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.
This study aims to elucidate the molecular mechanism of an alternative electron flow (AEF) functioning under suppressed (CO2-limited) photosynthesis in the cyanobacterium Synechocystis sp. PCC 6803. Photosynthetic linear electron flow, evaluated as the quantum yield of photosystem II [Y(II)], reaches a maximum shortly after the onset of actinic illumination. Thereafter, Y(II) transiently decreases concomitantly with a decrease in the photosynthetic oxygen evolution rate and then recovers to a rate that is close to the initial maximum. These results show that CO2 limitation suppresses photosynthesis and induces AEF. In contrast to the wild type, Synechocystis sp. PCC 6803 mutants deficient in the genes encoding FLAVODIIRON2 (FLV2) and FLV4 proteins show no recovery of Y(II) after prolonged illumination. However, Synechocystis sp. PCC 6803 mutants deficient in genes encoding proteins functioning in photorespiration show AEF activity similar to the wild type. In contrast to Synechocystis sp. PCC 6803, the cyanobacterium Synechococcus elongatus PCC 7942 has no FLV proteins with high homology to FLV2 and FLV4 in Synechocystis sp. PCC 6803. This lack of FLV2/4 may explain why AEF is not induced under CO2-limited photosynthesis in S. elongatus PCC 7942. As the glutathione S-transferase fusion protein overexpressed in Escherichia coli exhibits NADH-dependent oxygen reduction to water, we suggest that FLV2 and FLV4 mediate oxygen-dependent AEF in Synechocystis sp. PCC 6803 when electron acceptors such as CO2 are not available.In photosynthesis, photon energy absorbed by PSI and PSII in thylakoid membranes oxidizes the reaction center chlorophylls (Chls), P700 in PSI and P680 in PSII, and drives the photosynthetic electron transport (PET) system. In PSII, water is oxidized to oxygen as the oxidized P680 accepts electrons from water. These electrons then reduce the cytochrome b6/f complex through plastoquinone (PQ) in the thylakoid membranes. Photooxidized P700 in PSI accepts electrons from the reduced cytochrome b6/f complex through plastocyanin or cytochrome c6. Electrons released in the photooxidation of P700 are used to produce NADPH through ferredoxin and ferredoxin NADP+ reductase. Thus, electrons flow from water to NADPH in the so-called photosynthetic linear electron flow (LEF). Importantly, LEF induces a proton gradient across the thylakoid membranes, which provides the driving force for ATP production by ATP synthases in the thylakoid membranes. NADPH and ATP serve as chemical energy donors in the photosynthetic carbon reduction cycle (Calvin cycle).It recently has been proposed that, in cyanobacteria, the photorespiratory carbon oxidation cycle (photorespiration) functions simultaneously with the Calvin cycle to recover carbon for the regeneration of ribulose-1,5-bisphosphate, one of the substrates of Rubisco (Hagemann et al., 2013). Rubisco catalyzes the primary reactions of carbon reduction as well as oxidation cycles. However, the presence of a specific carbon concentration mechanism (CCM) in cyanobacteria had been thought to prevent the operation of photorespiration. CCM maintains a high concentration of CO2 around Rubisco so that the oxygenase activity of Rubisco is suppressed (Badger and Price, 1992). However, recent studies on mutants deficient in photorespiration enzymes have shown that photorespiration functions, particularly under CO2-limited conditions, in cyanobacteria as it does in higher plants (Eisenhut et al., 2006, 2008).Decreased consumption of NADPH under CO2-limited or high-light conditions causes electrons to accumulate in the PET system. As a result, the photooxidation and photoreduction cycles of the reaction center Chls in PSI and PSII become uncoupled from the production of NADPH, inducing alternative electron flow (AEF) pathways (Mullineaux, 2014). In cyanobacteria, several AEFs that differ from those in higher plants are proposed to function as electron sinks (Mullineaux, 2014). Electrons accumulated in the PET system flow to oxygen through FLAVODIIRON1 (FLV1) and FLV3 proteins in PSI and the terminal oxidase, cytochrome c oxidase complex, and cytochrome bd-quinol oxidase (Pils and Schmetterer, 2001; Berry et al., 2002; Helman et al., 2003; Nomura et al., 2006; Lea-Smith et al., 2013). Cyanobacterial FLV comprises a diiron center, a flavodoxin domain with an FMN-binding site, and a flavin reductase domain (Vicente et al., 2002). In Synechocystis sp. PCC 6803, Helman et al. (2003) identified four genes encoding FLV1 to FLV4 and showed that FLV1 and FLV3 were essential for the photoreduction of oxygen by PSI. FLV1 and FLV3 were proposed to function as a heterodimer (Allahverdiyeva et al., 2013). FLV2/4 have been proposed to function in energy dissipation associated with PSII (Zhang et al., 2012). In addition, hydrogenases convert H+ to H2 with NADPH as an electron donor (Appel et al., 2000). Furthermore, Flores et al. (2005) suggested that the nitrate assimilation pathway functions in AEF when the cells live in medium containing nitrate.To elucidate the physiological functions of these AEFs, evaluation of the presence and capacity of each AEF pathway is required. Therefore, in vivo analyses of electron fluxes are essential. We had found that an electron flow uncoupled from photosynthetic oxygen evolution functioned under suppressed (CO2-limited) photosynthesis in the cyanobacterium Synechocystis sp. PCC 6803 but not in Synechococcus elongatus PCC 7942 (Hayashi et al., 2014), indicating that an AEF operated in Synechocystis sp. PCC 6803. This AEF was induced in high-[CO2]-grown Synechocystis sp. PCC 6803 during the transition from CO2-saturated photosynthesis to CO2-limited photosynthesis (Hayashi et al., 2014). In contrast, in Synechocystis sp. PCC 6803 grown at ambient CO2 concentration, AEF was detected immediately following the transition to CO2-limited photosynthesis (Hayashi et al., 2014), suggesting that AEF was already induced under ambient atmospheric conditions.The expression of the AEF activity observed under CO2-limited photosynthesis required the presence of oxygen in Synechocystis sp. PCC 6803 (Hayashi et al., 2014). In Synechocystis sp. PCC 6803, FLV1/3 were proposed to catalyze the photoreduction of oxygen (Helman et al., 2003). However, Hayashi et al. (2014) found no evidence that FLV1/3 operated under CO2-limited photosynthesis: a mutant Synechocystis sp. PCC 6803 deficient in FLV1/3 maintained almost constant electron flux under CO2-limited photosynthesis after the transition from CO2-saturated conditions. Thus, the postulated photoreduction of oxygen by FLV1/3 was not responsible for the electron flux observed under CO2-limited photosynthesis in Synechocystis sp. PCC 6803.In this study, we aimed to elucidate the molecular mechanism of the oxygen-dependent AEF functioning under CO2-limited photosynthesis in Synechocystis sp. PCC 6803. The possibility that FLV2 and FLV4 catalyze the photoreduction of oxygen under CO2-limited photosynthesis could not be excluded, given that AEF in high-[CO2]-grown Synechocystis sp. PCC 6803 was induced following the transition to CO2-limited photosynthesis (Hayashi et al., 2014). Both FLV2 and FLV4 are predicted to possess oxidoreductase motifs, similar to FLV1 and FLV3 (Helman et al., 2003; Zhang et al., 2012). Furthermore, the expression of two FLV genes (flv2 and flv4) was enhanced under low-[CO2] conditions (Zhang et al., 2009). Zhang et al. (2012) proposed that FLV2 and FLV4 did not donate electrons to oxygen on the basis of the finding that the Synechocystis sp. PCC 6803 mutants deficient in FLV1/3 showed no light-dependent oxygen uptake (Helman et al., 2003). However, Helman et al. (2003) cultivated Synechocystis sp. PCC 6803 strains deficient in FLV1 and FLV3 proteins under high-[CO2] conditions, and we cannot exclude the possibility that the FLV2 and FLV4 proteins were not produced in the studied cells. Taken together, it seems plausible that FLV2 and FLV4 mediate oxygen-dependent AEF following the transition to CO2-limited photosynthesis. To evaluate this possibility, we constructed Synechocystis sp. PCC 6803 mutants deficient in flv2 and flv4 and measured their oxygen evolution and Chl fluorescence simultaneously. The mutants showed suppressed LEF after transition to CO2-limited photosynthesis, similar to S. elongatus PCC 7942. We also tested the possibility that photorespiration functions as an electron sink under CO2-limited photosynthesis in Synechocystis sp. PCC 6803. A recent study revealed photorespiratory oxygen uptake in a flv1/3 mutant under CO2-depleted conditions (Allahverdiyeva et al., 2011). In this study, we found that the quantum yield of photosystem II [Y(II)] of mutants deficient in genes encoding proteins that function in photorespiration was similar to that of wild-type Synechocystis sp. PCC 6803. Thus, FLV2 and FLV4 appear to function in the oxygen-dependent AEF under CO2-limited photosynthesis in Synechocystis sp. PCC 6803. This inference is further supported by the lack of FLV2 and FLV4 homologs in the genome of S. elongatus PCC 7942 (Bersanini et al., 2014). In addition, we found oxygen-reducing activities of recombinant glutathione S-transferase (GST)-FLV4 fusion protein, similar to those of recombinant FLV3 protein (Vicente et al., 2002). In light of these results, we discuss the molecular mechanism of the oxygen-dependent AEF under CO2-limited photosynthesis and the physiological function of FLV proteins in Synechocystis sp. PCC 6803.  相似文献   

