首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
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   

3.
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   

4.
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   

5.
6.
7.
8.
Recently, a feedback inhibition of the chloroplastic 1-deoxy-d-xylulose 5-phosphate (DXP)/2-C-methyl-d-erythritol 4-phosphate (MEP) pathway of isoprenoid synthesis by end products dimethylallyl diphosphate (DMADP) and isopentenyl diphosphate (IDP) was postulated, but the extent to which DMADP and IDP can build up is not known. We used bisphosphonate inhibitors, alendronate and zoledronate, that inhibit the consumption of DMADP and IDP by prenyltransferases to gain insight into the extent of end product accumulation and possible feedback inhibition in isoprene-emitting hybrid aspen (Populus tremula × Populus tremuloides). A kinetic method based on dark release of isoprene emission at the expense of substrate pools accumulated in light was used to estimate the in vivo pool sizes of DMADP and upstream metabolites. Feeding with fosmidomycin, an inhibitor of DXP reductoisomerase, alone or in combination with bisphosphonates was used to inhibit carbon input into DXP/MEP pathway or both input and output. We observed a major increase in pathway intermediates, 3- to 4-fold, upstream of DMADP in bisphosphonate-inhibited leaves, but the DMADP pool was enhanced much less, 1.3- to 1.5-fold. In combined fosmidomycin/bisphosphonate treatment, pathway intermediates accumulated, reflecting cytosolic flux of intermediates that can be important under strong metabolic pull in physiological conditions. The data suggested that metabolites accumulated upstream of DMADP consist of phosphorylated intermediates and IDP. Slow conversion of the huge pools of intermediates to DMADP was limited by reductive energy supply. These data indicate that the DXP/MEP pathway is extremely elastic, and the presence of a significant pool of phosphorylated intermediates provides an important valve for fine tuning the pathway flux.Isoprenoids constitute a versatile class of compounds fulfilling major physiological functions. They are formed by two pathways in plants, the mevalonate (MVA) pathway in the cytosol (Gershenzon and Croteau, 1993) and the 1-deoxy-d-xylulose 5-phosphate (DXP)/2-C-methyl-d-erythritol 4-phosphate (MEP) pathway in plastids (Gershenzon and Croteau, 1993; Jomaa et al., 1999; Li and Sharkey, 2013b). The MVA pathway is primarily responsible for the synthesis of sesquiterpenes (C15), triterpenes (C30) including brassinosteroids, and even larger molecules such as dolichols (Bick and Lange, 2003; Li and Sharkey, 2013b; Rajabi Memari et al., 2013; Rosenkranz and Schnitzler, 2013). The DXP/MEP pathway is responsible for the synthesis of the simplest isoprenoids, isoprene and 2-methyl-3-buten-2-ol (C5), monoterpenes (C10), diterpenes (C20) including gibberellins and phytol residue of chlorophylls, and tetraterpenes (C40) including carotenoids (Rodríguez-Concepción and Boronat, 2002; Roberts, 2007).Given that in plants both pathways produce ultimately the same substrates, dimethylallyl diphosphate (DMADP) and isopentenyl diphosphate (IDP), the pertinent question is to what extent the two pathways can exchange metabolites (Rodríguez-Concepción and Boronat, 2002). There is evidence of a certain exchange of IDP between cytosolic and plastidic compartments, although the contribution of IDP from one compartment to the pathway flux in the other seems to be relatively minor (Schwender et al., 2001; Rodríguez-Concepción and Boronat, 2002; Bick and Lange, 2003). Some studies have further demonstrated that the exchange of IDP is fully bidirectional (De-Eknamkul and Potduang, 2003; Rodríguez-Concepción, 2006), whereas other studies suggest that IDP export from plastids to cytosol operates with a greater efficiency than the opposite transport (Hemmerlin et al., 2003; Laule et al., 2003). However, although the overall intercompartmental exchange of isoprenoid substrates to pathway flux in the given compartment might seem minor under nonstressed conditions, the importance of cross talk among the pathways might increase under stress conditions that specifically inhibit isoprenoid synthesis in one pathway. In fact, the DXP/MEP pathway is strongly linked to photosynthetic metabolism, and therefore, inhibition of photosynthesis under stressful conditions such as heat stress or drought or photoinhibition could inhibit the synthesis of isoprenoids when they are most needed to fulfill their protective function (Loreto and Schnitzler, 2010; Niinemets, 2010; Possell and Loreto, 2013). There is some evidence demonstrating a certain cooperativity among the two isoprenoid synthesis pathways under conditions leading to a reduction of the activity of one of them (Piel et al., 1998; Jux et al., 2001; Page et al., 2004; Rodríguez-Concepción, 2006), but the capacity for such a replacement of function and regulation is poorly understood.Recent