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 156 毫秒
1.
The negatively charged lipid phosphatidylglycerol (PG) constitutes up to 10% of total lipids in photosynthetic membranes, and its deprivation in cyanobacteria is accompanied by chlorophyll (Chl) depletion. Indeed, radioactive labeling of the PG-depleted ΔpgsA mutant of Synechocystis sp. strain PCC 6803, which is not able to synthesize PG, proved the inhibition of Chl biosynthesis caused by restriction on the formation of 5-aminolevulinic acid and protochlorophyllide. Although the mutant accumulated chlorophyllide, the last Chl precursor, we showed that it originated from dephytylation of existing Chl and not from the block in the Chl biosynthesis. The lack of de novo-produced Chl under PG depletion was accompanied by a significantly weakened biosynthesis of both monomeric and trimeric photosystem I (PSI) complexes, although the decrease in cellular content was manifested only for the trimeric form. However, our analysis of ΔpgsA mutant, which lacked trimeric PSI because of the absence of the PsaL subunit, suggested that the virtual stability of monomeric PSI is a result of disintegration of PSI trimers. Interestingly, the loss of trimeric PSI was accompanied by accumulation of monomeric PSI associated with the newly synthesized CP43 subunit of photosystem II. We conclude that the absence of PG results in the inhibition of Chl biosynthetic pathway, which impairs synthesis of PSI, despite the accumulation of chlorophyllide released from the degraded Chl proteins. Based on the knowledge about the role of PG in prokaryotes, we hypothesize that the synthesis of Chl and PSI complexes are colocated in a membrane microdomain requiring PG for integrity.Photosynthetic membrane of oxygenic phototrophs has a unique lipid composition that has been conserved during billions of years of evolution from cyanobacteria and algae to modern higher plants. With no known exception, this membrane system always contains the uncharged glycolipids monogalactosyldiacylglycerol and digalactosyldiacylglycerol (DGDG) as well as the negatively charged lipids sulfoquinovosyldiacylglycerol (SQDG) and phosphatidylglycerol (PG; Murata and Siegenthaler, 1998). Interestingly, it seems that PG is the only lipid completely essential for the oxygenic photosynthesis. The loss of DGDG has only a mild impact on the cyanobacterial cell (Awai et al., 2007), and as shown recently in the cyanobacterium Synechocystis sp. strain PCC 6803, both galactolipids can be in fact replaced by glucolipids (Awai et al., 2014). SQDG and PG are only minor lipid components, each accounting for 5% to 12% of total lipids (Murata and Siegenthaler, 1998). SQDG is dispensable, although its lack results in various defects (Yu et al., 2002; Aoki et al., 2004), but PG plays an essential role in both cyanobacterial cells and plant chloroplasts (Hagio et al., 2000; Babiychuk et al., 2003).The critical role of PG has been mostly connected to the function of PSII. In both cyanobacteria and plants, lack of PG impairs the stability of PSII complexes and the electron transport between primary and secondary quinone acceptors inside the PSII reaction center. As shown in Synechocystis sp., PG molecules stabilize PSII dimers and facilitate the binding of inner antenna protein CP43 within the PSII core (Laczkó-Dobos et al., 2008). Indeed, according to the PSII crystal structure, two PG molecules are located at the interface between CP43 and the D1-D2 heterodimer (Guskov et al., 2009). As a consequence, the PG depletion inhibits and destabilizes PSII complexes and also, impairs assembly of new PSII complexes, although all PSII subunits are still synthesized in the cell (Laczkó-Dobos et al., 2008).Despite the fact that the vital link between PG and PSII is now well established, the phenotypic traits of PG-depleted cells signal that there are other sites in the photosynthetic membrane requiring strictly PG molecules. In Synechocystis sp., lack of PG triggers rapid loss of trimeric PSI complexes (Domonkos et al., 2004; Sato et al., 2004), and because PSI complexes bind more than 80% of chlorophyll (Chl) in the Synechocystis sp. cell, the PG depletion is accompanied by a characteristic Chl bleaching (Domonkos et al., 2004). However, the reasons for this symptom are still unclear. Chl metabolism is tightly coordinated with synthesis, assembly, and degradation of photosystem complexes (for review, see Komenda et al., 2012b; Sobotka, 2014), and we have shown recently that the PSI complexes are the main sink for de novo Chl produced in cyanobacteria (Kopečná et al., 2012). Given the drastic decrease in PSI content in the PG-depleted cells, Chl biosynthesis must be directly or indirectly affected after the PG concentration in membranes drops below a critical value. Although it was recently suggested that galactolipid and Chl biosyntheses are coregulated during chloroplast biogenesis (Kobayashi et al., 2014), a response of the Chl biosynthetic pathway to the altered lipid content has not been examined.To investigate Chl metabolism during PG starvation, we used the Synechocystis sp. ΔpgsA mutant, which is unable to synthesize PG (Hagio et al., 2000). The advantage of using the ΔpgsA strain is in its ability to utilize exogenous PG from growth medium, which allows monitoring of phenotypic changes from a wild type-like situation to completely PG-depleted cells. Chl biosynthesis shares the same metabolic pathway with heme and other tetrapyrroles. At the beginning of tetrapyrrole biosynthesis, the initial precursor, 5-aminolevulinic acid (ALA), is made from Glu through glutamyl-tRNA and subsequently converted in several steps to protoporphyrin IX. The pathway branches at the point where protoporphyrin IX is chelated by magnesium to produce Mg-protoporphyrin IX, the first intermediate on the Chl branch. This step is catalyzed by Mg-chelatase, a multisubunit enzyme that associates relatively weakly with the membrane; however, all following enzymes downward in the pathway are almost exclusively bound to membranes (Masuda and Fujita, 2008; Kopečná et al., 2012). The last enzyme of the Chl pathway, Chl synthase, is an integral membrane protein that attaches a phytyl chain to the last intermediate chlorophyllide (Chlide) to finalize Chl formation (Oster et al., 1997; Addlesee et al., 2000). According to current views, Chl synthase should also be involved in reutilization of Chl molecules from degraded Chl-binding proteins, which includes dephytylation and phytylation of Chl molecules with Chlide as an intermediate (Vavilin and Vermaas, 2007).In this study, we show a complex impact of PG deficiency on Chl metabolism. The lack of PG inhibited Chl biosynthesis at the two different steps: first, it drastically reduced formation of the initial precursor ALA, and second, it impaired the Mg-protoporphyrin methyl ester IX (MgPME) cyclase enzyme catalyzing synthesis of protochlorophyllide (Pchlide). The diminished rate of Chl formation was accompanied by impaired synthesis of both trimeric and monomeric PSI complexes and accumulation of a PSI monomer associated with the CP43 subunit of PSII. We also showed that the PG-depleted cells accumulated Chlide, originating from dephytylation of existing Chl, which suggests an inability to reutilize Chl for the PSI synthesis. We discuss a scenario that the Chl biosynthesis and synthesis of core PSI subunits are colocated in PG-enriched membrane microdomains.  相似文献   

2.
