1996; Hirsch et al., 1998; Spalding et al., 1999), while evidence is accumulating that many K transporters, including members of the K transporter (KT)/high affinity K transporter (HAK)/K uptake permease (KUP) family, are low-affinity K transporters (Quintero and Blatt, 1997; Senn et al., 2001), implying that functions of plant K channels and transporters overlap at different K concentration ranges.Out of the three families of K transporters, cation proton antiporter (CPA), high affinity K/Na transporter (HKT), and KT/HAK/KUP, CPA was characterized as a K+(Na+)/H+ antiporter, HKT may cotransport Na and K or transport Na only (Rubio et al., 1995; Uozumi et al., 2000), while KT/HAK/KUP were predicted to be H+-coupled K+ symporters (Mäser et al., 2001; Lebaudy et al., 2007). KT/HAK/KUP were named by different researchers who first identified and cloned them (Quintero and Blatt, 1997; Santa-María et al., 1997). In plants, the KT/HAK/KUP family is the largest K transporter family, including 13 members in Arabidopsis and 27 members in the rice (Oryza sativa) genome (Rubio et al., 2000; Mäser et al., 2001; Bañuelos et al., 2002; Gupta et al., 2008). Sequence alignments show that genes of this family share relatively low homology to each other. The KT/HAK/KUP family was divided into four major clusters (Rubio et al., 2000; Gupta et al., 2008), and in cluster I and II, they were further separated into A and B groups. Genes of cluster I or II likely exist in all plants, cluster III is composed of genes from both Arabidopsis and rice, while cluster IV includes only four rice genes (Grabov, 2007; Gupta et al., 2008).The functions of KT/HAK/KUP were studied mostly in heterologous expression systems. Transporters of cluster I, such as AtHAK5, HvHAK1, OsHAK1, and OsHAK5, are localized in the plasma membrane (Kim et al., 1998; Bañuelos et al., 2002; Gierth et al., 2005) and exhibit high-affinity K uptake in the yeast Saccharomyces cerevisiae (Santa-María et al., 1997; Fu and Luan, 1998; Rubio et al., 2000) and in Escherichia coli (Horie et al., 2011). Transporters of cluster II, like AtKUP4 (TINY ROOT HAIRS1, TRH1), HvHAK2, OsHAK2, OsHAK7, and OsHAK10, could not complement the K uptake-deficient yeast (Saccharomyces cerevisiae) but were able to mediate K fluxes in a bacterial mutant; they might be tonoplast transporters (Senn et al., 2001; Bañuelos et al., 2002; Rodríguez-Navarro and Rubio, 2006). The function of transporters in clusters III and IV is even less known (Grabov, 2007).Existing data suggest that some KT/HAK/KUP transporters also may respond to salinity stress (Maathuis, 2009). The cluster I transporters of HvHAK1 mediate Na influx (Santa-María et al., 1997), while AtHAK5 expression is inhibited by Na (Rubio et al., 2000; Nieves-Cordones et al., 2010). Expression of OsHAK5 in tobacco (Nicotiana tabacum) BY2 cells enhanced the salt tolerance of these cells by accumulating more K without affecting their Na content (Horie et al., 2011).There are only scarce reports on the physiological function of KT/HAK/KUP in planta. In Arabidopsis, mutation of AtKUP2 (SHORT HYPOCOTYL3) resulted in a short hypocotyl, small leaves, and a short flowering stem (Elumalai et al., 2002), while a loss-of-function mutation of AtKUP4 (TRH1) resulted in short root hairs and a loss of gravity response in the root (Rigas et al., 2001; Desbrosses et al., 2003; Ahn et al., 2004). AtHAK5 is the only system currently known to mediate K uptake at concentrations below 0.01 mm (Rubio et al., 2010) and provides a cesium uptake pathway (Qi et al., 2008). AtHAK5 and AtAKT1 are the two major physiologically relevant molecular entities mediating K uptake into roots in the range between 0.01 and 0.05 mm (Pyo et al., 2010; Rubio et al., 2010). AtAKT1 may contribute to K uptake within the K concentrations that belong to the high-affinity system described by Epstein et al. (1963).Among all 27 members of the KT/HAK/KUP family in rice, OsHAK1, OsHAK5, OsHAK19, and OsHAK20 were grouped in cluster IB (Gupta et al., 2008). These four rice HAK members share 50.9% to 53.4% amino acid identity with AtHAK5. OsHAK1 was expressed in the whole plant, with maximum expression in roots, and was up-regulated by K deficiency; it mediated high-affinity K uptake in yeast (Bañuelos et al., 2002). In this study, we examined the tissue-specific localization and the physiological functions of OsHAK5 in response to variation in K supply and to salt stress in rice. By comparing K uptake and translocation in OsHAK5 knockout (KO) mutants and in OsHAK5-overexpressing lines with those in their respective wild-type lines supplied with different K concentrations, we found that OsHAK5 not only mediates high-affinity K acquisition but also participates in root-to-shoot K transport as well as in K-regulated salt tolerance.  相似文献   

17.