2.
The chloroplast-encoded low molecular weight protein PsbN is annotated as a photosystem II (PSII) subunit. To elucidate the localization and function of PsbN, encoded on the opposite strand to the psbB gene cluster, we raised antibodies and inserted a resistance cassette into PsbN in both directions. Both homoplastomic tobacco (Nicotiana tabacum) mutants ∆psbN-F and ∆psbN-R show essentially the same PSII deficiencies. The mutants are extremely light sensitive and failed to recover from photoinhibition. Although synthesis of PSII proteins was not altered significantly, both mutants accumulated only ∼25% of PSII proteins compared with the wild type. Assembly of PSII precomplexes occurred at normal rates, but heterodimeric PSII reaction centers (RCs) and higher order PSII assemblies were not formed efficiently in the mutants. The ∆psbN-R mutant was complemented by allotopic expression of the PsbN gene fused to the sequence of a chloroplast transit peptide in the nuclear genome. PsbN represents a bitopic trans-membrane peptide localized in stroma lamellae with its highly conserved C terminus exposed to the stroma. Significant amounts of PsbN were already present in dark-grown seedling. Our data prove that PsbN is not a constituent subunit of PSII but is required for repair from photoinhibition and efficient assembly of the PSII RC.  相似文献   

3.
Under high-irradiance conditions, plants must efficiently protect photosystem II (PSII) from damage. In this study, we demonstrate that the chloroplast protein HYPERSENSITIVE TO HIGH LIGHT1 (HHL1) is expressed in response to high light and functions in protecting PSII against photodamage. Arabidopsis thaliana hhl1 mutants show hypersensitivity to high light, drastically decreased PSII photosynthetic activity, higher nonphotochemical quenching activity, a faster xanthophyll cycle, and increased accumulation of reactive oxygen species following high-light exposure. Moreover, HHL1 deficiency accelerated the degradation of PSII core subunits under high light, decreasing the accumulation of PSII core subunits and PSII–light-harvesting complex II supercomplex. HHL1 primarily localizes in the stroma-exposed thylakoid membranes and associates with the PSII core monomer complex through direct interaction with PSII core proteins CP43 and CP47. Interestingly, HHL1 also directly interacts, in vivo and in vitro, with LOW QUANTUM YIELD OF PHOTOSYSTEM II1 (LQY1), which functions in the repair and reassembly of PSII. Furthermore, the hhl1 lqy1 double mutants show increased photosensitivity compared with single mutants. Taken together, these results suggest that HHL1 forms a complex with LQY1 and participates in photodamage repair of PSII under high light.  相似文献   

4.
Photosystem II (PSII) requires constant disassembly and reassembly to accommodate replacement of the D1 protein. Here, we characterize Arabidopsis thaliana MET1, a PSII assembly factor with PDZ and TPR domains. The maize (Zea mays) MET1 homolog is enriched in mesophyll chloroplasts compared with bundle sheath chloroplasts, and MET1 mRNA and protein levels increase during leaf development concomitant with the thylakoid machinery. MET1 is conserved in C3 and C4 plants and green algae but is not found in prokaryotes. Arabidopsis MET1 is a peripheral thylakoid protein enriched in stroma lamellae and is also present in grana. Split-ubiquitin assays and coimmunoprecipitations showed interaction of MET1 with stromal loops of PSII core components CP43 and CP47. From native gels, we inferred that MET1 associates with PSII subcomplexes formed during the PSII repair cycle. When grown under fluctuating light intensities, the Arabidopsis MET1 null mutant (met1) showed conditional reduced growth, near complete blockage in PSII supercomplex formation, and concomitant increase of unassembled CP43. Growth of met1 in high light resulted in loss of PSII supercomplexes and accelerated D1 degradation. We propose that MET1 functions as a CP43/CP47 chaperone on the stromal side of the membrane during PSII assembly and repair. This function is consistent with the observed differential MET1 accumulation across dimorphic maize chloroplasts.  相似文献   

5.
Efficient assembly and repair of the oxygen-evolving photosystem II (PSII) complex is vital for maintaining photosynthetic activity in plants, algae, and cyanobacteria. How chlorophyll is delivered to PSII during assembly and how vulnerable assembly complexes are protected from photodamage are unknown. Here, we identify a chlorophyll and β-carotene binding protein complex in the cyanobacterium Synechocystis PCC 6803 important for formation of the D1/D2 reaction center assembly complex. It is composed of putative short-chain dehydrogenase/reductase Ycf39, encoded by the slr0399 gene, and two members of the high-light-inducible protein (Hlip) family, HliC and HliD, which are small membrane proteins related to the light-harvesting chlorophyll binding complexes found in plants. Perturbed chlorophyll recycling in a Ycf39-null mutant and copurification of chlorophyll synthase and unassembled D1 with the Ycf39-Hlip complex indicate a role in the delivery of chlorophyll to newly synthesized D1. Sequence similarities suggest the presence of a related complex in chloroplasts.  相似文献   

6.
The cytochrome b6f (cytb6f) complex plays a central role in photosynthesis, coupling electron transport between photosystem II (PSII) and photosystem I to the generation of a transmembrane proton gradient used for the biosynthesis of ATP. Photosynthesis relies on rapid shuttling of electrons by plastoquinone (PQ) molecules between PSII and cytb6f complexes in the lipid phase of the thylakoid membrane. Thus, the relative membrane location of these complexes is crucial, yet remains unknown. Here, we exploit the selective binding of the electron transfer protein plastocyanin (Pc) to the lumenal membrane surface of the cytb6f complex using a Pc-functionalized atomic force microscope (AFM) probe to identify the position of cytb6f complexes in grana thylakoid membranes from spinach (Spinacia oleracea). This affinity-mapping AFM method directly correlates membrane surface topography with Pc-cytb6f interactions, allowing us to construct a map of the grana thylakoid membrane that reveals nanodomains of colocalized PSII and cytb6f complexes. We suggest that the close proximity between PSII and cytb6f complexes integrates solar energy conversion and electron transfer by fostering short-range diffusion of PQ in the protein-crowded thylakoid membrane, thereby optimizing photosynthetic efficiency.  相似文献   