State transitions in photosynthesis provide for the dynamic allocation of a mobile fraction of light-harvesting complex II (LHCII) to photosystem II (PSII) in state I and to photosystem I (PSI) in state II. In the state I-to-state II transition, LHCII is phosphorylated by STN7 and associates with PSI to favor absorption cross-section of PSI. Here, we used Arabidopsis (Arabidopsis thaliana) mutants with defects in chlorophyll (Chl) b biosynthesis or in the chloroplast signal recognition particle (cpSRP) machinery to study the flexible formation of PS-LHC supercomplexes. Intriguingly, we found that impaired Chl b biosynthesis in chlorina1-2 (ch1-2) led to preferentially stabilized LHCI rather than LHCII, while the contents of both LHCI and LHCII were equally depressed in the cpSRP43-deficient mutant (chaos). In view of recent findings on the modified state transitions in LHCI-deficient mutants (Benson et al., 2015), the ch1-2 and chaos mutants were used to assess the influence of varying LHCI/LHCII antenna size on state transitions. Under state II conditions, LHCII-PSI supercomplexes were not formed in both ch1-2 and chaos plants. LHCII phosphorylation was drastically reduced in ch1-2, and the inactivation of STN7 correlates with the lack of state transitions. In contrast, phosphorylated LHCII in chaos was observed to be exclusively associated with PSII complexes, indicating a lack of mobile LHCII in chaos. Thus, the comparative analysis of ch1-2 and chaos mutants provides new evidence for the flexible organization of LHCs and enhances our understanding of the reversible allocation of LHCII to the two photosystems.In oxygenic photosynthesis, PSII and PSI function in series to convert light energy into the chemical energy that fuels multiple metabolic processes. Most of this light energy is captured by the chlorophyll (Chl) and carotenoid pigments in the light-harvesting antenna complexes (LHCs) that are peripherally associated with the core complexes of both photosystems (Wobbe et al., 2016). However, since the two photosystems exhibit different absorption spectra (Nelson and Yocum, 2006; Nield and Barber, 2006; Qin et al., 2015), PSI or PSII is preferentially excited under naturally fluctuating light intensities and qualities. To optimize photosynthetic electron transfer, the excitation state of the two photosystems must be rebalanced in response to changes in lighting conditions. To achieve this, higher plants and green algae require rapid and precise acclimatory mechanisms to adjust the relative absorption cross-sections of the two photosystems.To date, the phenomenon of state transitions is one of the well-documented short-term acclimatory mechanisms. It allows a mobile portion of the light-harvesting antenna complex II (LHCII) to be allocated to either photosystem, depending on the spectral composition and intensity of the ambient light (Allen and Forsberg, 2001; Rochaix, 2011; Goldschmidt-Clermont and Bassi, 2015; Gollan et al., 2015). State transitions are driven by the redox state of the plastoquinone (PQ) pool (Vener et al., 1997; Zito et al., 1999). When PSI is preferentially excited (by far-red light), the PQ pool is oxidized and all the LHCII is associated with PSII. This allocation of antenna complexes is defined as state I. When light conditions (blue/red light or low light) favor exciton trapping of PSII, the transition from state I to state II occurs. The over-reduced PQ pool triggers the activation of the membrane-localized Ser-Thr kinase STN7, which phosphorylates an N-terminal Thr on each of two major LHCII proteins, LHCB1 and LHCB2 (Allen, 1992; Bellafiore et al., 2005; Shapiguzov et al., 2016). Phosphorylation of LHCII results in the dissociation of LHCII from PSII and triggers its reversible relocation to PSI (Allen, 1992; Rochaix, 2011). Conversely, when the PQ pool is reoxidized, STN7 is inactivated and the constitutively active, thylakoid-associated phosphatase TAP38/PPH1 dephosphorylates LHCII, which then reassociates with PSII (Pribil et al., 2010; Shapiguzov et al., 2010). The physiological significance of state transitions has been demonstrated by the reduction in growth rate seen in the stn7 knock-out mutant under fluctuating light conditions (Bellafiore et al., 2005; Tikkanen et al., 2010).The canonical state transitions model implies spatial and temporal regulation of the allocation of LHC between the two spatially segregated photosystems (Dekker and Boekema, 2005). PSII-LHCII supercomplexes are organized in a tightly packed form in the stacked grana regions of thylakoid membranes, while PSI-LHCI supercomplexes are mainly localized in the nonstacked stromal lamellae and grana margin regions (Dekker and Boekema, 2005; Haferkamp et al., 2010). It has been proposed that, in the grana margin regions, which harbor LHCII and both photosystems, LHCII can migrate rapidly between them (Albertsson et al., 1990; Albertsson, 2001). This idea is supported by the recent discovery of mega complexes containing both photosystems in the grana margin regions (Yokono et al., 2015). Furthermore, phosphorylation of LHCII was found to increase not only the amount of PSI found in the grana margin region of thylakoid membranes (Tikkanen et al., 2008a), but also to modulate the pattern of PSI-PSII megacomplexes under changing light conditions (Suorsa et al., 2015). Nonetheless, open questions remain in relation to the physiological significance of the detection of phosphorylated LHCII in all thylakoid regions, even under the constant light conditions (Grieco et al., 2012; Leoni et al., 2013; Wientjes et al., 2013), although LHCII phosphorylation has been shown to modify the stacking of thylakoid membranes (Chuartzman et al., 2008; Pietrzykowska et al., 2014).State I-to-state II transition is featured by the formation of LHCII-PSI-LHCI supercomplexes, in which LHCII favors the light-harvesting capacity of PSI. Recently, LHCII-PSI-LHCI supercomplexes have been successfully isolated and purified using various detergents (Galka et al., 2012; Drop et al., 2014; Crepin and Caffarri, 2015) or a styrene-maleic acid copolymer (Bell et al., 2015). These findings yielded further insights into the reorganization of supercomplexes associated with state transitions, and it was suggested that phosphorylation of LHCB2 rather than LHCB1 is the essential trigger for the formation of state transition supercomplexes (Leoni et al., 2013; Pietrzykowska et al., 2014; Crepin and Caffarri, 2015; Longoni et al., 2015). Furthermore, characterization of mutants deficient in individual PSI core subunits indicates that PsaH, L, and I are required for docking of LHCII at PSI (Lunde et al., 2000; Zhang and Scheller, 2004; Kouril et al., 2005; Plöchinger et al., 2016).Recently, the state transition capacity has been characterized in the Arabidopsis (Arabidopsis thaliana) mutants with missing LHCI components. Although the Arabidopsis knock-out mutants lacking one of the four LHCI proteins (LHCA1-4) showed enhanced accumulation of LHCII-PSI complexes, the absorption cross-section of PSI under state II conditions was still compromised in the lhca1-4 mutants, and it is suggested that LHCI mediates the detergent-sensitive interaction between ‘extra LHCII’ and PSI (Benson et al., 2015; Grieco et al., 2015). Furthermore, the Arabidopsis mutant ΔLhca lacking all LHCA1-4 proteins was shown to be compensated for the deficiency of LHCI by binding LHCII under state II conditions (Bressan et al., 2016). In spite of this finding, the significant reduction in the absorption cross-section of PSI was still observed in the ΔLhca mutant, suggesting a substantial role of LHCI in light absorption under canopy conditions (Bressan et al., 2016). However, these findings emphasize the acclimatory function of state transitions in balancing light absorption capacity between the two photosystems by modifying their relative antenna size and imply the dynamic and variable organization of PS-LHC supercomplexes.LHC proteins are encoded by the nuclear Lhc superfamily (Jansson, 1994). The biogenesis of LHCs includes the cytoplasmic synthesis of the LHC precursor proteins, their translocation into chloroplasts via the TOC/TIC complex, and their posttranslational targeting and integration into the thylakoid membranes by means of the chloroplast signal recognition particle (cpSRP) machinery (Jarvis and Lopez-Juez, 2013). The posttranslational cpSRP-dependent pathway for the final translocation of LHC proteins into the thylakoid membrane includes interaction of cpSRP43 with LHC apo-proteins and recruitment of cpSRP54 to form a transit complex. Then binding of this tripartite cpSRP transit complex to the SRP receptor cpFtsY follows, which supports docking of the transit complex to thylakoid membranes and its association with the LHC translocase ALB3. Ultimately, ALB3 inserts LHC apo-proteins into the thylakoid membrane (Richter et al., 2010). Importantly, stoichiometric amounts of newly synthesized Chl a and Chl b as well as carotenoid are inserted into the LHC apo-proteins by unknown mechanisms to form the functional LHCs that associate with the core complexes of both photosystems in the thylakoid membranes (Dall’Osto et al., 2015; Wang and Grimm, 2015).The first committed steps in Chl synthesis occur in the Mg branch of the tetrapyrrole biosynthesis pathway. 5-Aminolevulinic acid synthesis provides the precursor for the formation of protoporphyrin IX, which is directed into the Mg branch (Tanaka and Tanaka, 2007; Brzezowski et al., 2015). Chl synthesis ends with the conversion of Chl a to Chl b catalyzed by Chl a oxygenase (CAO; Tanaka et al., 1998; Tomitani et al., 1999). It has been hypothesized that coordination between Chl synthesis and the posttranslational cpSRP pathway is a prerequisite for the efficient integration of Chls into LHC apo-proteins.In this study, we intend to characterize the assembly of LHCs when the availability of Chl molecules or the integration of LHC apo-proteins into thylakoid membranes is limiting. To this end, we compared the assembly of LHCs and the organization of PS-LHC complexes in two different sets of Arabidopsis mutants. Firstly, we used the chlorina1-2 (ch1-2) mutant, which is defective in the CAO gene. The members of the second set of mutants carry knock-out mutations in genes involved in the chloroplast SRP pathway (Richter et al., 2010).Our studies revealed distinct accumulation of PS-LHC supercomplexes between the two sets of mutant relative to wild-type plants. In spite of the defect in synthesis of Chl b, ch1-2 retains predominantly intact PSI-LHCI supercomplexes but has strongly reduced amounts of LHCII. In contrast, the chaos (cpSRP43) mutant exhibits synchronously reduced contents of both LHCI and LHCII, which results in the accumulation of PS core complexes without accompanying LHCs. Thus, the distribution of LHCs in the thylakoid membranes of the two mutants, ch1-2 and chaos, were explored under varying light conditions with the aim of elucidating the influence of modified LHCI/LHCII antenna size on state transitions. Our results contribute to an expanding view on the variety of photosynthetic complexes, which can be observed in Arabidopsis plants with specified mutations in LHC biogenesis.  相似文献   

3.