18.
19.
Transglutaminases (TGases) are ubiquitous enzymes that take part in a variety of cellular functions. In the pollen tube, cytoplasmic TGases are likely to be involved in the incorporation of primary amines at selected peptide-bound glutamine residues of cytosolic proteins (including actin and tubulin), while cell wall-associated TGases are believed to regulate pollen tube growth. Using immunological probes, we identified TGases associated with different subcellular compartments (cytosol, membranes, and cell walls). Binding of cytosolic TGase to actin filaments was shown to be Ca2+ dependent. The membrane TGase is likely associated with both Golgi-derived structures and the plasma membrane, suggesting a Golgi-based exocytotic delivery of TGase. Association of TGase with the plasma membrane was also confirmed by immunogold transmission electron microscopy. Immunolocalization of TGase indicated that the enzyme was present in the growing region of pollen tubes and that the enzyme colocalizes with cell wall markers. Bidimensional electrophoresis indicated that different TGase isoforms were present in distinct subcellular compartments, suggesting either different roles or different regulatory mechanisms of enzyme activity. The application of specific inhibitors showed that the distribution of TGase in different subcellular compartments was regulated by both membrane dynamics and cytoskeleton integrity, suggesting that delivery of TGase to the cell wall requires the transport of membranes along cytoskeleton filaments. Taken together, these data indicate that a cytoplasmic TGase interacts with the cytoskeleton, while a different TGase isoform, probably delivered via a membrane/cytoskeleton-based transport system, is secreted in the cell wall of pear (Pyrus communis) pollen tubes, where it might play a role in the regulation of apical growth.Transglutaminases (TGases [EC 2.3.2.13]; protein-Gln γ-glutamyltransferase) are a family of ubiquitous Ca2+-activated enzymes that are involved in animal cell morphogenesis and differentiation, apoptosis, cell death, inflammation, cell migration, and wound healing (Griffin et al., 2002; Mehta et al., 2006; Beninati et al., 2009). TGases are associated with different subcellular compartments, such as cytosol, plasma membrane, nucleus, mitochondria, and extracellular matrix. The specific localization of TGases is likely to determine both the biochemical activity and the type of proteins and/or substrates with which TGases react (Park et al., 2010). The distribution profile of TGase is affected by Ca2+, since the enzyme is preferentially associated with the lysosome compartment of liver cells in the absence of Ca2+ (Juprelle-Soret et al., 1984).TGase was initially detected in association with the cytosol, with the particulate (probably the microsomal) fraction (Birckbichler et al., 1976), and with the nucleus of animal cells (Remington and Russell, 1982). The association of TGase with the plasma membrane was related to its activity in promoting cell adhesion and to the interaction of cells with the extracellular matrix, while the presence of TGase in the nucleus is likely related to cell apoptosis (Griffin et al., 2002). How TGase is delivered to its final destination in animal cells remains to be clarified. Since the cytoskeleton is essential for the correct positioning of proteins in the cells, this interplay has often been studied in terms of potential substrates of TGase activity (Griffin et al., 2002). For example, the TGase-mediated incorporation of polyamines (PAs) stimulates actin polymerization (Takashi, 1988; Griffin et al., 2002). TGase was also found to associate with myosin in stress fibers of vascular smooth cells (Chowdhury et al., 1997). The association between TGase and microtubules (MTs) was initially studied in view of the importance of MTs in Alzheimer’s disease (Griffin et al., 2002), whereas the dynamics of MTs is also likely to be controlled by TGase (Al-Jallad et al., 2011). Interestingly, MTs are also a substrate of TGase activity in cells committed to apoptosis (Piredda et al., 1999). TGase was also shown to posttranslationally modify MT-associated proteins such as tau (Griffin et al., 2002).Information about the localization and function of TGases in plant cells is limited. Following the early evidence