studies using genetically modified plants accumulating end products of the DXP/MEP pathway or using natural variation in product accumulation have demonstrated the existence of a potentially important feedback regulation of the DXP/MEP pathway flux by primary end products of the pathway (Banerjee et al., 2013; Ghirardo et al., 2014; Wright et al., 2014). In particular, binding of DMADP and perhaps IDP to DXP synthase, the first enzyme in the DXP/MEP pathway, leads to downregulation of the pathway flux when the end products cannot be used, such as under stress conditions. However, the strength of such a feedback regulation can be importantly modified by accumulation of phosphorylated intermediates of the pathway, such that DMADP and IDP do not accumulate. Previous studies have demonstrated that there is a certain pool of phosphorylated intermediates in vivo, and that this pool can strongly increase under certain conditions, including experimental and genetic modification of DXP/MEP pathway input and output (Li et al., 2011; Rasulov et al., 2011; Li and Sharkey, 2013a; Ghirardo et al., 2014; Wright et al., 2014).It has further been shown that 2-C-methyl-d-erythritol 2,4-cyclodiphosphate (ME-cDP) is the metabolite accumulating in the plastids, and this accumulation can buffer DMADP and IDP changes in the case of varying DXP/MEP pathway input and consumption (Li and Sharkey, 2013a; Wright et al., 2014). A significant part of ME-cDP might even escape to cytosol, implying the existence of another interesting link between cytosolic and chloroplastic processes in isoprenoid synthesis (Wright et al., 2014). Furthermore, as ME-cDP is an important signaling molecule eliciting a number of gene expression responses (Xiao et al., 2012), accumulation of ME-cDP in plastids and flux to cytosol and further to the nucleus is particularly interesting from the perspective of long-term regulation of isoprenoid synthesis, and suggests a coordination of cellular stress responses by the plastidial isoprenoid synthesis pathway.Isoprene-emitting species constitute an exciting model system where a very large DXP/MEP pathway flux goes to isoprene synthesis under physiological conditions (Li and Sharkey, 2013b; Sharkey et al., 2013). In isoprene-emitting species, there is a concomitant use of the primary substrate DMADP between the plastidic synthesis of isoprene and isoprenoids with a larger molecular size, such as phytol residue of chlorophyll (C20) and carotenoids (C40; Ghirardo et al., 2014; Rasulov et al., 2014), and different from nonemitting species, isoprene emitters seem to support a much larger pool of DMADP without the onset of feedback inhibition (Ghirardo et al., 2014; Wright et al., 2014). However, it is poorly understood how inhibition of one branch of the pathway (isoprene versus larger isoprenoids) affects the other, to what extent it can lead to accumulation of phosphorylated intermediates, how it affects the overall pathway flux through the feedback regulation, and what is the possible role of cytosolic import and export of intermediates. These are all relevant questions to gain insight into the control of the partitioning of pathway flux between isoprene and larger isoprenoids and to understand the biological role of isoprene emission.Studies using metabolic inhibitors to deconvolute the factors involved in pathway regulation and understand the biological role of isoprene have so far used inhibitors that block the early steps of the corresponding pathways. In particular, fosmidomycin, a specific inhibitor of DXP reductoisomerase, the enzyme responsible for the synthesis of MEP from DXP, has been used to inhibit the DXP/MEP pathway (Loreto and Velikova, 2001; Sharkey et al., 2001; Loreto et al., 2004). In addition, lovastatin (mevinolin), the inhibitor of 3-hydroxy-3-methylglutaryl-CoA reductase that controls the MVA pathway flux, has been used to study the cooperativity of the two pathways (e.g. Laule et al., 2003; Mansouri and Salari, 2014). However, these inhibitors are not suitable to understand how end product accumulation can alter the pathway flux.Bisphosphonates constitute a promising class of inhibitors that could be particularly apt for studies on the effects of the inhibition of the end points of the pathway. They have been demonstrated to inhibit cytosolic farnesyl diphosphate (FDP) synthase activity (Oberhauser et al., 1998; Cromartie et al., 1999; van Beek et al., 1999; Bergstrom et al., 2000; Burke et al., 2004), аs well as geranyl diphosphate (GDP) and geranylgeranyl diphosphate (GGDP) synthase activities (Oberhauser et al., 1998; Cromartie et al., 1999; Kloer et al., 2006; No et al., 2012; Lindert et al., 2013). To our knowledge, bisphosphonates have not been used to study the effects of end product accumulation on the pathway flux in isoprene-emitting species, with the exception of one study that investigated the development of isoprene emission capacity through leaf ontogeny (Rasulov et al., 2014).A limitation