4.
5.
Examining the proteins that plants secrete into the apoplast in response to pathogen attack provides crucial information for understanding the molecular mechanisms underlying plant innate immunity. In this study, we analyzed the changes in the root apoplast secretome of the Verticillium wilt-resistant island cotton cv Hai 7124 (Gossypium barbadense) upon infection with Verticillium dahliae. Two-dimensional differential gel electrophoresis and matrix-assisted laser desorption/ionization tandem time-of-flight mass spectrometry analysis identified 68 significantly altered spots, corresponding to 49 different proteins. Gene ontology annotation indicated that most of these proteins function in reactive oxygen species (ROS) metabolism and defense response. Of the ROS-related proteins identified, we further characterized a thioredoxin, GbNRX1, which increased in abundance in response to V. dahliae challenge, finding that GbNRX1 functions in apoplastic ROS scavenging after the ROS burst that occurs upon recognition of V. dahliae. Silencing of GbNRX1 resulted in defective dissipation of apoplastic ROS, which led to higher ROS accumulation in protoplasts. As a result, the GbNRX1-silenced plants showed reduced wilt resistance, indicating that the initial defense response in the root apoplast requires the antioxidant activity of GbNRX1. Together, our results demonstrate that apoplastic ROS generation and scavenging occur in tandem in response to pathogen attack; also, the rapid balancing of redox to maintain homeostasis after the ROS burst, which involves GbNRX1, is critical for the apoplastic immune response.Cotton (Gossypium spp.) is one of the most economically important crops worldwide and a number of pathogens affect the growth and development of cotton plants. The soil-borne pathogen Verticillium dahliae (V. dahliae) causes the destructive vascular disease Verticillium wilt, which results in devastating reductions in plant mass, lint yield, and fiber quality (Bolek et al., 2005; Cai et al., 2009). To date, Verticillium wilt has not been effectively controlled in the most common cultivated cotton species, upland cotton (Gossypium hirsutum), and cultivars with stably inherited resistance to this disease are currently unavailable (Aguado et al., 2008; Jiang et al., 2009; Zhang et al., 2012a). Unlike upland cotton, sea-island cotton (Gossypium barbadense), which is only cultivated on a small scale, possesses Verticillium wilt resistance. Exploring the molecular mechanisms involved in the defense responses against V. dahliae invasion in G. barbadense can provide useful information for generating wilt-resistant G. hirsutum species through molecular breeding.During the past decades, progress has been made in studying the defense responses against V. dahliae infection in cotton. Global analyses have demonstrated that several signaling pathways, including those mediated by salicylic acid, ethylene, jasmonic acid, and brassinosteroids, activate distinct processes involved in V. dahliae defense (Bari and Jones, 2009; Grant and Jones, 2009; Gao et al., 2013a). Accumulating evidence indicates that many V. dahliae-responsive genes, such as GbWARKY1, GhSSN, GbERF, GhMLP28, GhNDR1, GhMKK2, and GhBAK1 (Qin et al., 2004; Gao et al., 2011, 2013b; Li et al., 2014a; Sun et al., 2014; Yang et al., 2015), play crucial roles in defense against Verticillium wilt. In addition, the biosynthesis of terpenoids, lignin, and gossypol also makes important contributions to V. dahliae resistance in cotton (Tan et al., 2000; Luo et al., 2001; Xu et al., 2011; Gao et al., 2013a). Together, these studies have greatly improved our understanding of the complex innate defense systems against V. dahliae infection in cotton.The initial interaction between plants and pathogens takes place in the apoplast, the compartment of the plant cell outside the cell membrane, including the cell wall and intercellular space (Dietz, 1997). In response to pathogen colonization, the attacked plant cells undergo significant cellular and molecular changes, such as reinforcement of the cell wall and secretion of antimicrobial molecules into the apoplastic space (Bednarek et al., 2010). Thus, the apoplast serves as the first line of defense against microbe invasion, and apoplast immunity can be considered an important component of the plant immune response to pathogens.Upon recognition of pathogen infection, rapid production of reactive oxygen species [the reactive oxygen species (ROS) burst] occurs in the apoplast (Lamb and Dixon, 1997; Torres et al., 2006; Torres, 2010). This ROS burst is regarded as a core component of the early plant immune response (Daudi et al., 2012; Doehlemann and Hemetsberger, 2013). During defense responses, apoplastic ROS can diffuse into the cytoplasm and serve as signals, interacting with other signaling processes such as phosphorylation cascades, calcium signaling, and hormone-mediated pathways (Kovtun et al., 2000; Mou et al., 2003). Apoplastic ROS can also directly strengthen the host cell walls by oxidative cross linking of glycoproteins (Bradley et al., 1992; Lamb and Dixon, 1997) or the precursors of lignin and suberin polymers (Hückelhoven, 2007). Moreover, apoplastic ROS can directly affect pathogens by degrading nucleic acids and peptides from microbes or causing lipid peroxidation and membrane damage in the microbe (Mehdy, 1994; Lamb and Dixon, 1997; Apel and Hirt, 2004; Montillet et al., 2005).ROS levels in the apoplast increase rapidly in response to a variety of pathogens, but subsequently return to basal levels. The rapid production and dissipation of apoplastic ROS indicate that this process is finely regulated. Two classes of enzymes, NADPH oxidases and class III peroxidases, account for the rapid ROS burst in the apoplast (Bolwell et al., 1995; O’Brien et al., 2012). NADPH oxidases are directly phosphorylated by the receptor-like kinase BIK1 to enhance ROS generation (Li et al., 2014b). Also, due to the toxicity of high levels of ROS, plants have evolved enzymatic and nonenzymatic mechanisms to eliminate ROS, thereby preventing or reducing oxidative damage (Rahal et al., 2014; Torres et al., 2006). However, the molecular system responsible for the regulation of apoplastic ROS homeostasis during the immune response is not well understood.In this study, we performed a comparative analysis of the apoplastic proteomes in control roots compared with V. dahliae-inoculated roots of Gossypium barbadense (wilt-resistant sea-island cotton) using the two-dimensional differential gel electrophoresis (2D-DIGE) technique. Among the differentially expressed apoplastic proteins, ROS-related proteins were found to be major components, including a thioredoxin, GbNRX1, which functions as an ROS scavenger in response to V. dahliae infection. Knock-down of GbNRX1 expression in cotton by virus-induced gene silencing (VIGS) resulted in reduced resistance to V. dahliae. Our results demonstrate that maintaining apoplastic ROS homeostasis is a crucial component of the apoplastic immune response and that GbNRX1 is an important regulator of this process.  相似文献   

6.