of an enzyme-based incorporation of PAs in plants (Serafini-Fracassini et al., 1988), a number of reports described the presence and role of TGase in nonphotosynthetic/photosynthetic tissues and in isolated chloroplasts (Serafini-Fracassini and Del Duca, 2008, and refs. therein). Attempts have also been made to examine the differences and similarities between plant and animal TGases. For example, a tobacco (Nicotiana tabacum) TGase was proposed to be involved in the programmed cell death (PCD) of the flower corolla (Della Mea et al., 2007); in such a case, TGase is likely to be released into the cell wall by a Golgi vesicle-based transport. Plant TGases might also be involved in protection against viruses (Del Duca et al., 2007) and in the self-incompatibility (SI) response involving pollen and stigma during sexual reproduction (Del Duca et al., 2010). Recently, different TGase isoforms were detected in meristematic apices of Jerusalem artichoke (Helianthus tuberosus) tuber sprouts (Beninati et al., 2013).The pollen tube is a widely investigated tip-growing plant cell (Lee and Yang, 2008). Studies are generally aimed at clarifying the many aspects related either to its growth or to rejection by the stigma/style. Early evidence for a role of PAs during pollen tube emergence (Bagni et al., 1981) was confirmed through the detection of PA binding via a Ca2+-activated TGase activity (Del Duca et al., 1997) and later by the identification of actin and tubulin as substrates of purified pollen TGase (Del Duca et al., 2009). In pollen, the enzyme affected the polymerization state and activity of actin filaments (AFs) and MTs (Del Duca et al., 2009) and existed as both soluble and cell wall associated (Di Sandro et al., 2010). Visualization of fluorescently labeled TGase products indicated that the cross-linking activity of TGase occurred at the apex of pollen tubes, in a basal region close to the pollen grain and within the pollen grain itself (Iorio et al., 2008). The enzyme was found as a soluble cytoplasmic form likely involved in the regulation of unspecified physiological processes (possibly associated with the cytoskeleton; Del Duca et al., 2009).Although the association of pollen TGases with organelles/vesicles has not been reported, an extracellular form of a Ca2+-dependent TGase was shown to be involved in pollen tube growth (likely as a modulator of cell wall building and strengthening). Moreover, pollen TGase was secreted in the incubation medium during germination, where it might catalyze the cross linking of PAs with secreted proteins (Di Sandro et al., 2010). This suggests that pollen TGase may be secreted through a vesicle-based mechanism. Finally, a TGase activity was also observed in planta, consistent with a possible role of TGase during tube migration through the style (Di Sandro et al., 2010) or in the SI response of pollen tubes (Del Duca et al., 2010).The pollen tube is an excellent model to study how a given plant protein is either secreted or delivered to its final destination. Although we know that actin and tubulin are substrates of TGase activity, and that the active enzyme is located in the cell wall and released outside, how TGase is distributed in the cells and how this process is dependent on cytoskeleton and membrane dynamics remain unknown. Here, we wanted to study in detail the localization and distribution of TGase in growing pollen tubes of pear (Pyrus communis) in relation to both cytoskeleton and membrane dynamics. The aim was to shed light on the mechanism by which TGase is transported and secreted, a process that is still not well understood even in animal cells. Specific antibodies that cross react with the TGase of pollen tubes were used to localize the enzyme in different membrane compartments and in the cell wall. The use of specific inhibitors indicated that the delivery of extracellular TGase is dependent on both AFs and membrane dynamics. Analysis by bidimensional electrophoresis (2-DE) showed that distinct TGase isoforms are associated with different cell compartments, suggesting that TGase might be differently regulated according to its position in the cell. Together, these data may contribute to our understanding of the mechanisms underlying pollen tube growth, an essential aspect of fertilization processes.  相似文献   

20.