7.
Photosynthetic organisms developed multiple strategies for balancing light-harvesting versus intracellular energy utilization to survive ever-changing environmental conditions. The light-harvesting complex (LHC) protein family is of paramount importance for this function and can form light-harvesting pigment protein complexes. In this work, we describe detailed analyses of the photosystem II (PSII) LHC protein LHCBM9 of the microalga Chlamydomonas reinhardtii in terms of expression kinetics, localization, and function. In contrast to most LHC members described before, LHCBM9 expression was determined to be very low during standard cell cultivation but strongly increased as a response to specific stress conditions, e.g., when nutrient availability was limited. LHCBM9 was localized as part of PSII supercomplexes but was not found in association with photosystem I complexes. Knockdown cell lines with 50 to 70% reduced amounts of LHCBM9 showed reduced photosynthetic activity upon illumination and severe perturbation of hydrogen production activity. Functional analysis, performed on isolated PSII supercomplexes and recombinant LHCBM9 proteins, demonstrated that presence of LHCBM9 resulted in faster chlorophyll fluorescence decay and reduced production of singlet oxygen, indicating upgraded photoprotection. We conclude that LHCBM9 has a special role within the family of LHCII proteins and serves an important protective function during stress conditions by promoting efficient light energy dissipation and stabilizing PSII supercomplexes.  相似文献   

8.
9.
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   

10.
11.
Yariv phenylglycosides [1,3,5-tri(p-glycosyloxyphenylazo)-2,4,6-trihydroxybenzene] are a group of chemical compounds that selectively bind to arabinogalactan proteins (AGPs), a type of plant proteoglycan. Yariv phenylglycosides are widely used as cytochemical reagents to perturb the molecular functions of AGPs as well as for the detection, quantification, purification, and staining of AGPs. However, the target structure in AGPs to which Yariv phenylglycosides bind has not been determined. Here, we identify the structural element of AGPs required for the interaction with Yariv phenylglycosides by stepwise trimming of the arabinogalactan moieties using combinations of specific glycoside hydrolases. Whereas the precipitation with Yariv phenylglycosides (Yariv reactivity) of radish (Raphanus sativus) root AGP was not reduced after enzyme treatment to remove α-l-arabinofuranosyl and β-glucuronosyl residues and β-1,6-galactan side chains, it was completely lost after degradation of the β-1,3-galactan main chains. In addition, Yariv reactivity of gum arabic, a commercial product of acacia (Acacia senegal) AGPs, increased rather than decreased during the repeated degradation of β-1,6-galactan side chains by Smith degradation. Among various oligosaccharides corresponding to partial structures of AGPs, β-1,3-galactooligosaccharides longer than β-1,3-galactoheptaose exhibited significant precipitation with Yariv in a radial diffusion assay on agar. A pull-down assay using oligosaccharides cross linked to hydrazine beads detected an interaction of β-1,3-galactooligosaccharides longer than β-1,3-galactopentaose with Yariv phenylglycoside. To the contrary, no interaction with Yariv was detected for β-1,6-galactooligosaccharides of any length. Therefore, we conclude that Yariv phenylglycosides should be considered specific binding reagents for β-1,3-galactan chains longer than five residues, and seven residues are sufficient for cross linking, leading to precipitation of the Yariv phenylglycosides.Arabinogalactan proteins (AGPs) are a type of plant proteoglycans consisting of a Hyp-rich core protein and large arabinogalactan (AG) moieties (Fincher et al., 1983; Nothnagel, 1997). Although there are many molecular species of AGP differentiated by their core proteins, the AG moieties commonly comprise β-1,3-galactan main chains and β-1,6-galactan side chains, to which l-Ara and other auxiliary sugars, such as GlcA, 4-O-methyl-GlcA, l-Fuc, l-Rha, and Xyl, are attached (Fincher et al., 1983; Nothnagel, 1997; Seifert and Roberts, 2007). A commercial product of AGPs prepared from the acacia (Acacia senegal) tree is known as gum arabic and utilized as a food stabilizer. In the Japanese herbal remedy Juzen-Taiho-To, AGs from Astragalus membranaceus are the active ingredient (Majewska-Sawka and Nothnagel, 2000; Kiyohara et al., 2002). In intact plants, AGPs are implicated in various physiological events and serve as extracellular constituents and signaling molecules. For instance, an AGP from stylar transmitting tissue attracts pollen tubes and stimulates their elongation in tobacco (Nicotiana tabacum; Cheung et al., 1995).Yariv phenylglycosides [1,3,5-tri(p-glycosyloxyphenylazo)-2,4,6-trihydroxybenzene] are a group of chemical compounds that were initially developed as carbohydrate antigens for the purification of anti-glycoside antibody and sugar-binding protein (Yariv et al., 1962, 1967a). It then turned out that Yariv phenylglycosides specifically precipitate AGPs (Yariv et al., 1967b; Jermyn and Yeow, 1975). The specific interaction of AGPs with Yariv phenylglycosides forming brown-red precipitate is called Yariv reactivity and has been recognized as an important criterion in the definition of AGPs, even though a number of AGPs do not exhibit Yariv reactivity. Nevertheless, the structure involved in the interaction with Yariv phenylglycoside is presumed to be conserved in many AGPs. The interaction of Yariv phenylglycosides with AGP depends on the glycosyl residues attached to the phenylazotrihydroxybenzene core. In particular, β-glucosyl Yariv phenylglycoside (β-Glc-Yariv) and β-galactosyl Yariv phenylglycoside (β-Gal-Yariv) bind to AGPs, whereas α-glucosyl Yariv and α-galactosyl Yariv (α-Gal-Yariv) do not bind to AGPs (Jermyn and Yeow, 1975; Larkin, 1977, 1978; Nothnagel and Lyon, 1986). Because of the specific interaction with the β-glycosyl Yariv phenylglycosides (β-Yarivs), AGPs were formerly called “β-lectins” (Jermyn and Yeow, 1975; Gleeson and Jermyn, 1979; Nothnagel and Lyon, 1986).The β-Yarivs are useful tools for staining, detection, and quantification of AGPs. Using β-Glc-Yariv, β-lectins were shown to exist in angiosperm, gymnosperm, fern, moss, and liverwort, illustrating the wide distribution of AGPs in the plant kingdom (Jermyn and Yeow, 1975; Clarke et al., 1978). In addition, β-Yarivs are also used as chemical reagents in the purification of AGPs. A nonclassical AGP, xylogen, which is a signaling molecule inducing the differentiation to tracheary elements, has been purified from the culture medium of zinnia (Zinnia elegans) cells by precipitation with β-Glc-Yariv (Motose et al., 2004). As the treatment with β-Yarivs causes the perturbation of various physiological processes in plants, β-Yarivs are reliable cytochemical reagents to explore AGP functions. Application of β-Yarivs to cultured cells of Arabidopsis (Arabidopsis thaliana) induced programmed cell death, demonstrating the involvement of AGPs in the determination of cell fate (Gao and Showalter, 1999). In tobacco cultured cells, the treatment with β-Yarivs has indicated a possible role of AGPs in the orientation of cortical microtubules and the polymerization of F-actin (Sardar et al., 2006).Although Yariv phenylglycosides have been extensively utilized in studies of AGPs over 40 years, the identification of the target structures on AGPs required for β-Yariv reactivity remains elusive (Nothnagel, 1997; Seifert and Roberts, 2007). It has been proposed that β-Yarivs bind to the Hyp-rich core protein, based on the observation that deglycosylation treatment with hydrogen fluoride did not abolish the Yariv reactivity of gum arabic and a tobacco AGP (Akiyama et al., 1984). To the contrary, other reports have asserted the importance of the carbohydrate moieties for Yariv reactivity (Komalavilas et al., 1991). However, with regard to the specific carbohydrate structure required for interaction with β-Yarivs, the results were not always consistent: neither α-l-arabinofuranosyl residues nor β-1,6-galactan side chains were found to be involved in Yariv reactivity of AGPs from Gladiolus spp., radish (Raphanus sativus), and grape (Vitis vinifera; Gleeson and Clarke, 1979; Tsumuraya et al., 1987; Saulnier et al., 1992); partial acid hydrolysis to remove α-l-arabinofuranosyl residues diminished Yariv reactivity of a rose (Rosa spp.) AGP (Komalavilas et al., 1991); and mugwort (Artemisia vulgaris) pollen O-glycans consisting of a β-1,6-galactan core and branched α-l-arabinofuranosyl side chains precipitated with β-Glc-Yariv (Léonard et al., 2005). Accordingly, it has also been suggested that Yariv reactivity depends on the overall physical and chemical properties rather than a specific structural feature of AGPs.In this study, we demonstrate that the peptide component of AGPs is not required for Yariv reactivity. By sequentially trimming the AG moieties of AGPs with sets of specific glycoside hydrolases, we show that β-Gal-Yariv binds to the β-1,3-galactan main chains of radish root AGP. We confirm that β-1,6-galactan side chains are not necessary for Yariv reactivity, we identify β-1,3-galactopentaose (β-1,3-Gal5) as the smallest carbohydrate structure to interact with β-Gal-Yariv, and we show that β-1,3-galactoheptaose (β-1,3-Gal7) or longer β-1,3-galactosyl chains are required for the formation of insoluble precipitate with Yariv phenylglycoside. Based on computational modeling, a possible interaction mechanism between β-Gal-Yariv and β-1,3-galactan is suggested.  相似文献   