The default growth pattern of primary roots of land plants is directed by gravity. However, roots possess the ability to sense and respond directionally to other chemical and physical stimuli, separately and in combination. Therefore, these root tropic responses must be antagonistic to gravitropism. The role of reactive oxygen species (ROS) in gravitropism of maize and Arabidopsis (Arabidopsis thaliana) roots has been previously described. However, which cellular signals underlie the integration of the different environmental stimuli, which lead to an appropriate root tropic response, is currently unknown. In gravity-responding roots, we observed, by applying the ROS-sensitive fluorescent dye dihydrorhodamine-123 and confocal microscopy, a transient asymmetric ROS distribution, higher at the concave side of the root. The asymmetry, detected at the distal elongation zone, was built in the first 2 h of the gravitropic response and dissipated after another 2 h. In contrast, hydrotropically responding roots show no transient asymmetric distribution of ROS. Decreasing ROS levels by applying the antioxidant ascorbate, or the ROS-generation inhibitor diphenylene iodonium attenuated gravitropism while enhancing hydrotropism. Arabidopsis mutants deficient in Ascorbate Peroxidase 1 showed attenuated hydrotropic root bending. Mutants of the root-expressed NADPH oxidase RBOH C, but not rbohD, showed enhanced hydrotropism and less ROS in their roots apices (tested in tissue extracts with Amplex Red). Finally, hydrostimulation prior to gravistimulation attenuated the gravistimulated asymmetric ROS and auxin signals that are required for gravity-directed curvature. We suggest that ROS, presumably H2O2, function in tuning root tropic responses by promoting gravitropism and negatively regulating hydrotropism.Plants evolved the ability to sense and respond to various environmental stimuli in an integrated fashion. Due to their sessile nature, they respond to directional stimuli such as light, gravity, touch, and moisture by directional organ growth (curvature), a phenomenon termed tropism. Experiments on coleoptiles conducted by Darwin in the 1880s revealed that in phototropism, the light stimulus is perceived by the tip, from which a signal is transmitted to the growing part (Darwin and Darwin, 1880). Darwin postulated that in a similar manner, the root tip perceives stimuli from the environment, including gravity and moisture, processes them, and directs the growth movement, acting like “the brain of one of the lower animals” (Darwin and Darwin, 1880). The transmitted signal in phototropism and gravitropism was later found to be a phytohormone, and its redistribution on opposite sides of the root or shoot was hypothesized to promote differential growth and bending of the organ (Went, 1926; Cholodny, 1927). Over the years, the phytohormone was characterized as indole-3-acetic acid (IAA, auxin; Kögl et al., 1934; Thimann, 1935), and the ‘Cholodny-Went’ theory was demonstrated for gravitropism and phototropism (Rashotte et al., 2000; Friml et al., 2002). In addition to auxin, second messengers such as Ca2+, pH oscillations, reactive oxygen species (ROS) and abscisic acid (ABA) were shown to play an essential role in gravitropism (Young and Evans, 1994; Fasano et al., 2001; Joo et al., 2001; Ponce et al., 2008). Auxin was shown to induce ROS accumulation during root gravitropism, where the gravitropic bending is ROS dependent (Joo et al., 2001; Peer et al., 2013).ROS such as superoxide and hydrogen peroxide were initially considered toxic byproducts of aerobic respiration but currently are known also for their essential role in myriad cellular and physiological processes in animals and plants (Mittler et al., 2011). ROS and antioxidants are essential components of plant cell growth (Foreman et al., 2003), cell cycle control, and shoot apical meristem maintenance (Schippers et al., 2016) and play a crucial role in protein modification and cellular redox homeostasis (Foyer and Noctor, 2005). ROS function as signal molecules by mediating both biotic- (Sagi and Fluhr, 2006; Miller et al., 2009) and abiotic- (Kwak et al., 2003; Sharma and Dietz, 2009) stress responses. Joo et al. (2001) reported a transient increase in intracellular ROS concentrations early in the gravitropic response, at the concave side of maize roots, where auxin concentrations are higher. Indeed, this asymmetric ROS distribution is required for gravitropic bending, since maize roots treated with antioxidants, which act as ROS scavengers, showed reduced gravitropic root bending (Joo et al., 2001). The link between auxin and ROS production was later shown to involve the activation of NADPH oxidase, a major membrane-bound ROS generator, via a PI3K-dependent pathway (Brightman et al., 1988; Joo et al., 2005; Peer et al., 2013). Peer et al. (2013) suggested that in gravitropism, ROS buffer auxin signaling by oxidizing the active auxin IAA to the nonactive and nontransported form, oxIAA.Gravitropic-oriented growth is the default growth program of the plant, with shoots growing upwards and roots downward. However, upon exposure to specific external stimuli, the plant overcomes its gravitropic growth program and bends toward or away from the source of the stimulus. For example, as roots respond to physical obstacles or water deficiency. The ability of roots to direct their growth toward environments of higher water potential was described by Darwin and even earlier and was later defined as hydrotropism (Von Sachs, 1887; Jaffe et al., 1985; Eapen et al., 2005).In Arabidopsis (Arabidopsis thaliana), wild-type seedlings respond to moisture gradients (hydrostimulation) by bending their primary roots toward higher water potential. Upon hydrostimulation, amyloplasts, the starch-containing plastids in root-cap columella cells, which function as part of the gravity sensing system, are degraded within hours and recover upon water replenishment (Takahashi et al., 2003; Ponce et al., 2008; Nakayama et al., 2012). Moreover, mutants with a reduced response to gravity (pgm1) and to auxin (axr1 and axr2) exhibit higher responsiveness to hydrostimulation, manifested as accelerated bending compared to wild-type roots (Takahashi et al., 2002, 2003). Recently, we have shown that hydrotropic root bending does not require auxin redistribution and is accelerated in the presence of auxin polar transport inhibitors and auxin-signaling antagonists (Shkolnik et al., 2016). These results reflect the competition, or interference, between root gravitropism and hydrotropism (Takahashi et al., 2009). However, which cellular signals participate in the integration of the different environmental stimuli that direct root tropic curvature is still poorly understood. Here we sought to assess the potential role of ROS in regulating hydrotropism and gravitropism in Arabidopsis roots.  相似文献   

7.