with any inhibitor study could be a certain nonspecificity, inhibition of additional nondesired reactions, but so far there are no data on such nonspecificity of bisphosphonates. However, there is evidence that diphosphate and its analogs are inhibitors of any ferredoxin (Fd)-dependent reaction (Forti and Meyer, 1969; Bojko and Więckowski, 1999). This could be potentially relevant given that DXP/MEP pathway-reducing steps, at the level of 4-hydroxy-3-methyl-2-(E)-butenyl diphosphate (HMBDP) synthase (HDS) and reductase (HDR), directly accept electrons from Fd in light (Eisenreich et al., 2001; Seemann et al., 2006; Li and Sharkey, 2013a). In addition to the DXP/MEP pathway, inhibition at the level of Fd could also affect photosynthetic reactions and thereby alter energy supply for the DXP/MEP pathway.In this study, we have investigated the effects of inhibition of the initial and final steps of the DXP/MEP pathway by fosmidomycin and bisphosphonate inhibitors alendronate and zoledronate in a strong isoprene emitter hybrid aspen (Populus tremula × Populus tremuloides). Alendronate is a highly specific inhibitor of GDP (Lange et al., 2001; Burke et al., 2004) and FDP synthases (Bergstrom et al., 2000; Burke et al., 2004), and a less specific inhibitor of GGDP synthase (Szabo et al., 2002). Zoledronate operates similarly to alendronate, but is a much stronger inhibitor, being operationally active in concentrations several orders of magnitude less than alendronate (Lange et al., 2001; Henneman et al., 2011; Wasko, 2011). A unique in vivo method was used to study dynamic changes in DMADP and phosphorylated intermediate pool sizes (Rasulov et al., 2009a, 2011; Li et al., 2011), and different inhibitors were applied alone or in sequence to study the regulation of the pathway flux in conditions when the flux out of the pathway or into the pathway is curbed and when both the input and the output are curbed. Dynamic model calculations were used to quantitatively evaluate the significance of the cytosolic intermediate input into chloroplastic isoprenoid synthesis under different conditions of the DXP/MEP pathway entrance and exit, and to evaluate the possible nonspecific inhibition of other steps controlling DXP/MEP pathway flux. The study demonstrates the important regulation of DXP/MEP pathway input and output under conditions of end product accumulation and partial cooperativity among chloroplastic and cytosolic isoprenoid synthesis pathways.  相似文献   

9.
10.
We have established an efficient transient expression system with several vacuolar reporters to study the roles of endosomal sorting complex required for transport (ESCRT)-III subunits in regulating the formation of intraluminal vesicles of prevacuolar compartments (PVCs)/multivesicular bodies (MVBs) in plant cells. By measuring the distributions of reporters on/within the membrane of PVC/MVB or tonoplast, we have identified dominant negative mutants of ESCRT-III subunits that affect membrane protein degradation from both secretory and endocytic pathways. In addition, induced expression of these mutants resulted in reduction in luminal vesicles of PVC/MVB, along with increased detection of membrane-attaching vesicles inside the PVC/MVB. Transgenic Arabidopsis (Arabidopsis thaliana) plants with induced expression of ESCRT-III dominant negative mutants also displayed severe cotyledon developmental defects with reduced cell size, loss of the central vacuole, and abnormal chloroplast development in mesophyll cells, pointing out an essential role of the ESCRT-III complex in postembryonic development in plants. Finally, membrane dissociation of ESCRT-III components is important for their biological functions and is regulated by direct interaction among Vacuolar Protein Sorting-Associated Protein20-1 (VPS20.1), Sucrose Nonfermenting7-1, VPS2.1, and the adenosine triphosphatase VPS4/SUPPRESSOR OF K+ TRANSPORT GROWTH DEFECT1.Endomembrane trafficking in plant cells is complicated such that secretory, endocytic, and recycling pathways are usually integrated with each other at the post-Golgi compartments, among which, the trans-Golgi network (TGN) and prevacuolar compartment (PVC)/multivesicular body (MVB) are best studied (Tse et al., 2004; Lam et al., 2007a, 2007b; Müller et al., 2007; Foresti and Denecke, 2008; Hwang, 2008; Otegui and Spitzer, 2008; Robinson et al., 2008; Richter et al., 2009; Ding et al., 2012; Gao et al., 2014). Following the endocytic trafficking of a lipophilic dye, FM4-64, the TGN and PVC/MVB are sequentially labeled and thus are defined as the early and late endosome, respectively, in plant cells (Lam et al., 2007a; Chow et al., 2008). While the TGN is a tubular vesicular-like structure that may include several different microdomains and fit its biological function as a sorting station (Chow et al., 2008; Kang et al., 2011), the PVC/MVB is 200 to 500 nm in size with multiple luminal vesicles of approximately 40 nm (Tse et al., 2004). Membrane cargoes destined