12.
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   

13.
Photosynthetic organisms have the ability to adapt to changes in light quality by readjusting the cross sections of the light-harvesting systems of photosystem II (PSII) and photosystem I (PSI). This process, called state transitions, maintains the redox poise of the photosynthetic electron transfer chain and ensures a high photosynthetic yield when light is limiting. It is mediated by the Stt7/STN7 protein kinase, which is activated through the cytochrome b6f complex upon reduction of the plastoquinone pool. Its probable major substrate, the light-harvesting complex of PSII, once phosphorylated, dissociates from PSII and docks to PSI, thereby restoring the balance of absorbed light excitation energy between the two photosystems. Although the kinase is known to be inactivated under high-light intensities, the molecular mechanisms governing its regulation remain unknown. In this study we monitored the redox state of a conserved and essential Cys pair of the Stt7/STN7 kinase and show that it forms a disulfide bridge. We could not detect any change in the redox state of these Cys during state transitions and high-light treatment. It is only after prolonged anaerobiosis that this disulfide bridge is reduced. It is likely to be mainly intramolecular, although kinase activation may involve a transient covalently linked kinase dimer with two intermolecular disulfide bonds. Using the yeast two-hybrid system, we have mapped one interaction site of the kinase on the Rieske protein of the cytochrome b6f complex.Photosynthetic organisms are subjected to constant changes in light quality and quantity and need to adapt to these changes in order to optimize, on the one hand, their photosynthetic yield, and to minimize photo-oxidative damage on the other. The photosynthetic electron transfer chain consists of photosystem II (PSII), the plastoquinone (PQ) pool, the cytochrome b6f complex (Cyt b6f), plastocyanin, and photosystem I (PSI). All of these complexes and components are integrated or closely associated with the thylakoid membrane. The two antenna systems of PSII and PSI capture and direct the light excitation energy to the corresponding reaction centers in which a chlorophyll dimer is oxidized and charge separation occurs across the thylakoid membrane. These processes lead to the onset of electron flow from water on the donor side of PSII to ferredoxin on the acceptor side of PSI coupled with proton translocation across the thylakoid membrane. In order to sustain optimal electron flow along this electron transfer chain, the redox poise needs to be maintained under changing environmental conditions. Several mechanisms have evolved for the maintenance of this redox balance. In the case of over-reduction of the acceptor side of PSI, excess electrons can reduce molecular oxygen through the Mehler reaction to superoxide, which is then converted to hydrogen peroxide by a plastid superoxide dismutase and ultimately to water by a peroxidase (Asada, 2000). Over-reduction of the PQ pool can be alleviated by PTOX, the plastid terminal oxidase responsible for oxidizing PQH2 to form hydrogen peroxide, which is subsequently converted to water (Carol et al., 1999; Cournac et al., 2000; Wu et al., 1999).In addition to these electron sinks that prevent the over-reduction of the electron transfer chain, the photosynthetic apparatus is able to maintain the redox poise of the PQ pool by readjusting the relative cross sections of the light harvesting systems of PSII and PSI upon unequal excitation of the two photosystems. This readjustment can occur both in the short term through state transitions and in the long term by changing the stoichiometry between PSII and PSI (Bonaventura and Myers, 1969; Murata, 1969; Pfannschmidt, 2003). State transitions occur because of perturbations of the redox state of the PQ pool due to unequal excitation of PSII and PSI, limitations in electron acceptors downstream of PSI, and/or in CO2 availability. Excess excitation of PSII relative to PSI leads to reduction of the PQ pool and thus favors the docking of PQH2 to the Qo site of the Cyt b6f complex. This process activates the Stt7/STN7 protein kinase (Vener et al., 1997; Zito et al., 1999), which is closely associated with this complex and leads to the phosphorylation of some LHCII proteins and to their detachment from PSII and binding to PSI (Depège et al., 2003; Lemeille et al., 2009). Although both Lhcb1 and Lhcb2 are phosphorylated, only the phosphorylated form of Lhcb2 is associated with PSI whereas phosphorylated Lhcb1 is excluded from this complex (Longoni et al., 2015). This state corresponds to state 2. In this way the change in the relative antenna sizes of the two photosystems restores the redox poise of the PQ pool. The process is reversible as over-excitation of PSI relative to PSII leads to the oxidation of the PQ pool and to the inactivation of the kinase. Under these conditions, phosphorylated LHCII associated with PSI is dephosphorylated by the PPH1/TAP38 phosphatase (Pribil et al., 2010; Shapiguzov et al., 2010) and returns to PSII (state 1). It should be noted, however, that a strict causal link between LHCII phosphorylation and its migration from PSII to PSI has been questioned recently by the finding that some phosphorylated LHCII remains associated with PSII supercomplexes and that LHCII serves as antenna for both photosystems under most natural light conditions (Drop et al., 2014; Wientjes et al., 2013).State transitions are important at low light but do not occur under high light because the LHCII kinase is inactivated under these conditions (Schuster et al., 1986). It was proposed that inactivation of the kinase is mediated by the ferredoxin-thioredoxin system and that a disulfide bond in the kinase rather than in the substrate may be the target site of thioredoxin (Rintamäki et al., 1997, 2000). Analysis of the Stt7/STN7 protein sequences indeed reveals the presence of two conserved Cys residues close to the N-terminal end of this kinase, which are conserved in all species examined and both are essential for kinase activity although they are located outside of the kinase catalytic domain (Fig. 1) (Depège et al., 2003; Lemeille et al., 2009). Based on protease protection studies, this model of the Stt7/STN7 kinase proposes that the N-terminal end of the kinase is on the lumen side of the thylakoid membrane separated from the catalytic domain on the stromal side by an unusual transmembrane domain containing several Pro residues (Lemeille et al., 2009). This configuration of the kinase allows its catalytic domain to act on the substrate sites of the LHCII proteins, which are exposed to the stroma. Although in this model the conserved Cys residues in the lumen are on the opposite side from the stromal thioredoxins, it is possible that thiol-reducing equivalents are transferred across the thylakoid membrane through the CcdA and Hcf164 proteins, which have been shown to operate in this way during heme and Cyt b6f assembly (Lennartz et al., 2001; Page et al., 2004) or through the LTO1 protein (Du et al., 2015; Karamoko et al., 2011).Figure 1.Conserved Cys in the Stt7/STN7 kinase. Alignment of the sequences of the Stt7/STN protein kinase from Selaginella moelendorffii (Sm), Physcomitrella patens (Pp), Oryza sativa (Os), Populus trichocarpa (Pt), Arabidopsis thaliana (At), Chlamydomonas reinhardtii ...Here we have examined the redox state of the Stt7/STN7 kinase during state transitions and after illumination with high light to test the proposed model. We find that the Stt7/STN7 kinase contains a disulfide bridge that appears to be intramolecular and maintained not only during state transitions but also in high light when the kinase is inactive. Although these results suggest at first sight that the disulfide bridge of Stt7/STN7 is maintained during its activation and inactivation, we propose that a transient opening of this bridge occurs during the activation process followed by the formation of an intermolecular disulfide bridge and the appearance of a short-lived, covalently linked kinase dimer.  相似文献   