In Solanaceae, the self-incompatibility S-RNase and S-locus F-box interactions define self-pollen recognition and rejection in an S-specific manner. This interaction triggers a cascade of events involving other gene products unlinked to the S-locus that are crucial to the self-incompatibility response. To date, two essential pistil-modifier genes, 120K and High Top-Band (HT-B), have been identified in Nicotiana species. However, biochemistry and genetics indicate that additional modifier genes are required. We recently reported a Kunitz-type proteinase inhibitor, named NaStEP (for Nicotiana alata Stigma-Expressed Protein), that is highly expressed in the stigmas of self-incompatible Nicotiana species. Here, we report the proteinase inhibitor activity of NaStEP. NaStEP is taken up by both compatible and incompatible pollen tubes, but its suppression in Nicotiana spp. transgenic plants disrupts S-specific pollen rejection; therefore, NaStEP is a novel pistil-modifier gene. Furthermore, HT-B levels within the pollen tubes are reduced when NaStEP-suppressed pistils are pollinated with either compatible or incompatible pollen. In wild-type self-incompatible N. alata, in contrast, HT-B degradation occurs preferentially in compatible pollinations. Taken together, these data show that the presence of NaStEP is required for the stability of HT-B inside pollen tubes during the rejection response, but the underlying mechanism is currently unknown.To avoid low-fitness progeny, many plants have developed a cell-cell interaction mechanism to promote outcrossing, through the recognition and discrimination of both self and nonself pollen. This recognition system is controlled by the highly polymorphic self-incompatibility S-locus, which determines pollination specificity in both the pollen and pistil. Pollen is rejected when male and female S-haplotypes coincide (de Nettancourt, 1977, 2001; Franklin et al., 1995).In Solanaceae, Plantaginaceae, and Rosaceae, the S-locus product in the pistil is an extracellular glycoprotein named S-RNase (Anderson et al., 1986; McClure et al., 1989). During pollination, S-RNase is taken up by both compatible and incompatible pollen tubes (Luu et al., 2000) and targeted to a vacuole (Goldraij et al., 2006). In the later stages of an incompatible cross, the S-RNase-containing vacuole is disrupted and the S-RNases are released to the pollen tube cytoplasm, where RNA degradation can occur (McClure et al., 2011).The S-pollen gene encodes an SLF or SFB (SLF/SFB; for S-locus F-box) protein, which is a member of the F-box protein family (Entani et al., 2003; Sijacic et al., 2004). In vitro binding assays show that PiSLF in Petunia inflata physically interacts with S-RNases, although this interaction is stronger with nonself S-RNases than with self S-RNases (Hua and Kao, 2006). Additional protein-protein interaction assays suggest that SLF/SFB may be a component of an SCF (for Skp1-Cullin1-F-box) or SCF-like complex (Qiao et al., 2004; Hua and Kao, 2006). Notably, data from Zhao et al. (2010) in Petunia hybrida show that reduction of PhSSK1 (for P. hybrida SLF-interacting Skp-like1) and its Antirrhinum hispanicum ortholog, AhSSK1, is also required for cross-pollen compatibility.Although S-RNase and SLF/SFB define pollen rejection S-specificity, modifier genes unlinked to the S-locus are required for self-incompatibility (SI; Martin, 1968; Ai et al., 1991; Murfett et al., 1996; Tsukamoto et al., 1999).To date, only two pistil-modifier genes have been identified: High Top-Band (HT-B) and 120K. In Nicotiana spp., HT-B is an 8.6-kD acidic protein with a domain consisting of 20 Asn and Asp residues toward its C terminus (McClure et al., 1999; Kondo and McClure, 2008). Loss-of-function assays prove HT-B to be essential for pollen rejection in Nicotiana spp., Solanum spp., and Petunia