for degradation are sequestered into these tiny luminal vesicles and delivered to the lumen of the lytic vacuole (LV) via direct fusion between the PVC/MVB and the LV (Spitzer et al., 2009; Viotti et al., 2010; Cai et al., 2012). Therefore, the PVC/MVB functions between the TGN and LV as an intermediate organelle and decides the fate of membrane cargoes in the LV.In yeast (Saccharomyces cerevisiae), carboxypeptidase S (CPS) is synthesized as a type II integral membrane protein and sorted from the Golgi to the lumen of the vacuole (Spormann et al., 1992). Genetic analyses on the trafficking of CPS have led to the identification of approximately 17 class E genes (Piper et al., 1995; Babst et al., 1997, 2002a, 2002b; Odorizzi et al., 1998; Katzmann et al., 2001) that constitute the core endosomal sorting complex required for transport (ESCRT) machinery. The evolutionarily conserved ESCRT complex consists of several functionally different subcomplexes, ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III and the ESCRT-III-associated/Vacuolar Protein Sorting4 (VPS4) complex. Together, they form a complex protein-protein interaction network that coordinates sorting of cargoes and inward budding of the membrane on the MVB (Hurley and Hanson, 2010; Henne et al., 2011). Cargo proteins carrying ubiquitin signals are thought to be passed from one ESCRT subcomplex to the next, starting with their recognition by ESCRT-0 (Bilodeau et al., 2002, 2003; Hislop and von Zastrow, 2011; Le Bras et al., 2011; Shields and Piper, 2011; Urbé, 2011). ESCRT-0 recruits the ESCRT-I complex, a heterotetramer of VPS23, VPS28, VPS37, and MVB12, from the cytosol to the endosomal membrane (Katzmann et al., 2001, 2003). The C terminus of VPS28 interacts with the N terminus of VPS36, a member of the ESCRT-II complex (Kostelansky et al., 2006; Teo et al., 2006). Then, cargoes passed from ESCRT-I and ESCRT-II are concentrated in certain membrane domains of the endosome by ESCRT-III, which includes four coiled-coil proteins and is sufficient to induce the membrane invagination (Babst et al., 2002b; Saksena et al., 2009; Wollert et al., 2009). Finally, the ESCRT components are disassociated from the membrane by the adenosine triphosphatase (ATPase) associated with diverse cellular activities (AAA) VPS4/SUPPRESSOR OF K+ TRANSPORT GROWTH DEFECT1 (SKD1) before releasing the internal vesicles (Babst et al., 1997, 1998).Putative homologs of ESCRT-I–ESCRT-III and ESCRT-III-associated components have been identified in plants, except for ESCRT-0, which is only present in Opisthokonta (Winter and Hauser, 2006; Leung et al., 2008; Schellmann and Pimpl, 2009). To date, only a few plant ESCRT components have been studied in detail. The Arabidopsis (Arabidopsis thaliana) AAA ATPase SKD1 localized to the PVC/MVB and showed ATPase activity that was regulated by Lysosomal Trafficking Regulator-Interacting Protein5, a plant homolog of Vps Twenty Associated1 Protein (Haas et al., 2007). Expression of the dominant negative form of SKD1 caused an increase in the size of the MVB and a reduction in the number of internal vesicles (Haas et al., 2007). This protein also contributes to the maintenance of the central vacuole and might be associated with cell cycle regulation, as leaf trichomes expressing its dominant negative mutant form lost the central vacuole and frequently contained multiple nuclei (Shahriari et al., 2010). Double null mutants of CHARGED MULTIVESICULAR BODY PROTEIN, chmp1achmp1b, displayed severe growth defects and were seedling lethal. This may be due to the mislocalization of plasma membrane (PM) proteins, including those involved in auxin transport such as PINFORMED1, PINFORMED2, and AUXIN-RESISTANT1, from the vacuolar degradation pathway to the tonoplast of the LV (Spitzer et al., 2009).Plant ESCRT components usually contain several homologs, with the possibility of functional redundancy. Single mutants of individual ESCRT components may not result in an obvious phenotype, whereas knockout of all homologs of an ESCRT component by generating double or triple mutants may be lethal to the plant. As a first step to carry out systematic analysis on each ESCRT complex in plant cells, here, we established an efficient analysis system to monitor the localization changes of four vacuolar reporters that accumulate either in the lumen (LRR84A-GFP, EMP12-GFP, and aleurain-GFP) or on the tonoplast (GFP-VIT1) of the LV and identified several ESCRT-III dominant negative mutants. We reported that ESCRT-III subunits were involved in the release of PVC/MVB’s internal vesicles from the limiting membrane and were required for membrane protein degradation from secretory and endocytic pathways. In addition, transgenic Arabidopsis plants with induced expression of ESCRT-III dominant negative mutants showed severe cotyledon developmental defects. We also showed that membrane dissociation of ESCRT-III subunits was regulated by direct interaction with SKD1.  相似文献   

11.
12.
13.
14.