14.
Two LHC-like proteins, Photosystem II Subunit S (PSBS) and Light-Harvesting Complex Stress-Related (LHCSR), are essential for triggering excess energy dissipation in chloroplasts of vascular plants and green algae, respectively. The mechanism of quenching was studied in Physcomitrella patens, an early divergent streptophyta (including green algae and land plants) in which both proteins are active. PSBS was localized in grana together with photosystem II (PSII), but LHCSR was located mainly in stroma-exposed membranes together with photosystem I (PSI), and its distribution did not change upon high-light treatment. The quenched conformation can be preserved by rapidly freezing the high-light-treated tissues in liquid nitrogen. When using green fluorescent protein as an internal standard, 77K fluorescence emission spectra on isolated chloroplasts allowed for independent assessment of PSI and PSII fluorescence yield. Results showed that both photosystems underwent quenching upon high-light treatment in the wild type in contrast to mutants depleted of LHCSR, which lacked PSI quenching. Due to the contribution of LHCII, P. patens had a PSI antenna size twice as large with respect to higher plants. Thus, LHCII, which is highly abundant in stroma membranes, appears to be the target of quenching by LHCSR.  相似文献   

15.
The photoinhibition of photosystem I (PSI) is lethal to oxygenic phototrophs. Nevertheless, it is unclear how photodamage occurs or how oxygenic phototrophs prevent it. Here, we provide evidence that keeping P700 (the reaction center chlorophyll in PSI) oxidized protects PSI. Previous studies have suggested that PSI photoinhibition does not occur in the two model cyanobacteria, Synechocystis sp. PCC 6803 and Synechococcus elongatus PCC 7942, when photosynthetic CO2 fixation was suppressed under low CO2 partial pressure even in mutants deficient in flavodiiron protein (FLV), which mediates alternative electron flow. The lack of FLV in Synechococcus sp. PCC 7002 (S. 7002), however, is linked directly to reduced growth and PSI photodamage under CO2-limiting conditions. Unlike Synechocystis sp. PCC 6803 and S. elongatus PCC 7942, S. 7002 reduced P700 during CO2-limited illumination in the absence of FLV, resulting in decreases in both PSI and photosynthetic activities. Even at normal air CO2 concentration, the growth of S. 7002 mutant was retarded relative to that of the wild type. Therefore, P700 oxidation is essential for protecting PSI against photoinhibition. Here, we present various strategies to alleviate PSI photoinhibition in cyanobacteria.Low CO2 fixation efficiency in the Calvin-Benson cycle prevents the utilization of NADPH and ATP in photosynthesis and causes these molecules to accumulate, resulting in oxidative photosynthetic cell damage. High light, low temperature, and CO2 limitation increase NADPH and ATP levels beyond the Calvin-Benson cycle requirements. Electrons and H+ accumulate in the photosynthetic electron transport (PET) system. Excess electrons in the PET system trigger oxidative damage to PSI by forming reactive oxygen species (ROS), including the superoxide anion radical (O2) and singlet oxygen (1O2), within PSI and degrading the P700 reaction center chlorophyll (P700; Sonoike, 1996; Sejima et al., 2014; Zivcak et al., 2015a, 2015b; Takagi et al., 2016). PSI repair has been reported to be a very slow process (Kudoh and Sonoike, 2002), and a recent study showed that it took more than 12 d for damaged PSI in wheat (Triticum aestivum) leaves to recover completely (Zivcak et al., 2015b). PSI photoinhibition, therefore, is very detrimental to oxygenic phototroph growth. Nevertheless, PSI photoinhibition is alleviated by keeping P700 oxidized (Sejima et al., 2014).In the PET system of oxygenic phototrophs, P700 oxidation is a physiological response to environmental variations. In C3 plants, low CO2 and/or high light intensity induce P700 oxidation in vivo (Klughammer and Schreiber, 1994; Laisk and Oja, 1994; Miyake et al., 2004, 2005). Several molecular mechanisms are proposed for P700 oxidation wherein the PSI acceptor does not limit the PET reaction. First, H+ accumulation on the lumenal side of thylakoid membranes lowers reduced plastoquinone (plastoquinol) oxidation rates in the cytochrome (Cyt) b6/f complex (Kramer et al., 1999). Second, plastidial terminal oxidase and cyanobacterial respiratory terminal oxidases on the thylakoid membranes suppress PSI electron influx by accepting upstream PSI electrons in the PET system. Oxygen is the final electron acceptor (Beardall et al., 2003; Trouillard et al., 2012; Lea-Smith et al., 2013). Finally, plastoquinol accumulation inhibits the Q-cycle turnover in the Cyt b6/f complex, which suppresses electron flow from the Cyt b6/f complex to P700. This reaction is called the reduction-induced suppression of electron flow (RISE; Shaku et al., 2016). Overall, these molecular mechanisms contribute to P700 oxidation, thereby preventing PSI photoinhibition and enabling oxygenic phototrophs to thrive. The proton gradient regulation5 (pgr5) mutant of Arabidopsis (Arabidopsis thaliana) cannot keep P700 oxidized and shows PSI photoinhibition under high-light and fluctuating light conditions (Munekage et al., 2002; Suorsa et al., 2012), which shows the importance of the oxidation of P700 for the protection of PSI in plants.Unlike green plants, P700 oxidation mechanisms in cyanobacteria are unclear. It is known that flavodiiron protein (FLV) could contribute to P700 oxidation. Four FLV isozymes (FLV1−FLV4) have been identified in the model cyanobacterium Synechocystis sp. PCC 6803 (S. 6803; Helman et al., 2003). FLV1 and FLV3 (FLV1/3) function as a heterodimer and catalyze the reduction of oxygen to water on the acceptor side of PSI using NAD(P)H as electron donors (Vicente et al., 2002; Helman et al., 2003; Allahverdiyeva et al., 2013). Unlike FLV1/3, FLV2/4 is induced only under low CO2 (Zhang et al., 2009) and mediates an oxygen-dependent alternative electron flow (AEF; Shimakawa et al., 2015). In S. 6803, FLV-dependent electron fluxes are coupled to photosynthesis and should alleviate electron overaccumulation in PSI (Helman et al., 2003, 2005; Allahverdiyeva et al., 2013; Shimakawa et al., 2015). Therefore, FLV is expected to contribute to P700 oxidation. The lack of FLV1/3 in S. 6803 causes PSI photoinhibition under artificial fluctuating light (Allahverdiyeva et al., 2013). However, under CO2 limitation (which suppresses photosynthetic CO2 fixation), deletions of FLV1/3 and FLV2/4 do not cause PSI photoinhibition in S. 6803 (Zhang et al., 2009) or Synechococcus elongatus PCC 7942 (S. 7942; Shaku et al., 2015), possibly because P700 stays oxidized under CO2 limitation regardless of the existence of FLV (Shaku et al., 2015). These data imply that FLV is not essential to keep P700 oxidized under CO2 limitation, at least in S. 6803 and S. 7942.In this study, we found that the lack of FLV1/3 leads to growth inhibition under ambient [CO2] concentration ([CO2]) in the cyanobacterium Synechococcus sp. PCC 7002 (S. 7002), unlike S. 6803 and S. 7942 (Zhang et al., 2009; Shaku et al., 2015). The S. 7002 genome, like that of S. 7942, includes genes coding for FLV1/3 isozymes but not for FLV2/4 (Fujisawa et al., 2014). The genetic profiles of flv and other genes related to cyanobacterial AEF, including those of S. 7002, S. 6803, and S. 7942, are summarized in P700 to approximately 10% in the flv knockout mutant of S. 7002 but not in those of S. 6803 or S. 7942. We demonstrated that the deletion of FLV in S. 7002 rendered it unable to oxidize P700, resulting in PSI photoinhibition. These findings show that there are different strategies in cyanobacteria to protect PSI against photooxidative damage under CO2 limitation.