spp. (McClure et al., 1999; Kondo et al., 2002; O’Brien et al., 2002; Sassa and Hirano, 2006; Puerta et al., 2009), although it is not expressed in SI Solanum habrochaites, prompting the speculation that in this species a related gene, HT-A, may function as a substitute (Covey et al., 2010). Immunolocalization shows that HT-B is readily taken up by pollen tubes during pollination. Its steady-state levels decrease slightly in pollen tubes from incompatible pollinations. However, in compatible crosses, HT-B levels decrease 75% to 97%, probably as a result of protein degradation (Goldraij et al., 2006).120K is a style-specific 120-kD arabinogalactan protein (Schultz et al., 1997) that is taken up by pollen tubes (Lind et al., 1996) and appears to be associated with S-RNase-containing vacuoles (Goldraij et al., 2006). 120K forms complexes with S-RNases and other proteins (Cruz-Garcia et al., 2005) in vitro, and suppression of 120K expression prevents S-specific pollen rejection (Hancock et al., 2005). Protein-protein interaction assays demonstrate that 120K interacts with the pollen-specific protein NaPCCP (a pollen C2 domain-containing protein), a protein that binds phosphatidylinositol 3-phosphate and is associated with the pollen tube endomembrane system (Lee et al., 2008, 2009).Two models have been proposed to explain pollen rejection in Solanaceae. (1) The S-RNase degradation model (Hua and Kao, 2006; Hua et al., 2007, 2008; Kubo et al., 2010) focuses on S-RNase-SLF interactions that bring about preferential nonself S-RNase degradation. In this model, strong nonself S-RNase-SLF interactions lead to the degradation of nonself S-RNases by the ubiquitin-26S proteasome system, allowing pollen tubes to escape from its cytotoxic effect. Weak self S-RNase-SLF interactions, in contrast, permit the persistence of sufficient free S-RNase that pollen tube RNA is degraded, resulting in self-pollen rejection. Notably, by functional and protein-protein interaction assays in Petunia spp., Kubo et al. (2010) found at least three types of divergent SLF proteins encoded at the S-locus, each recognizing a subgroup of nonself S-RNases. The authors proposed the collaborative nonself recognition model, where multiple SLF proteins interact with nonself S-RNases to protect nonself pollen from degradation (Kubo et al., 2010). (2) The compartmentalization model incorporates the observations that pollen tubes internalize both self and nonself S-RNases and targets them to vacuoles and that HT-B is degraded in compatible crosses but is stable in incompatible crosses (Goldraij et al., 2006). In incompatible crosses, the S-RNase-containing vacuoles are ultimately disrupted and S-RNases are released to the cytoplasm, where they degrade RNA, leading to rejection of self-pollen. In compatible crosses, the integrity of the S-RNase-containing vacuoles is preserved, allowing pollen tube growth to continue. Thus, in this model, self or nonself S-RNase-SLF interactions determine the specificity of pollen rejection indirectly.Biochemical and genetic data show that pistil-modifier genes apart from HT-B and 120K are required for SI. We recently described NaStEP (for N. alata Stigma-Expressed Protein), an abundant, pistil-specific stigma protein found in SI Nicotiana spp. (Busot et al., 2008). Its abundance in SI species made NaStEP a strong modifier gene candidate. Here, we demonstrate that NaStEP is taken up by pollen tubes, has subtilisin inhibitory activity, and that suppressing its expression in transgenic hybrids disrupts pollen rejection. Moreover, when NaStEP-suppressed hybrids are pollinated, HT-B protein is degraded in both compatible and incompatible pollen tubes, while in wild-type SI N. alata, HT-B is preferentially stabilized in incompatible pollen tubes.  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