Red algae represent an evolutionarily important group that gave rise to the whole red clade of photosynthetic organisms. They contain a unique combination of light-harvesting systems represented by a membrane-bound antenna and by phycobilisomes situated on thylakoid membrane surfaces. So far, very little has been revealed about the mobility of their phycobilisomes and the regulation of their light-harvesting system in general. Therefore, we carried out a detailed analysis of phycobilisome dynamics in several red alga strains and compared these results with the presence (or absence) of photoprotective mechanisms. Our data conclusively prove phycobilisome mobility in two model mesophilic red alga strains, Porphyridium cruentum and Rhodella violacea. In contrast, there was almost no phycobilisome mobility in the thermophilic red alga Cyanidium caldarium that was not caused by a decrease in lipid desaturation in this extremophile. Experimental data attributed this immobility to the strong phycobilisome-photosystem interaction that highly restricted phycobilisome movement. Variations in phycobilisome mobility reflect the different ways in which light-harvesting antennae can be regulated in mesophilic and thermophilic red algae. Fluorescence changes attributed in cyanobacteria to state transitions were observed only in mesophilic P. cruentum with mobile phycobilisomes, and they were absent in the extremophilic C. caldarium with immobile phycobilisomes. We suggest that state transitions have an important regulatory function in mesophilic red algae; however, in thermophilic red algae, this process is replaced by nonphotochemical quenching.Photosynthetic light reactions are mediated by pigment-binding protein complexes located either inside the thylakoid membrane (e.g. chlorophyll-binding proteins of both photosystems) or associated on the membrane surface (e.g. phycobilisomes [PBsomes] in cyanobacteria and red algae). Recent progress in structural biology has allowed the construction of high-resolution structural models of most photosynthetic protein complexes (for review, see Fromme, 2008) together with their large-scale organization into supercomplexes (for review, see Dekker and Boekema, 2005). However, the dynamics of these supercomplexes and the mobility of particular light-harvesting proteins in vivo are still poorly understood (for review, see Mullineaux, 2008a; Kaňa, 2013; Kirchhoff, 2014) The importance of protein mobility in various photosynthetic processes, like nonphotochemical quenching and state transitions, has been explored mostly based on indirect in vitro experiments, including single-particle analysis (Kouřil et al., 2005), or by biochemical methods (Betterle et al., 2009; Caffarri et al., 2009). Recent studies on the mobility of light-harvesting proteins using live-cell imaging (for review, see Mullineaux, 2008a; Kaňa, 2013) have elucidated the importance of protein mobility for photosynthetic function (Joshua and Mullineaux, 2004; Joshua et al., 2005; Goral et al., 2010, 2012; Johnson et al., 2011). In addition, the redistribution of respiratory complexes in cyanobacterial thylakoid membranes plays an essential role in controlling electron flow (Liu et al., 2012).It is generally accepted that the mobility of most of the transmembrane photosynthetic proteins is very restricted in the thylakoid. The typical effective diffusion coefficient of photosynthetic proteins is somewhere between 0.01 and 0.001 μm−2 s−1 (Kaňa, 2013). A similar restriction in membrane protein mobility has also been described for bacterial membranes (Dix and Verkman, 2008; Mika and Poolman, 2011). In fact, this is very different in comparison with what we know for other eukaryotic membranes (e.g. plasma membrane and endoplasmic reticulum), where membrane-protein diffusion can be faster by 1 or 2 orders of magnitude (Lippincott-Schwartz et al., 2001). Therefore, macromolecular crowding of proteins has been used to rationalize the restricted protein mobility in thylakoid membranes of chloroplasts (Kirchhoff, 2008a, 2008b). Indeed, atomic force microscopy studies have shown that there is a dense packing and interaction of complexes in the photosynthetic membranes (Liu et al., 2011). Therefore, the diffusion of photosynthetic proteins in the thylakoid membrane is rather slow, and it increases only in less crowded parts of thylakoids (Kirchhoff et al., 2013). The current model of photosynthetic protein mobility thus proposes the immobility of protein supercomplexes, such as PSII (Mullineaux et al., 1997; Kirchhoff, 2008b), with only a small mobile fraction of chlorophyll-binding proteins represented by external antennae of photosystems, including light harvesting complex of PSII in higher plants (Consoli et al., 2005; Kirchhoff et al., 2008) or iron stress-induced chlorophyll-binding protein A in cyanobacteria (Sarcina and Mullineaux, 2004).The restricted mobility of internal membrane supercomplexes (photosystems) contrasts with the relatively mobile PBsomes (Mullineaux et al., 1997; Sarcina et al., 2001). PBsomes are sizeable biliprotein supercomplexes (5–10 MD) attached to the thylakoid membrane surface with dimensions of approximately 64 × 42 × 28 nm (length × width × height; Arteni et al., 2008; Liu et al., 2008a). PBsomes are composed of chromophore-bearing phycobiliproteins and colorless linker polypeptides (Adir, 2005; Liu et al., 2005). They serve as the main light-harvesting antennae in various species, including