Table I.

Genetic background of AEF in three cyanobacteria species used in this studyGene homology analyses were performed in Cyanobase (http://genome.microbedb.jp/CyanoBase; Fujisawa et al., 2014). cox, aa3-type cytochrome c oxidase; cyd, cytochrome bd-type quinol oxidase; arto, cytochrome bo-type quinol oxidase; ndhD, D subunit of NAD(P)H dehydrogenase.
CyanobacteriaFlavodiiron Proteins
Respiratory Terminal Oxidases
Cyclic Electron Flow
flv1/3flv2/4coxcydartondhD1/2pgr5
S. 7002 (SYNPCC7002_)A1743/A1321A1162A1164A0725A0727A2000/A1973A1477
S. 6803sll1521/sll0550sll0219/sll0217slr1136slr1138slr1379/slr1380slr2082/slr2083/sll0813slr0331/slr1291ssr2016
S. 7942 (SYNPCC7942_)1810/1809260226041766/17671976/1439
Open in a separate window  相似文献   

16.
17.
Cyanobacteria have developed a photoprotective mechanism that decreases the energy arriving at the reaction centers by increasing thermal energy dissipation at the level of the phycobilisome (PB), the extramembranous light-harvesting antenna. This mechanism is triggered by the photoactive Orange Carotenoid Protein (OCP), which acts both as the photosensor and the energy quencher. The OCP binds the core of the PB. The structure of this core differs in diverse cyanobacterial strains. Here, using two isolated OCPs and four classes of PBs, we demonstrated that differences exist between OCPs related to PB binding, photoactivity, and carotenoid binding. Synechocystis PCC 6803 (hereafter Synechocystis) OCP, but not Arthrospira platensis PCC 7345 (hereafter Arthrospira) OCP, can attach echinenone in addition to hydroxyechinenone. Arthrospira OCP binds more strongly than Synechocystis OCP to all types of PBs. Synechocystis OCP can strongly bind only its own PB in 0.8 m potassium phosphate. However, if the Synechocystis OCP binds to the PB at very high phosphate concentrations (approximately 1.4 m), it is able to quench the fluorescence of any type of PB, even those isolated from strains that lack the OCP-mediated photoprotective mechanism. Thus, the determining step for the induction of photoprotection is the binding of the OCP to PBs. Our results also indicated that the structure of PBs, at least in vitro, significantly influences OCP binding and the stabilization of OCP-PB complexes. Finally, the fact that the OCP induced large fluorescence quenching even in the two-cylinder core of Synechococcus elongatus PBs strongly suggested that OCP binds to one of the basal allophycocyanin cylinders.The cyanobacterial Orange Carotenoid Protein (OCP) is a photoactive soluble protein of 35 kD that binds a ketocarotenoid, 3′-hydroxyechinenone (hECN). It is present in the majority of phycobilisome (PB)-containing cyanobacterial strains (Kirilovsky and Kerfeld, 2012, 2013). The PBs are light-harvesting extramembrane complexes formed by a core from which rods radiate. The core and rods are constituted of water-soluble blue and red phycobiliproteins, which covalently attach bilins (for review, see Glazer, 1984; Grossman et al., 1993; MacColl, 1998; Tandeau de Marsac, 2003; Adir, 2005). The OCP was first described by Holt and Krogmann (1981), and its structure was determined in 2003 (Kerfeld et al., 2003). However, its function was discovered only in 2006 (Wilson et al., 2006) and its photoactivity in 2008 (Wilson et al., 2008). The OCP is essential in a photoprotective mechanism that decreases the energy arriving at the reaction centers under high irradiance. Strong light induces thermal dissipation of the energy absorbed by the PBs, resulting in a decrease of PB fluorescence emission and of energy transfer from the PBs to the reaction centers (Wilson et al., 2006). This process, which is light intensity dependent, is induced by blue or green light but not by orange or red light (Rakhimberdieva et al., 2004; Wilson et al., 2006). The absorption of strong blue-green light by the OCP induces changes in the conformation of the carotenoid, converting the inactive orange dark form (OCPo) into an active red form (OCPr; Wilson et al., 2008). In OCPo, the hECN is in an all-trans-configuration (Kerfeld et al., 2003; Polívka et al., 2005). In OCPr, the apparent conjugation length of the carotenoid increases, resulting in a less distorted, more planar structure (Wilson et al., 2008). Fourier transform infrared spectra showed that conformational changes in the protein are also induced (Wilson et al., 2008) that are essential for the induction of the photoprotective mechanism. Only OCPr is able to bind to the core of PBs and to induce thermal energy dissipation (Wilson et al., 2008; Punginelli et al., 2009; Gorbunov et al., 2011; Gwizdala et al., 2011). Since the photoactivation of the OCP has a very low quantum yield (0.03; Wilson et al., 2008), the concentration of activated protein is zero in darkness and very low under low-light conditions (Wilson et al., 2008; Gorbunov et al., 2011). Thus, the photoprotective mechanism functions only under high-light conditions.The crystal structures of the Arthrospira maxima OCP and of the Synechocystis PCC 6803 (hereafter Synechocystis) OCP were solved in 2003 and 2010, respectively (Kerfeld et al., 2003; Wilson et al., 2010). These structures, assumed to correspond to the dark OCPo form, are essentially identical. The OCP consists of an all-α-helical N-terminal domain (residues 1–165), unique to cyanobacteria, and an α-helical/β-sheet C-terminal domain that is a member of the Nuclear Transport Factor2 superfamily (residues 191–320; Synechocystis numbering). Both domains are joined by a linker (residues 166–190; Synechocystis numbering) that appears to be flexible. The hECN molecule spans the N- and C-terminal domains of the protein, with its carbonyl end embedded in and hydrogen bonded to two absolutely conserved residues (Tyr-201 and Trp-288) in the C-terminal domain. The carotenoid is almost entirely buried; only 3.4% of the 3′ hECN is solvent exposed (Kerfeld et al., 2003). Synechocystis OCP can also bind with high-affinity echinenone (ECN) and zeaxanthin. While the ECN OCP is photoactive, the zeaxanthin OCP is photoinactive (Punginelli et al., 2009), indicating the importance of the carotenoid carbonyl group for photoactivity. The largest interface through which the two domains interact and through which the carotenoid passes is stabilized by a small number of hydrogen bonds, including one formed between Arg-155 and Glu-244 (Wilson et al., 