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.
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   

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

4.
5.
6.
Metals exert important functions in the chloroplast of plants, where they act as cofactors and catalysts in the photosynthetic electron transport chain. In particular, manganese (Mn) has a key function because of its indispensable role in the water-splitting reaction of photosystem II (PSII). More and better knowledge is required on how the various complexes of PSII are affected in response to, for example, nutritional disorders and other environmental stress conditions. We here present, to our knowledge, a new method that allows the analysis of metal binding in intact photosynthetic complexes of barley (Hordeum vulgare) thylakoids. The method is based on size exclusion chromatography coupled to inductively coupled plasma triple-quadrupole mass spectrometry. Proper fractionation of PSII super- and subcomplexes was achieved by critical selection of elution buffers, detergents for protein solubilization, and stabilizers to maintain complex integrity. The applicability of the method was shown by quantification of Mn binding in PSII from thylakoids of two barley genotypes with contrasting Mn efficiency exposed to increasing levels of Mn deficiency. The amount of PSII supercomplexes was drastically reduced in response to Mn deficiency. The Mn efficient genotype bound significantly more Mn per unit of PSII under control and mild Mn deficiency conditions than the inefficient genotype, despite having lower or similar total leaf Mn concentrations. It is concluded that the new method facilitates studies of the internal use of Mn and other biometals in various PSII complexes as well as their relative dynamics according to changes in environmental conditions.Several metals are important for chloroplast functioning, particularly in the photosynthetic apparatus, where they act as cofactors and catalysts in electron transport processes (Merchant, 2006; Nouet et al., 2011; Yruela, 2013). The photosynthetic biometals include iron (Fe) in the form of Fe-S clusters in PSI, heme-bridged Fe (cytochrome b559) and nonheme Fe in PSII, copper (Cu) in plastocyanin, magnesium (Mg) in chlorophyll (Chl), and calcium (Ca) and manganese (Mn) in PSII. Mn has a very special role because a metal cluster of four Mn ions and one Ca ion comprises the catalytic center of the oxygen evolving complex (OEC) in PSII (Ono et al., 1992; Umena et al., 2011). In the OEC, water is split, and molecular oxygen is produced by the photosynthetic light reactions. The photosynthetic biometals are, however, highly reactive and involved in a multitude of side reactions, which constitute a challenge for metal homeostasis. Accordingly, the handling of metals must be tightly regulated, and they must be kept within specific concentration ranges inside living cells to ensure adequate supply, while at the same time, avoiding oxidative stress (Pakrasi et al., 2001; Shcolnick and Keren, 2006; Møller et al., 2007).PSII is a large pigment-protein complex localized in the grana regions of the thylakoid membrane of chloroplasts. The basic structure of PSII is a monomer, and each complex contains more than 40 different proteins bound either stably or transiently (Nelson and Yocum, 2006; Shi et al., 2012; Järvi et al., 2015). The luminal surfaces of PSII are associated with the extrinsic proteins PsbO, PsbP, and PsbQ, which shield and support the catalytic Mn cluster and are required for efficient oxygen evolution (Roose et al., 2007; Bricker et al., 2012; Liu et al., 2014). After dimerization of the monomer, the complex associates with multiple copies of the light-harvesting antenna complex II (LHCII), forming various types of functional PSII-LHCII supercomplexes (Tikkanen et al., 2008; Kouřil et al., 2012; Shi et al., 2012).Intact PSII-LHCII supercomplexes have been successfully isolated, characterized, and refined from, for example, pea (Pisum sativum; Barera et al., 2012), Arabidopsis (Arabidopsis thaliana; Caffarri et al., 2009), and green algae (Chlamydomonas reinhardtii; Tokutsu et al., 2012). The procedure has typically involved Suc density gradient ultracentrifugation. Also, blue native (BN)-PAGE has been optimized for the separation and proteomic characterization of thylakoid PSII-LHCII supercomplexes (Heinemeyer et al., 2004; Järvi et al., 2011; Pagliano et al., 2014). The supramolecular organization of isolated PSII is very much dependent on the choice of detergent for efficient solubilization of the membrane-bound photosynthetic pigment-protein complexes. In recent years, dodecyl maltoside (DM) has become a commonly used detergent for one-step isolation of integral membrane proteins and complexes from thylakoids (Eshaghi et al., 1999; van Roon et al., 2000; Dekker et al., 2002; Pagliano et al., 2011). This detergent exists in two isomeric forms (α-DM and β-DM), of which α-DM is a milder detergent than β-DM, thereby better preserving the integrity of large PSII-LHCII supercomplexes (Pagliano et al., 2012).The major challenge associated with purification of higher plant PSII-LHCII supercomplexes is to obtain and subsequently, maintain the integrity of PSII super- and subcomplexes, including cofactors and the extrinsic proteins. To prevent dissociation of biometals and the extrinsic proteins from PSII, the osmoprotectant betaine (Papageorgiou et al., 1991; Papageorgiou and Murata, 1995) has successfully been included in the buffer of Suc gradients (Boekema et al., 1998; Tokutsu et al., 2012). Although the above-mentioned methods primarily have focused on the characterization and structural organization of isolated PSII-LHCII supercomplexes, no bench-top method has been available that allows direct analysis of the actual metal binding in PSII super- and subcomplexes. Such a method is required in order to fully understand how Mn and other photosynthetic biometals interact with the photosynthetic complexes, in particular PSII, and how the metal binding affects PSII dynamics under changing environmental conditions, including plant nutritional disorders.We here present a robust and highly sensitive method for analysis of metal binding in PSII-LHCII super- and subcomplexes from isolated barley (Hordeum vulgare) thylakoids. The method is based on size exclusion chromatography (SEC) coupled to inductively coupled plasma (ICP) triple-quadrupole (QQQ) mass spectrometry (MS). SEC is a gentle protein separation technique, provided that the stationary and mobile phases are carefully selected. Using an optimized set of analytical conditions, it is possible to maintain the integrity of metalloprotein complexes (Persson et al., 2009; Husted et al., 2011). We systematically evaluate the essential and important factors required to obtain optimal chromatographic resolution while maintaining PSII integrity, focusing on choice of mobile phase, detergents, stabilizers, and the most suitable chromatographic columns for efficient protein fractionation and elution. The optimized method, with its multielement ability, enables the study of metal binding in PSII-LHCII super- and subcomplexes. To show the applicability of the method, we studied the metal profiles of barley thylakoids that had been isolated from plants with different levels of Mn deficiency. Mn binding in size-fractionated PSII complexes was evaluated in response to increasing Mn deficiency, and two genotypes differing in their tolerance to Mn deficiency were compared.  相似文献   