cyanobacteria, red algae, glaucocystophytes, and cryptophytes. Although a single PBsome is composed of hundreds of biliproteins, absorbed light energy is efficiently transferred toward a specific biliprotein that functions as a terminal energy emitter (Glazer, 1989). From there, energy can be transferred to either PSI or PSII and used in photosynthesis (Mullineaux et al., 1990; Mullineaux, 1992, 1994). In typical prokaryotic cyanobacteria and eukaryotic red algae, PBsomes are composed of two main parts: (1) allophycocyanin (APC) core proteins adjacent to the thylakoid membrane; and (2) peripheral rod proteins made from phycocyanin only or from a combination of phycocyanin together with phycoerythrin. Such complex and modular composition allows for different spectroscopic properties of PBsomes and thus their complementary absorption in the spectral region that is not covered by chlorophyll-binding proteins.PBsome mobility has been studied only in a few types of cyanobacteria (for review, see Kaňa, 2013). PBsomes have been recognized as a mobile element with an effective diffusion coefficient of about 0.03 μm2 s−1 for Synechococcus sp. PCC 7942 (Mullineaux et al., 1997; Sarcina et al., 2001). The effective diffusion coefficient value depends on lipid composition, temperature, and the size of the PBsome (Sarcina et al., 2001). The diffusion coefficient reflects PBsome mobility, but it is not affected singularly by physical diffusion processes, and the role of PBsome-photosystem interaction is an open question (Kaňa, 2013). PBsome mobility seems to be related to the requirement of light-induced PBsome redistribution during state transitions (Joshua and Mullineaux, 2004). The mechanism of state transitions in cyanobacteria is still rather questionable (for review, see Kirilovsky et al., 2014). As PSII seems to be immobile, it has been suggested that PBsomes interact with photosystems only transiently and that physical redistribution (diffusion) of PBsomes is crucial for the state transition (Mullineaux et al., 1997). The importance of such long-distance diffusions, however, should be tested experimentally in more detail (Kaňa, 2013), as an alternative theory of the state transition proposed only slight PBsome movement (shifting) between photosystems (McConnell et al., 2002). However, in both cases, PBsome mobility (i.e. the PBsome’s ability to move) is required (Kaňa, 2013).Red algae are the eukaryotic representatives of phototrophs containing PBsomes (Su et al., 2010). They represent the ancestor of photosynthetic microorganisms from the red clade of photosynthesis (Yoon et al., 2006; Wang et al., 2013), which includes various model organisms such as diatoms, chromerids, or dinoflagellates. Red algae contain a unique combination of antennae systems on their membrane surfaces, which are formed mostly by hemispherical PBsomes (Mimuro and Kikuchi, 2003; Arteni et al., 2008). Red algae also contain transmembrane light-harvesting antennae (Vanselow et al., 2009; Neilson and Durnford, 2010; Green, 2011) associated mostly with PSI (Wolfe et al., 1994). Therefore, red algae represent a functionally important eukaryotic model organism; however, few facts are known about the regulation of its light-harvesting efficiency, although it seems to be connected with photoprotection in the reaction center (Delphin et al., 1996, 1998; Krupnik et al., 2013). The presence of photoprotective NPQ in PBsomes of prokaryotic cyanobacteria has been conclusively proven (Kirilovsky et al., 2014); however, this mechanism seems to be missing in eukaryotic phycobiliproteins of cryptophytes (Kaňa et al., 2012b) and red algae. Moreover, the presence (or absence) of PBsome mobility has not been confirmed conclusively (Liu et al., 2009).Therefore, we carried out a detailed study of PBsome mobility in red algal chloroplasts to determine the role of mobility in the regulation of light-harvesting efficiency. We found that red alga PBsomes are a mobile protein complex with effective diffusion coefficient between 2.7 × 10−3 and 13 × 10−3 μm−2 s−1 in all studied mesophilic strains. It contrasted with PBsomes in extremophilic red algal strains (Cyanidium caldarium), where PBsome mobility under physiological conditions was highly restricted (effective diffusion coefficient of approximately 0.6 × 10−3 μm−2 s−1). The restriction of PBsome mobility in extremophilic C. caldarium was due to a tight interaction of PBsomes with both photosystems and not to changes in lipid desaturation, an effect typical for extremophiles. The PBsome-photosystem interaction was weakened for C. caldarium grown at suboptimal temperatures, resulting in a pronounced increase in PBsome mobility thanks to PBsome decoupling from the photosystem. This result shows that PBsome mobility in this strain is limited by the strength of the PBsome-photosystem interaction rather than by the restriction of diffusion by factors such as macromolecular crowding. Moreover, our study allows us to describe two different models of light-harvesting antenna regulation in red algae. In mesophilic strains (Porphyridium cruentum and Rhodella violacea), absorbed light is redistributed between photosystems in a process of state transition that requires PBsome mobility. On the contrary, in extremophilic C. caldarium, PBsome are strongly coupled to photosystems and excess light is dissipated by a process of nonphotochemical quenching, as has been described recently (Krupnik et al., 2013).  相似文献   

15.