2010). This salt bridge stabilizes the closed structure of OCPo. Upon illumination, protein conformational changes cause the breakage of this bond and the opening of the protein (Wilson et al., 2012). Arg-155, which becomes more exposed upon the separation of the two domains, is essential for the OCPr binding to the PBs (Wilson et al., 2012).After exposure to high irradiance, when the light intensity decreases, recovery of full antenna capacity and fluorescence requires another protein, the Fluorescence Recovery Protein (FRP; Boulay et al., 2010). The active form of this soluble 13-kD protein is a dimer (Sutter et al., 2013). It interacts with the OCPr C-terminal domain (Boulay et al., 2010; Sutter et al., 2013). This accelerates the red-to-orange OCP conversion and helps the OCP to detach from the PB (Boulay et al., 2010; Gwizdala et al., 2011).Genes encoding the full-length OCP are found in the vast majority of cyanobacteria but not in all; 90 of 127 genomes recently surveyed contain at least one gene for a full-length OCP (Kirilovsky and Kerfeld, 2013). The genomes of Synechococcus elongatus and Thermosynechococcus elongatus, two cyanobacterial strains used as model organisms in photosynthesis and stress studies, do not contain a full-length ocp gene. These strains also lack FRP and β-carotene ketolase (involved in ketocarotenoid synthesis). As a consequence, these strains lack the OCP-related photoprotective mechanism and are more sensitive to fluctuating light intensities (Boulay et al., 2008).The core of the hemidiscoidal PBs of Synechocystis, the model organism routinely used for the study of the OCP-related photoprotective mechanism, consists of three cylinders, each one formed by four trimers of allophycocyanin (APC; Fig. 1; for review, see Glazer, 1984; Bryant, 1991; Grossman et al., 1993; MacColl, 1998; Adir, 2005). The APC trimers are predominantly assembled from a two-subunit heterodimer, αAPC-βAPC, which binds two phycocyanobilins, one in each subunit. Of the 12 total APC trimers in the PB core, eight are trimers of αAPC-βAPC. These trimers have a maximal emission at 660 nm (APC660). The upper cylinder contains only APC660 trimers. In contrast, each basal cylinder contains only two APC660 trimers. Each basal cylinder also contains the following: (1) a trimer in which one αAPC subunit is replaced by a special αAPC-like subunit called ApcD, and (2) a trimer in which one β-subunit is replaced by ApcF, a βAPC-like subunit, and one α-subunit is replaced by the N-terminal domain of ApcE, an αAPC-like domain (Fig. 1). The trimers containing one or two of these special subunits have a maximal emission at 680 nm (APC680). In each cylinder, the two external trimers are stabilized by an 8.7-kD linker protein.Open in a separate windowFigure 1.Schematic orthogonal projections of the various PB cores. In the PBs containing three or five cylinders, the top complete cylinder is formed by four αAPC-βAPC trimers emitting at 660 nm. Each of the basal cylinders of three types of PBs contains two αAPC-βAPC trimers emitting at 660 nm and two trimers emitting at 683 nm. In one of them, one αAPC is replaced by ApcD, and in the other one, αAPC-βAPC is replaced by the dimer ApcF-ApcE. In the five cylinder PBs, two additional semicylinders formed by two αAPC-βAPC trimers are present. In all the cylinders, the two external trimers include an 8.7-kD linker protein (ApcC).The C-terminal part of Synechocystis ApcE contains three interconnected repeated domains of about 120 residues (called Rep domains) that are similar to the conserved domains of rod linkers. Each Rep domain interacts with an APC trimer situated in different cylinders, which stabilizes the core of PB (Zhao et al., 1992; Shen et al., 1993; Ajlani et al., 1995; Ajlani and Vernotte, 1998). The ApcE protein also determines the number of APC cylinders that form the PB core (Capuano et al., 1991, 1993). Indeed, there are PBs containing only the two basal cylinders, as in S. elongatus (ex S. elongatus PCC 7942) and Synechococcus PCC 6301. In these strains, the approximately 72-kD ApcE possesses only two Rep domains. There also exist pentacylindrical cores in which, in addition to the three cylinders existing in Synechocystis PBs, there are two other cylinders, each formed by two APC660 trimers, for example in Anabaena variabilis, Anabaena PCC 7120, and Mastigocladus laminosus (Glauser et al., 1992; Ducret et al., 1998). In the pentacylindrical PBs, ApcE (approximately 125 kD) contains four Rep domains (Capuano et al., 1993). Finally, ApcE is also involved in the interaction between the PB and the thylakoids.The bicylindric and tricylindric cores are surrounded by six rods formed generally by three hexamers of the blue phycocyanin (PC) or two PC hexamers and one hexamer containing phycoerythrin or phycoerythrincyanonin. The rods and the hexamers are stabilized by nonchromophorylated linker proteins. A linker protein, LRC also stabilizes the binding of the rods to the core. The pentacylindric PBs can contain up to eight rods. The quantity and length of rods and the presence of phycoerythrin or phycoerythrocyanin at the periphery of the rods depends on environmental conditions like light intensity or quality (Kipe-Nolt et al., 1982; Glauser et al., 1992).The OCP probably binds to one of the APC660 trimers (Tian et al., 2011, 2012; Jallet et al., 2012), and the presence of the rods stabilizes this binding to Synechocystis PBs (Gwizdala et al., 2011). The different structures of PBs in other strains could affect the binding of the OCP. Thus, we undertook a study about the relationship between the structure of PBs and OCP binding in preparation for introducing the OCP-related photoprotective mechanism into S. elongatus using Synechocystis genes. In this study, we used the in vitro reconstitution system developed by Gwizdala et al. (2011) with three different types of isolated PBs: Arthrospira platensis PCC 7345 (hereafter Arthrospira) PBs, having a tricylindrical core like Synechocystis PBs; Anabaena variabilis (hereafter Anabaena) PBs, having a pentacylindrical core; and S. elongatus PCC 7942 (hereafter Synechococcus) PBs, having a bicylindrical core. We also used two different OCPs, the Synechocystis OCP and the Arthrospira OCP. Each OCP was isolated from mutant Synechocystis cells overexpressing one or the other ocp gene with a C-terminal His tag.  相似文献   