The plastidic 2-C-methyl-d-erythritol-4-phosphate (MEP) pathway is one of the most important pathways in plants and produces a large variety of essential isoprenoids. Its regulation, however, is still not well understood. Using the stable isotope 13C-labeling technique, we analyzed the carbon fluxes through the MEP pathway and into the major plastidic isoprenoid products in isoprene-emitting and transgenic isoprene-nonemitting (NE) gray poplar (Populus × canescens). We assessed the dependence on temperature, light intensity, and atmospheric [CO2]. Isoprene biosynthesis was by far (99%) the main carbon sink of MEP pathway intermediates in mature gray poplar leaves, and its production required severalfold higher carbon fluxes compared with NE leaves with almost zero isoprene emission. To compensate for the much lower demand for carbon, NE leaves drastically reduced the overall carbon flux within the MEP pathway. Feedback inhibition of 1-deoxy-d-xylulose-5-phosphate synthase activity by accumulated plastidic dimethylallyl diphosphate almost completely explained this reduction in carbon flux. Our data demonstrate that short-term biochemical feedback regulation of 1-deoxy-d-xylulose-5-phosphate synthase activity by plastidic dimethylallyl diphosphate is an important regulatory mechanism of the MEP pathway. Despite being relieved from the large carbon demand of isoprene biosynthesis, NE plants redirected only approximately 0.5% of this saved carbon toward essential nonvolatile isoprenoids, i.e. β-carotene and lutein, most probably to compensate for the absence of isoprene and its antioxidant properties.Isoprenoids represent the largest and most diverse group (over 50,000) of natural compounds and are essential in all living organisms (Gershenzon and Dudareva, 2007; Thulasiram et al., 2007). They are economically important for humans as flavor and fragrance, cosmetics, drugs, polymers for rubber, and precursors for the chemical industry (Chang and Keasling, 2006). The broad variety of isoprenoid products is formed from two building blocks, dimethylallyl diphosphate (DMADP) and isopentenyl diphosphate (IDP). In plants, the plastidic 2-C-methyl-d-erythritol-4-phosphate (MEP) pathway (Zeidler et al., 1997) produces physiologically and ecologically important volatile organic compounds (VOCs), the carotenoids (tetraterpenes; Giuliano et al., 2008; Cazzonelli and Pogson, 2010), diterpenes, the prenyl side-chains of chlorophylls (Chls) and plastoquinones, isoprenylated proteins, the phytohormones gibberellins, and side-chain of cytokinins (for review, see Dudareva et al., 2013; Moses et al., 2013). Industrially important prokaryotes (e.g. Escherichia coli) also use the MEP pathway for the biosynthesis of isoprenoids (Vranová et al., 2012), and there is an increasing interest in manipulating the MEP pathway of engineered microbes to increase production of economically relevant isoprenoids (Chang and Keasling, 2006). To achieve this, a mechanistic understanding of the regulation of the MEP pathway is needed (Vranová et al., 2012).Some plants, including poplars (Populus spp.), produce large amounts of the hemiterpene VOC isoprene. Worldwide isoprene emissions from plants are estimated to be 600 teragrams per year and to account for one-third of all hydrocarbons emitted to the atmosphere (Arneth et al., 2008; Guenther, 2013). Isoprene has strong effects on air chemistry and climate by participating in ozone formation reactions (Fuentes et al., 2000), by prolonging the lifespan of methane, a greenhouse gas (Poisson et al., 2000; Archibald et al., 2011), and by taking part in the formation of secondary organic aerosols (Kiendler-Scharr et al., 2012).Poplar leaves invest a significant amount of recently fixed carbon in isoprene biosynthesis (Delwiche and Sharkey, 1993; Schnitzler et al., 2010; Ghirardo et al., 2011) to cope with abiotic stresses (Sharkey, 1995; Velikova and Loreto, 2005; Behnke et al., 2007, 2010b, 2013; Vickers et al., 2009; Loreto and Schnitzler, 2010; Sun et al., 2013b), although there are indications that other protective mechanisms can partially compensate the lack of isoprene emission in genetically transformed poplars (Behnke et al., 2012; Way et al., 2013). It has been suggested that in isoprene-emitting (IE) species, most of the carbon that passes through the MEP pathway is used for isoprene biosynthesis (Sharkey and Yeh, 2001). However, a recent study using pulse-chase labeling with 14C has shown continuous synthesis and degradation of carotenes and Chl a in mature leaves of Arabidopsis (Arabidopsis thaliana; Beisel et al., 2010), and the amount of flux diverted to carotenoid and Chl synthesis compared with isoprene biosynthesis in poplar leaves is not known.Isoprene emission is temperature, light, and CO2 dependent (Schnitzler et al., 2005; Rasulov et al., 2010; Way et al., 2011; Monson et al., 2012; Li and Sharkey, 2013a). It has been demonstrated that isoprene biosynthesis depends on the activities of IDP isomerase (EC 5.3.3.2), isoprene synthase (ISPS; EC 4.2.3.27), and the amount of ISPS substrate, DMADP (Brüggemann and Schnitzler, 2002a, 2002b; Schnitzler et al., 2005; Rasulov et al., 2009b). In turn, DMADP concentration has been hypothesized to act as a feedback regulator of the MEP pathway by inhibiting 1-deoxy-d-xylulose-5-phosphate synthase (DXS; EC 2.2.1.7), the first enzyme of the MEP pathway (Banerjee et al., 2013). Understanding the controlling mechanism of isoprene biosynthesis is not only of fundamental relevance, but also necessary for engineering the MEP pathway in various organisms and for accurate simulation of isoprene emissions by plants in predicting atmospheric reactivity (Niinemets and Monson, 2013).There is ample evidence that silencing the ISPS in poplar has a broad effect on the leaf metabolome (Behnke et al., 2009, 2010a, 2013; Way et al., 2011; Kaling et al., 2014). While some of those changes (e.g. ascorbate and α-tocopherol) are compensatory mechanisms to cope with abiotic stresses, others (e.g. shikimate pathway and phenolic compounds) might be related to the alteration of the MEP pathway (Way et al., 2013; Kaling et al., 2014). The perturbation of these metabolic pathways can be attributed to the removal of a major carbon sink of the MEP pathway and the resulting change in the energy balance within the plant cell (Niinemets et al., 1999; Ghirardo et al., 2011). In this work, we analyzed the carbon fluxes through the MEP pathway into the main plastidic isoprenoid products.We used the 13C-labeling technique as a tool to measure the carbon fluxes through the MEP pathway at different temperatures, light intensities, and CO2 concentrations in mature leaves of IE and transgenic, isoprene-nonemitting (NE) gray poplar (Populus × canescens). Isoprene emission was drastically reduced in the transgenic trees through knockdown of PcISPS gene expression by RNA interference, resulting in plants with only 1% to 5% of isoprene emission potential compared with wild-type plants (Behnke et al., 2007).We measured the appearance of 13C in the isoprenoid precursors 2-C-methyl-d-erythritol-2,4-cyclodiphosphate (MEcDP) and DMADP as well as isoprene and the major downstream products of the MEP pathway, i.e. carotenoids and Chls. To reliably detect de novo synthesis of the pigments, which occur at very low rates (Beisel et al., 2010), we used isotope ratio mass spectrometry (IRMS).Here, (1) we quantify the effect of isoprene biosynthesis on the MEP pathway in poplar, and (2) we show that suppression of isoprene biosynthesis negatively affects the carbon flux through the MEP pathway by accumulating plastidic DMADP, which feeds back to inhibit PcDXS, leading to (3) a slight increase of carbon flux toward production of greater chain-length isoprenoids and (4) a strong decrease in the overall isoprenoid carbon fluxes to compensate for the much lower MEP pathway demand for carbon. This study strongly supports the hypothesis that an important regulatory mechanism of the MEP pathway is the feedback regulation of plastidic DMADP on DXS. The large carbon flux through the MEP pathway of IE poplar plastids demonstrates the potential of transgenically altered IE plant species to produce economically valuable isoprenoids at high rates in, for instance, industrial applications.  相似文献   

8.
9.
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   

10.
11.