Heavy metal-transporting P-type ATPase (HMA) has been implicated in the transport of heavy metals in plants. Here, we report the function and role of an uncharacterized member of HMA, OsHMA5 in rice (Oryza sativa). Knockout of OsHMA5 resulted in a decreased copper (Cu) concentration in the shoots but an increased Cu concentration in the roots at the vegetative stage. At the reproductive stage, the concentration of Cu in the brown rice was significantly lower in the mutants than in the wild-type rice; however, there was no difference in the concentrations of iron, manganese, and zinc between two independent mutants and the wild type. The Cu concentration of xylem sap was lower in the mutants than in the wild-type rice. OsHMA5 was mainly expressed in the roots at the vegetative stage but also in nodes, peduncle, rachis, and husk at the reproductive stage. The expression was up-regulated by excess Cu but not by the deficiency of Cu and other metals, including zinc, iron, and manganese, at the vegetative stage. Analysis of the transgenic rice carrying the OsHMA5 promoter fused with green fluorescent protein revealed that it was localized at the root pericycle cells and xylem region of diffuse vascular bundles in node I, vascular tissues of peduncle, rachis, and husk. Furthermore, immunostaining with an antibody against OsHMA5 revealed that it was localized to the plasma membrane. Expression of OsHMA5 in a Cu transport-defective mutant yeast (Saccharomyces cerevisiae) strain restored the growth. Taken together, OsHMA5 is involved in loading Cu to the xylem of the roots and other organs.Plants require nutrient elements to maintain normal growth and development. A number of different transporters, such as Cation Diffusion Facilitator, Natural resistance-associated macrophage protein, ATP-Binding Cassette, Zinc- and Iron-regulated-like Protein, and P-type ATPase, have been reported to be involved in the uptake, translocation, distribution, and homeostasis of nutrients (Hall and Williams, 2003; Krämer et al., 2007; Palmer and Guerinot, 2009). Among them, heavy metal-transporting P-type ATPase (HMA), the P1B subfamily of the P-type ATPase superfamily, has been implicated in heavy metal transport (Williams and Mills, 2005; Grotz and Guerinot, 2006; Argüello et al., 2007; Burkhead et al., 2009). There are eight and nine members of P1B-ATPase in Arabidopsis thaliana and rice (Oryza sativa), respectively (Williams and Mills, 2005). They are divided into two groups: zinc (Zn)/cadmium (Cd)/cobalt/lead (Pb) and copper (Cu)/silver transporters (Williams and Mills, 2005). AtHMA1 to AtHMA4 in A. thaliana and OsHMA1 to OsHMA3 in rice belong to the former group, while AtHMA5 to AtHMA8 and OsHMA4 to OsHMA9 belong to the latter group, although AtHMA1 has also been shown to transport Zn, Cu, and calcium (Axelsen and Palmgren, 2001; Williams and Mills, 2005; Seigneurin-Berny et al., 2006; Moreno et al., 2008; Kim et al., 2009).All members of HMAs in A. thaliana have been functionally characterized. AtHMA1 is involved in delivering Cu to the stroma, exporting Zn2+ from the chloroplast, or as a Ca2+/heavy metal transporter to the intracellular organelle (Seigneurin-Berny et al., 2006; Moreno et al., 2008; Kim et al., 2009). AtHMA2 and AtHMA4 localized at the pericycle are partially redundant and responsible for the release of Zn into the xylem (xylem loading) as well as Cd (Hussain et al., 2004; Verret et al., 2004; Wong and Cobbett, 2009; Wong et al., 2009), while AtHMA3 localized at the tonoplast plays a role in the detoxification of Zn/Cd/cobalt/Pb by mediating them into the vacuole (Morel et al., 2009; Chao et al., 2012). On the other hand, AtHMA5 is involved in the Cu translocation from roots to shoots or Cu detoxification of roots (Andrés-Colás et al., 2006; Kobayashi et al., 2008). AtHMA6 (PAA1, for P-type ATPase of Arabidopsis1) localized at the chloroplast periphery has been proposed to transport Cu over the chloroplast envelope, whereas AtHMA8 (PAA2) localized at the thylakoid membranes most likely transports Cu into the thylakoid lumen to supply plastocyanin (Shikanai et al., 2003; Abdel-Ghany et al., 2005). Finally, AtHMA7 (RESPONSIVE-TO-ANTAGONIST1) is responsible for delivering Cu to ethylene receptors and Cu homeostasis in the seedlings (Hirayama et al., 1999; Woeste and Kieber, 2000; Binder et al., 2010).By contrast, only three out of nine P-type ATPase members have been functionally characterized in rice. OsHMA2 was recently reported to be involved in the root-shoot translocation of Zn and Cd (Satoh-Nagasawa et al., 2012; Takahashi et al., 2012; Yamaji et al., 2013). Furthermore, OsHMA2 at the node is required for preferential distribution of Zn to young leaves and panicles (Yamaji et al., 2013). OsHMA3 is localized to the tonoplast of the root cells and responsible for the sequestration of Cd into the vacuoles (Ueno et al., 2010; Miyadate et al., 2011). On the other hand, OsHMA9 was mainly expressed in vascular tissues, including the xylem and phloem (Lee et al., 2007). The knockout lines accumulated more Zn, Cu, Pb, and Cd, suggesting its role in the efflux of these metals from the cells (Lee et al., 2007).Some members of P-type ATPase have also been identified in other plant species, including barley (Hordeum vulgare), wheat (Triticum aestivum), Thlaspi caerulescens (Noccaea caerulescens), and Arabidopsis halleri. HvHMA1 from barley might be involved in mobilizing Zn and Cu during the stage of grain filling (Mikkelsen et al., 2012). HvHMA2 from barley and TaHMA2 from wheat showed similar functions as OsHMA2 in rice (Mills et al., 2012; Tan et al., 2013). AhHMA3 in A. halleri, a Zn hyperaccumulator, is probably involved in high Zn accumulation (Becher et al., 2004; Chiang et al., 2006). Furthermore, AhHMA4 for Zn translocation showed a higher expression level (Chiang et al., 2006; Hanikenne et al., 2008). On the other hand, TcHMA3 from ecotype Ganges of T. caerulescens, a Cd hyperaccumulator, plays an important role in the detoxification of Cd by sequestering Cd into the vacuole of the leaves (Ueno et al., 2011). High expression of TcHMA4 (NcHMA4) was also reported in T. caerulescens (Bernard et al., 2004; Papoyan and Kochian, 2004; Craciun et al., 2012).In this study, we investigated the function and role of an uncharacterized member of P-type ATPase in rice, OsHMA5. We found that OsHMA5 is involved in the xylem loading of Cu at both the vegetative and reproductive growth stages.  相似文献   

16.