18.
19.
Divinyl reductase (DVR) converts 8-vinyl groups on various chlorophyll intermediates to ethyl groups, which is indispensable for chlorophyll biosynthesis. To date, five DVR activities have been detected, but adequate evidence of enzymatic assays using purified or recombinant DVR proteins has not been demonstrated, and it is unclear whether one or multiple enzymes catalyze these activities. In this study, we systematically carried out enzymatic assays using four recombinant DVR proteins and five divinyl substrates and then investigated the in vivo accumulation of various chlorophyll intermediates in rice (Oryza sativa), maize (Zea mays), and cucumber (Cucumis sativus). The results demonstrated that both rice and maize DVR proteins can convert all of the five divinyl substrates to corresponding monovinyl compounds, while both cucumber and Arabidopsis (Arabidopsis thaliana) DVR proteins can convert three of them. Meanwhile, the OsDVR (Os03g22780)-inactivated 824ys mutant of rice exclusively accumulated divinyl chlorophylls in its various organs during different developmental stages. Collectively, we conclude that a single DVR with broad substrate specificity is responsible for reducing the 8-vinyl groups of various chlorophyll intermediates in higher plants, but DVR proteins from different species have diverse and differing substrate preferences, although they are homologous.Chlorophyll (Chl) molecules universally exist in photosynthetic organisms. As the main component of the photosynthetic pigments, Chl molecules perform essential processes of absorbing light and transferring the light energy in the reaction center of the photosystems (Fromme et al., 2003). Based on the number of vinyl side chains, Chls are classified into two groups, 3,8-divinyl (DV)-Chl and 3-monovinyl (MV)-Chl. The DV-Chl molecule contains two vinyl groups at positions 3 and 8 of the tetrapyrrole macrocycle, whereas the MV-Chl molecule contains a vinyl group at position 3 and an ethyl group at position 8 of the macrocycle. Almost all of the oxygenic photosynthetic organisms contain MV-Chls, with the exceptions of some marine picophytoplankton species that contain only DV-Chls as their primary photosynthetic pigments (Chisholm et al., 1992; Goericke and Repeta, 1992; Porra, 1997).The classical single-branched Chl biosynthetic pathway proposed by Granick (1950) and modified by Jones (1963) assumed the rapid reduction of the 8-vinyl group of DV-protochlorophyllide (Pchlide) catalyzed by a putative 8-vinyl reductase. Ellsworth and Aronoff (1969) found evidence for both MV and DV forms of several Chl biosynthetic intermediates between magnesium-protoporphyrin IX monomethyl ester (MPE) and Pchlide in Chlorella spp. mutants. Belanger and Rebeiz (1979, 1980) reported that the Pchlide pool of etiolated higher plants contains both MV- and DV-Pchlide. Afterward, following the further detection of MV- and DV-tetrapyrrole intermediates and their biosynthetic interconversion in tissues and extracts of different plants (Belanger and Rebeiz, 1982; Duggan and Rebeiz, 1982; Tripathy and Rebeiz, 1986, 1988; Parham and Rebeiz, 1992, 1995; Kim and Rebeiz, 1996), a multibranched Chl biosynthetic heterogeneity was proposed (Rebeiz et al., 1983, 1986, 1999; Whyte and Griffiths, 1993; Kolossov and Rebeiz, 2010).Biosynthetic heterogeneity refers to the biosynthesis of a particular metabolite by an organelle, tissue, or organism via multiple biosynthetic routes. Varieties of reports lead to the assumption that Chl biosynthetic heterogeneity originates mainly in parallel DV- and MV-Chl biosynthetic routes. These routes are interconnected by 8-vinyl reductases that convert DV-tetrapyrroles to MV-tetrapyrroles by conversion of the vinyl group at position 8 of ring B to the ethyl group (Parham and Rebeiz, 1995; Rebeiz et al., 2003). DV-MPE could be converted to MV-MPE in crude homogenates from etiolated wheat (Triticum aestivum) seedlings (Ellsworth and Hsing, 1974). Exogenous DV-Pchlide could be partially converted to MV-Pchlide in barley (Hordeum vulgare) plastids (Tripathy and Rebeiz, 1988). 8-Vinyl chlorophyllide (Chlide) a reductases in etioplast membranes isolated from etiolated cucumber (Cucumis sativus) cotyledons and barley and maize (Zea mays) leaves were found to be very active in the conversion of exogenous DV-Chlide a to MV-Chlide a (Parham and Rebeiz, 1992, 1995). Kim and Rebeiz (1996) suggested that Chl biosynthetic heterogeneity in higher plants may originate at the level of DV magnesium-protoporphyrin IX (Mg-Proto) and would be mediated by the activity of a putative 8-vinyl Mg-Proto reductase in barley etiochloroplasts and plastid membranes. However, since these reports did not use purified or recombinant enzyme, it is not clear whether the reductions of the 8-vinyl groups of various Chl intermediates are catalyzed by one enzyme of broad specificity or by multiple enzymes of narrow specificity, which actually has become one of the focus issues in Chl biosynthesis.Nagata et al. (2005) and Nakanishi et al. (2005) independently identified the AT5G18660 gene of Arabidopsis (Arabidopsis thaliana) as an 8-vinyl reductase, namely, divinyl reductase (DVR). Chew and Bryant (2007) identified the DVR BciA (CT1063) gene of the green sulfur bacterium Chlorobium tepidum, which is homologous to AT5G18660. An enzymatic assay using a recombinant Arabidopsis DVR (AtDVR) on five DV substrates revealed that the major substrate of AtDVR is DV-Chlide a, while the other four DV substrates could not be converted to corresponding MV compounds (Nagata et al., 2007). Nevertheless, a recombinant BciA is able to reduce the 8-vinyl group of DV-Pchlide to generate MV-Pchlide (Chew and Bryant, 2007). Recently, we identified the rice (Oryza sativa) DVR encoded by Os03g22780 that has sequence similarity with the Arabidopsis DVR gene AT5G18660. We also confirmed that the recombinant rice DVR (OsDVR) is able to not only convert DV-Chlide a to MV-Chlide a but also to convert DV-Chl a to MV-Chl a (Wang et al., 2010). Thus, it is possible that the reductions of the 8-vinyl groups of various Chl biosynthetic intermediates are catalyzed by one enzyme of broad specificity.In this report, we extended our studies to four DVR proteins and five DV substrates. First, ZmDVR and CsDVR genes were isolated from maize and cucumber genomes, respectively, using a homology-based cloning approach. Second, enzymatic assays were systematically carried out using recombinant OsDVR, ZmDVR, CsDVR, and AtDVR as representative DVR proteins and using DV-Chl a, DV-Chlide a, DV-Pchlide a, DV-MPE, and DV-Mg-Proto as DV substrates. Third, we examined the in vivo accumulations of various Chl intermediates in rice, maize, and cucumber. Finally, we systematically investigated the in vivo accumulations of Chl and its various intermediates in the OsDVR (Os03g22780)-inactivated 824ys mutant of rice (Wang et al., 2010). The results strongly suggested that a single DVR protein with broad substrate specificity is responsible for reducing the 8-vinyl groups of various intermediate molecules of Chl biosynthesis in higher plants, but DVR proteins from different species could have diverse and differing substrate preferences even though they are homologous.  相似文献   

20.
Brassinosteroid (BR) and gibberellin (GA) are two predominant hormones regulating plant cell elongation. A defect in either of these leads to reduced plant growth and dwarfism. However, their relationship remains unknown in rice (Oryza sativa). Here, we demonstrated that BR regulates cell elongation by modulating GA metabolism in rice. Under physiological conditions, BR promotes GA accumulation by regulating the expression of GA metabolic genes to stimulate cell elongation. BR greatly induces the expression of D18/GA3ox-2, one of the GA biosynthetic genes, leading to increased GA1 levels, the bioactive GA in rice seedlings. Consequently, both d18 and loss-of-function GA-signaling mutants have decreased BR sensitivity. When excessive active BR is applied, the hormone mostly induces GA inactivation through upregulation of the GA inactivation gene GA2ox-3 and also represses BR biosynthesis, resulting in decreased hormone levels and growth inhibition. As a feedback mechanism, GA extensively inhibits BR biosynthesis and the BR response. GA treatment decreases the enlarged leaf angles in plants with enhanced BR biosynthesis or signaling. Our results revealed a previously unknown mechanism underlying BR and GA crosstalk depending on tissues and hormone levels, which greatly advances our understanding of hormone actions in crop plants and appears much different from that in Arabidopsis thaliana.  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