Plant resistance to phytopathogenic microorganisms mainly relies on the activation of an innate immune response usually launched after recognition by the plant cells of microbe-associated molecular patterns. The plant hormones, salicylic acid (SA), jasmonic acid, and ethylene have emerged as key players in the signaling networks involved in plant immunity. Rhamnolipids (RLs) are glycolipids produced by bacteria and are involved in surface motility and biofilm development. Here we report that RLs trigger an immune response in Arabidopsis (Arabidopsis thaliana) characterized by signaling molecules accumulation and defense gene activation. This immune response participates to resistance against the hemibiotrophic bacterium Pseudomonas syringae pv tomato, the biotrophic oomycete Hyaloperonospora arabidopsidis, and the necrotrophic fungus Botrytis cinerea. We show that RL-mediated resistance involves different signaling pathways that depend on the type of pathogen. Ethylene is involved in RL-induced resistance to H. arabidopsidis and to P. syringae pv tomato whereas jasmonic acid is essential for the resistance to B. cinerea. SA participates to the restriction of all pathogens. We also show evidence that SA-dependent plant defenses are potentiated by RLs following challenge by B. cinerea or P. syringae pv tomato. These results highlight a central role for SA in RL-mediated resistance. In addition to the activation of plant defense responses, antimicrobial properties of RLs are thought to participate in the protection against the fungus and the oomycete. Our data highlight the intricate mechanisms involved in plant protection triggered by a new type of molecule that can be perceived by plant cells and that can also act directly onto pathogens.In their environment, plants are challenged by potentially pathogenic microorganisms. In response, they express a set of defense mechanisms including preformed structural and chemical barriers, as well as an innate immune response quickly activated after microorganism perception (Boller and Felix, 2009). Plant innate immunity is triggered after recognition by pattern recognition receptors of conserved pathogen- or microbe-associated molecular patterns (PAMPs or MAMPs, respectively) or by plant endogenous molecules released by pathogen invasion and called danger-associated molecular patterns (Boller and Felix, 2009; Dodds and Rathjen, 2010). This first step of recognition leads to the activation of MAMP-triggered immunity (MTI). Successful pathogens can secrete effectors that interfere or suppress MTI, resulting in effector-triggered susceptibility. A second level of perception involves the direct or indirect recognition by specific receptors of pathogen effectors leading to effector-triggered immunity (ETI; Boller and Felix, 2009; Dodds and Rathjen, 2010). Whereas MTI and ETI are thought to involve common signaling network, ETI is usually quantitatively stronger than MTI and associated with more sustained and robust immune responses (Katagiri and Tsuda, 2010; Tsuda and Katagiri, 2010).The plant hormones, salicylic acid (SA), jasmonic acid (JA), and ethylene (ET) have emerged as key players in the signaling networks involved in MTI and ETI (Robert-Seilaniantz et al., 2007; Tsuda et al., 2009; Katagiri and Tsuda, 2010; Mersmann et al., 2010; Tsuda and Katagiri, 2010; Robert-Seilaniantz et al., 2011). Interactions between these signal molecules allow the plant to activate and/or modulate an appropriate spectrum of responses, depending on the pathogen lifestyle, necrotroph or biotroph (Glazebrook, 2005; Koornneef and Pieterse, 2008). It is assumed that JA and ET signaling pathways are important for resistance to necrotrophic fungi including Botrytis cinerea and Alternaria brassicicola (Thomma et al., 2001; Ferrari et al., 2003; Glazebrook, 2005). Infection of Arabidopsis (Arabidopsis thaliana) with B. cinerea causes the induction of the JA/ET responsive gene PLANT DEFENSIN1.2 (PDF1.2; Penninckx et al., 1996; Zimmerli et al., 2001). Induction of PDF1.2 by B. cinerea is blocked in ethylene-insensitive2 (ein2) and coronatine-insensitive1 (coi1) mutants that are respectively defective in ET and JA signal transduction pathways. Moreover, ein2 and coi1 plants are highly susceptible to B. cinerea infection (Thomma et al., 1998; Thomma et al., 1999). JA/ET-dependent responses do not seem to be usually induced during resistance to biotrophs, but they can be effective if they are stimulated prior to pathogen challenge (Glazebrook, 2005). Plants impaired in SA signaling are highly susceptible to biotrophic and hemibiotrophic pathogens. Following pathogen infection, SA hydroxylase (NahG), enhanced disease susceptibility5 (eds5), or SA induction-deficient2 (sid2) plants are unable to accumulate high SA levels and they display heightened susceptibility to Pseudomonas syringae pv tomato (Pst), Hyaloperonospora arabidopsidis, or Erysiphe orontii (Delaney et al., 1994; Lawton et al., 1995; Wildermuth et al., 2001; Nawrath et al., 2002; Vlot et al., 2009). Mutants that are insensitive to SA, such as nonexpressor of PATHOGENESIS-RELATED (PR) genes1 (npr1), have enhanced susceptibility to these pathogens (Cao et al., 1994; Glazebrook et al., 1996; Shah et al., 1997; Dong, 2004). According to some reports, plant defense against necrotrophs also involves SA. Arabidopsis plants expressing the nahG gene and infected with B. cinerea show larger lesions compared with wild-type plants (Govrin and Levine, 2002). In tobacco (Nicotiana tabacum), acidic isoforms of PR3 and PR5 gene that are specifically induced by SA (Ménard et al., 2004) are up-regulated after challenge by B. cinerea (El Oirdi et al., 2010). Resistance to some necrotrophs like Fusarium graminearum involves both SA and JA signaling pathways (Makandar et al., 2010). It is assumed that SA and JA signaling can be antagonistic (Bostock, 2005; Koornneef and Pieterse, 2008; Pieterse et al., 2009; Thaler et al., 2012). In Arabidopsis, SA inhibits JA-dependent resistance against A. brassicicola or B. cinerea (Spoel et al., 2007; Koornneef et al., 2008). Recent studies demonstrated that ET modulates the NPR1-mediated antagonism between SA and JA (Leon-Reyes et al., 2009; Leon-Reyes et al., 2010a) and suppression by SA of JA-responsive gene expression is targeted at a position downstream of the JA biosynthesis pathway (Leon-Reyes et al., 2010b). Synergistic effects of SA- and JA-dependent signaling are also well documented (Schenk et al., 2000; van Wees et al., 2000; Mur et al., 2006) and induction of some defense responses after pathogen challenge requires intact JA, ET, and SA signaling pathways (Campbell et al., 2003).Isolated MAMPs trigger defense responses that also require the activation of SA, JA, and ET signaling pathways (Tsuda et al., 2009; Katagiri and Tsuda, 2010). For instance, treatment with the flagellin peptide flg22 induces many SA-related genes including SID2, EDS5, NPR1, and PR1 (Ferrari et al., 2007; Denoux et al., 2008), causes SA accumulation (Tsuda et al., 2008; Wang et al., 2009), and activates ET signaling (Bethke et al., 2009; Mersmann et al., 2010). Local application of lipopolysaccharides elevates the level of SA (Mishina and Zeier, 2007). The oomycete Pep13 peptide induces defense responses in potato (Solanum tuberosum) that require both SA and JA (Halim et al., 2009). Although signaling networks induced by isolated MAMPs are well documented, the contribution of SA, JA, and ET in MAMP- or PAMP-induced resistance to biotrophs and necrotrophs is poorly understood.Rhamnolipids (RLs) are glycolipids produced by various bacteria species including some Pseudomonas and Burkholderia species. They are essential for bacterial surface motility and biofilm development (Vatsa et al., 2010; Chrzanowski et al., 2012). RLs are potent stimulators of animal immunity (Vatsa et al., 2010). They have recently been shown to elicit plant defense responses and to induce resistance against B. cinerea in grapevine (Vitis vinifera; Varnier et al., 2009). They also participate to biocontrol activity of the plant beneficial bacteria Pseudomonas aeruginosa PNA1 against oomycetes (Perneel et al., 2008). However, the signaling pathways used by RLs to stimulate plant innate immunity are not known. To gain more insights into RL-induced MTI, we investigated RL-triggered defense responses and resistance to the necrotrophic fungus B. cinerea, the biotroph oomycete H. arabidopsidis, and the hemibiotroph bacterium Pst in Arabidopsis. Our results show that RLs trigger an innate immune response in Arabidopsis that protects the plant against these different lifestyle pathogens. We demonstrate that RL-mediated resistance involves separated signaling sectors that depend on the type of pathogen. In plants challenged by RLs, SA has a central role and participates to the restriction of the three pathogens. ET is fully involved in RL-induced resistance to the biotrophic oomycete and to the hemibiotrophic bacterium whereas JA is essential for the resistance to the necrotrophic fungus.  相似文献   

12.