17.
18.
19.
Two mutants sensitive to heat stress for growth and impaired in NADPH dehydrogenase (NDH-1)-dependent cyclic electron transport around photosystem I (NDH-CET) were isolated from the cyanobacterium Synechocystis sp. strain PCC 6803 transformed with a transposon-bearing library. Both mutants had a tag in the same sll0272 gene, encoding a protein highly homologous to NdhV identified in Arabidopsis (Arabidopsis thaliana). Deletion of the sll0272 gene (ndhV) did not influence the assembly of NDH-1 complexes and the activities of CO2 uptake and respiration but reduced the activity of NDH-CET. NdhV interacted with NdhS, a ferredoxin-binding subunit of cyanobacterial NDH-1 complex. Deletion of NdhS completely abolished NdhV, but deletion of NdhV had no effect on the amount of NdhS. Reduction of NDH-CET activity was more significant in ΔndhS than in ΔndhV. We therefore propose that NdhV cooperates with NdhS to accept electrons from reduced ferredoxin.Cyanobacterial NADPH dehydrogenase (NDH-1) complexes are localized in the thylakoid membrane (Ohkawa et al., 2001, 2002; Zhang et al., 2004; Xu et al., 2008; Battchikova et al., 2011b) and participate in a variety of bioenergetic reactions, such as respiration, cyclic electron transport around photosystem I (NDH-CET), and CO2 uptake (Ogawa, 1991; Mi et al., 1992; Ohkawa et al., 2000). Structurally, the cyanobacterial NDH-1 complexes closely resemble energy-converting complex I in eubacteria and the mitochondrial respiratory chain regardless of the absence of homologs of three subunits in cyanobacterial genomes that constitute the catalytically active core of complex I (Friedrich et al., 1995; Friedrich and Scheide, 2000; Arteni et al., 2006). Over the past decade, new subunits of NDH-1 complexes specific to oxygenic photosynthesis have been identified in several cyanobacterial strains. They are NdhM to NdhQ and NdhS (Prommeenate et al., 2004; Battchikova et al., 2005, 2011b; Nowaczyk et al., 2011; Wulfhorst et al., 2014; Zhang et al., 2014; Zhao et al., 2014b, 2015), in addition to NdhL first identified in the cyanobacterium Synechocystis sp. strain PCC 6803 (hereafter Synechocystis 6803) about 20 years ago (Ogawa, 1992). Among them, NdhS possesses a ferredoxin (Fd)-binding motif and was shown to bind Fd, which suggested that Fd is one of the electron donors to NDH-1 complexes (Mi et al., 1995; Battchikova et al., 2011b; Ma and Ogawa, 2015). Deletion of NdhS strongly reduced the activity of NDH-CET but had no effect on respiration and CO2 uptake (Battchikova et al., 2011b; Ma and Ogawa, 2015). The NDH-CET plays an important role in coping with various environmental stresses regardless of its elusive mechanism. For example, this function can greatly alleviate heat-sensitive growth phenotypes (Wang et al., 2006a; Zhao et al., 2014a). Thus, heat treatment strategy can help in identifying the proteins essential to NDH-CET.Here, a new oxygenic photosynthesis-specific (OPS) subunit NdhV was identified in Synechocystis 6803 with the help of heat treatment strategy, and its deletion did not influence the assembly of NDH-1L and NDH-1MS complexes and the activities of CO2 uptake and respiration but impaired the NDH-CET activity. We give evidence that NdhV interacts with NdhS and is another component of Fd-binding domain of cyanobacterial NDH-1 complex. A possible role of NdhV on the NDH-CET activity is discussed.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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