13.
14.
15.
16.
17.
18.
A method is presented for rapid extraction of the total plastoquinone (PQ) pool from Synechocystis sp. strain PCC 6803 cells that preserves the in vivo plastoquinol (PQH2) to -PQ ratio. Cells were rapidly transferred into ice-cold organic solvent for instantaneous extraction of the cellular PQ plus PQH2 content. After high-performance liquid chromatography fractionation of the organic phase extract, the PQH2 content was quantitatively determined via its fluorescence emission at 330 nm. The in-cell PQH2-PQ ratio then followed from comparison of the PQH2 signal in samples as collected and in an identical sample after complete reduction with sodium borohydride. Prior to PQH2 extraction, cells from steady-state chemostat cultures were exposed to a wide range of physiological conditions, including high/low availability of inorganic carbon, and various actinic illumination conditions. Well-characterized electron-transfer inhibitors were used to generate a reduced or an oxidized PQ pool for reference. The in vivo redox state of the PQ pool was correlated with the results of pulse-amplitude modulation-based chlorophyll a fluorescence emission measurements, oxygen exchange rates, and 77 K fluorescence emission spectra. Our results show that the redox state of the PQ pool of Synechocystis sp. strain PCC 6803 is subject to strict homeostatic control (i.e. regulated between narrow limits), in contrast to the more dynamic chlorophyll a fluorescence signal.The photosynthetic apparatus of oxygenic phototrophs consists of two types of photosynthetic reaction centers: PSII and PSI. Both photosystems are connected in series, with electrons flowing from PSII toward PSI through an intermediate electron transfer chain, which comprises the so-called plastoquinone (PQ) pool, plastocyanin and/or cytochrome c553, and the cytochrome b6f complex. The redox potential of the PQ pool is clamped by the relative rates of electron release into and uptake from this pool. Within the PSII complex, electrons are extracted from water at the lumenal side of the thylakoid membrane and transferred to the primary accepting quinone (QA) at the stromal side. The electron is subsequently transferred to a PQ molecule in the secondary accepting quinone (QB) of PSII. The intermediate QB semiquinone, which is formed accordingly, is stable in the QB site for several seconds (Diner et al., 1991; Mitchell, 1993) and subsequently can be reduced to plastoquinol (PQH2). The midpoint potential of QA reduction is approximately −100 mV (Krieger-Liszkay and Rutherford, 1998; Allakhverdiev et al., 2011), whereas the corresponding midpoint potential of the QB semiquinone is close to zero (Nicholls and Ferguson, 2013). PQH2 equilibrates with the PQ pool in the thylakoid membranes, which has a size that is approximately 1 order of magnitude larger than the number of PSII reaction centers (Melis and Brown, 1980; Aoki and Katoh, 1983).PQ is a lipophilic, membrane-bound electron carrier, with a midpoint potential of +80 mV (Okayama, 1976), that can accept two electrons and two protons to form PQH2 (Rich and Bendall, 1980). PQH2 can donate both electrons to the cytochrome b6f complex, one to low-potential cytochrome b6, by which reduced high-potential cytochrome b6 is formed, and one to the cytochrome f moiety on the lumenal side of the thylakoid membrane, where the two protons are released. High-potential cytochrome b6 then donates an electron back to PQ on the stromal side of the membrane, rendering a semiquinone in the PQ-binding pocket on the cytoplasmic face of the b6f complex ready as an acceptor of another electron from PSII, and reduced cytochrome f feeds an electron to a water-soluble electron carrier (i.e. either plastocyanin or cytochrome c553) for subsequent transfer to the reaction center of PSI or to cytochrome c oxidase, respectively (Rich et al., 1991; Geerts et al., 1994; Schubert et al., 1995; Paumann et al., 2004; Mulkidjanian, 2010).Electron transfer through the cytochrome b6f complex proceeds according to the Q-cycle mechanism (Rich et al., 1991). As a result, maximally two protons from the stroma are released into the lumen per electron transferred. This electrochemical proton gradient can be used for the synthesis of ATP by the ATP synthase complex (Walker, 1998). In PSI, another transthylakoid membrane charge separation process is energized by light. Electron transfer within the PSI complex involves iron-sulfur clusters and quinones and leads to the reduction of ferredoxin, the reduced form of which serves as the electron donor for NADPH by the ferredoxin:NADP+ oxidoreductase enzyme (van Thor et al., 1999). The ATP and NADPH generated this way are used for CO2 fixation in a mutual stoichiometry that is close to the stoichiometry at which these two energy-rich compounds are formed at the thylakoid membrane. Normally, this ratio is ATP:NADPH = 3:2 (Behrenfeld et al., 2008).Photosynthetic and respiratory electron transport in cyanobacteria share a single PQ pool (Aoki and Katoh, 1983; Aoki et al., 1983; Matthijs et al., 1984; Scherer, 1990). Respiratory electron transfer provides cells the ability to form ATP in the dark, but this ability is not limited to those conditions. Transfer of electrons into the PQ pool is the result of the joint activity of PSII, respiratory dehydrogenases [in particular those specific for NAD(P)H and succinate], and cyclic electron transport around PSI (Mi et al., 1995; Cooley et al., 2000; Howitt et al., 2001;Yeremenko et al., 2005), whereas oxidation of PQH2 is catalyzed by the PQH2 oxidase, the cytochrome b6f complex, the respiratory cytochrome c oxidase (Nicholls et al., 1992; Pils and Schmetterer, 2001; Berry et al., 2002), and possibly plasma terminal oxidase (Peltier et al., 2010). Multiples of these partial reactions can proceed simultaneously, including respiratory electron transfer during illumination (Schubert et al., 1995), which includes oxygen uptake through a Mehler-like reaction (Helman et al., 2005; Allahverdiyeva et al., 2013).Because of its central location between the two photosystems, the redox state of the PQ pool has been identified as an important parameter that can signal photosynthetic imbalances (Mullineaux and Allen, 1990; Allen, 1995; Ma et al., 2010; Allen et al., 2011). Yet, an accurate estimation of the in vivo redox state of this pool has not been reported in cyanobacteria so far. Instead, the redox state of the PQ pool is widely assumed to be reflected in, or related to, the intensity of the chlorophyll a fluorescence emissions (Prasil et al., 1996; Yang et al., 2001; Gotoh et al., 2010; Houyoux et al., 2011). Imbalance in electron transport through the two photosystems may lead to a loss of excitation energy and, hence, to a loss of chlorophyll a fluorescence emission (Schreiber et al., 1986). Therefore, patterns of chlorophyll a fluorescence (pulse-amplitude modulated [PAM] fluorimetry; Baker, 2008) have widely been adopted for the analysis of (un)balanced photosynthetic electron transfer and, by inference, for indirect recording of the redox state of the PQ pool. However, the multitude of electron transfer pathways in the thylakoid membranes of cyanobacteria (see above) makes it much more complex to explain PAM signals in these organisms than in chloroplasts (Campbell et al., 1998). Additional regulatory mechanisms of nonphotochemical quenching, via the xanthophyll cycle in chloroplasts (Demmig-Adams et al., 2012) and the orange carotenoid protein (Kirilovsky and Kerfeld, 2012) in cyanobacteria, and energy redistribution via state transitions (Allen, 1995; Van Thor et al., 1998) complicate such comparisons even further.Several years ago, an HPLC-based technique was developed for the detection of the redox state of PQH2 in isolated thylakoids (Kruk and Karpinski, 2006), but these results have neither been related to physiological conditions nor to the results of chlorophyll a fluorescence measurements. In this report, we describe an adaptation of this method with elements of a method for estimation of the redox state of the ubiquinone pool in Escherichia coli (Bekker et al., 2007). This modified method allows for reliable measurements of the redox state of the PQ pool of Synechocystis sp. strain PCC 6803 under physiologically relevant conditions. The method uses rapid cell lysis in an organic solvent to arrest all physiological processes, followed by extraction and identification of PQH2 by HPLC separation with fluorescence detection. Next, we manipulated the redox state of the PQ pool with various redox-active agents, with inhibitors of photosynthetic electron flow, and by illumination with light specific for either PSII or PSI. The measured redox state of the PQ pool was then related to the chlorophyll a fluorescence signal and 77 K fluorescence emission spectra of cell samples taken in parallel and to oxygen-exchange rates measured separately. These experiments reveal that, despite highly fluctuating conditions of photosynthetic and respiratory electron flow, a remarkably stable redox state of the PQ pool is maintained. This homeostatically regulated redox state correlates poorly in many of the conditions tested with the more dynamic signal of chlorophyll a fluorescence emission, as measured with PAM fluorimetry. The latter signal only reflects the redox state of QA and not that of the PQ pool.  相似文献   

19.
20.
We have investigated the importance of carotenoids on the accumulation and function of the photosynthetic apparatus using a mutant of the green alga Chlamydomonas reinhardtii lacking carotenoids. The FN68 mutant is deficient in phytoene synthase, the first enzyme of the carotenoid biosynthesis pathway, and therefore is unable to synthesize any carotenes and xanthophylls. We find that FN68 is unable to accumulate the light-harvesting complexes associated with both photosystems as well as the RC subunits of photosystem II. The accumulation of the cytochrome b6f complex is also strongly reduced to a level approximately 10% that of the wild type. However, the residual fraction of assembled cytochrome b6f complexes exhibits single-turnover electron transfer kinetics comparable to those observed in the wild-type strain. Surprisingly, photosystem I is assembled to significant levels in the absence of carotenoids in FN68 and possesses functional properties that are very similar to those of the wild-type complex.Carotenoids (Cars) are fundamental components of the photosynthetic apparatus (Young and Britton, 1993, and refs. therein). The vast majority of Cars are noncovalently bound to either the core or the antenna subunits of PSI or PSII (Siefermann-Harms, 1985; Bassi et al., 1993). The most abundant Car bound to the core subunits of both photosystems is β-carotene, which is found in the vast majority of oxygenic organisms (Siefermann-Harms, 1985; Bassi et al., 1993). The light-harvesting complexes (LHCs) that act as the outer antenna in plants and green algae bind a wider range of oxygenated Cars, known as xanthophylls, the most abundant of which is lutein (Siefermann-Harms, 1985; Bassi et al., 1993; Jennings et al., 1996). The stoichiometry of xanthophylls binding to LHC complexes depends on the particular complexes and often on the illumination conditions during the organism’s growth (Siefermann-Harms, 1985; Demmig-Adams, 1990; Horton et al., 1996). Intriguingly, a molecule of β-carotene (as well as a molecule of chlorophyll [Chl] a) is found also in the cytochrome (Cyt) b6f complex (Kurisu et al., 2003; Stroebel et al., 2003).Cars have multiple functions in the photosynthetic process; they act as light-harvesting pigments (Frank and Cogdell, 1993), enlarging the optical cross section to radiation that is poorly absorbed by Chl. Moreover, Cars play a crucial role in processes such as nonphotochemical quenching that control the efficiency of light harvesting in response to the intensity of the incident radiation (for review, see Demmig-Adams, 1990; Horton et al., 1996; Niyogi, 1999). Probably the most important role of Cars in photosynthesis is the quenching of the excited triplet state of Chl (for review, see Frank and Cogdell, 1993; Giacometti et al., 2007), preventing the formation of highly reactive singlet oxygen, which represents the principal species active under high light stress (Hideg et al., 1994; Krieger-Liszkay, 2005). The importance of Cars is demonstrated by the observation that disruption of their biosynthesis through mutation, or by inhibition of a key enzyme in the pathway, leads to either lethal phenotypes or to rapid photobleaching of the photosynthetic tissue (Claes, 1957; Faludi-Dániel et al., 1968, 1970; Bolychevtseva et al., 1995; Trebst and Depka, 1997).Moreover, it has been shown that the presence of xanthophylls is absolutely necessary for refolding in vitro of LHC I and LHC II antenna complexes (Plumley and Schmidt, 1987; Paulsen et al., 1993; Sandonà et al., 1998). Such Cars, therefore, have a structural role, as well as their involvement in light harvesting, nonphotochemical quenching regulation, and the quenching of the Chl triplet state. Whether Cars also play a key structural role in the formation and stability of the core complexes of both PSI and PSII has not been systematically explored, since assembly of these complexes in vitro is not feasible. Studies in vivo using higher plants are complicated by the fact that Car deficiency is lethal and can be studied only during the early stages of greening and leaf development (Faludi-Dániel et al., 1968, 1970; Inwood et al., 2008). In these studies, it was shown that the accumulation of PSII complexes was greatly impaired in mutants of maize (Zea mays; Faludi-Dániel et al., 1968, 1970; Inwood et al., 2008), while the assembly of PSI appeared to be less sensitive to Car availability. In mutants of the cyanobacterium Synechocystis sp. PCC 6803 lacking the genes for phytoene desaturase or ζ-carotene desaturase, there was a complete loss of PSII assembly, while functional PSI complexes were assembled, albeit with slightly altered electron transfer kinetics with respect to the wild-type complex (Bautista et al., 2005). In agreement with the higher sensitivity of PSII assembly to Car availability, Trebst and Depka (1997) reported a specific effect on the synthesis of the D1 subunit of PSII RC upon treatment with phytoene desaturase inhibitors. On the other hand, it has recently been reported that in lycopene-β-cyclase mutants of Arabidopsis (Arabidopsis thaliana) that have a decreased amount of β-carotene (bound to the RC) with respect to most of the xanthophyll pool pigments (bound to the LHCs), the level of accumulation of PSI complexes, particularly that of the LHC I complement, was more affected that that of PSII, probably also because of an increased sensitivity to photodamage of mutated PSI RC (Cazzaniga et al., 2012; Fiore et al., 2012).In this investigation, we have studied the accumulation and functionality of the major chromophore-binding complexes of the photosynthetic apparatus, PSI, PSII, and Cyt b6f, in a Car-less mutant of the green alga Chlamydomonas reinhardtii (FN68) that is blocked at the first committed step of Car biosynthesis, namely, phytoene synthesis (McCarthy et al., 2004). Although the mutant is incapable of growing under phototrophic or photomixotrophic conditions, it can grow in complete darkness on a medium supplemented with a carbon source. Here, we show that the PSII core and antenna complexes fail to accumulate in the mutant and that the Cyt b6f complex accumulates to approximately one-tenth of the wild-type level. On the other hand, the PSI reaction center accumulates in FN68 and possesses electron transfer properties that are remarkably similar to those of wild-type PSI. Interestingly, we find that the level of PSI accumulation differs in other phytoene synthase null mutants, suggesting that additional mutations in one or other of these strains affect PSI stability. Nevertheless, our findings demonstrate that Cars are not required for either the assembly or the functionality of PSI in vivo.  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