7.
8.
9.
10.
11.
12.
13.
14.
Oxygenic photosynthesis evolved with cyanobacteria, the ancestors of plant chloroplasts. The highly oxidizing chemistry of water splitting required concomitant evolution of efficient photoprotection mechanisms to safeguard the photosynthetic machinery. The role of flavodiiron proteins (FDPs), originally called A-type flavoproteins or Flvs, in this context has only recently been appreciated. Cyanobacterial FDPs constitute a specific protein group that evolved to protect oxygenic photosynthesis. There are four FDPs in Synechocystis sp. PCC 6803 (Flv1 to Flv4). Two of them, Flv2 and Flv4, are encoded by an operon together with a Sll0218 protein. Their expression, tightly regulated by CO2 levels, is also influenced by changes in light intensity. Here we describe the overexpression of the flv4-2 operon in Synechocystis sp. PCC 6803 and demonstrate that it results in improved photochemistry of PSII. The flv4-2/OE mutant is more resistant to photoinhibition of PSII and exhibits a more oxidized state of the plastoquinone pool and reduced production of singlet oxygen compared with control strains. Results of biophysical measurements indicate that the flv4-2 operon functions in an alternative electron transfer pathway from PSII, and thus alleviates PSII excitation pressure by channeling up to 30% of PSII-originated electrons. Furthermore, intact phycobilisomes are required for stable expression of the flv4-2 operon genes and for the Flv2/Flv4 heterodimer-mediated electron transfer mechanism. The latter operates in photoprotection in a complementary way with the orange carotenoid protein-related nonphotochemical quenching. Expression of the flv4-2 operon and exchange of the D1 forms in PSII centers upon light stress, on the contrary, are mutually exclusive photoprotection strategies among cyanobacteria.Photosynthetic light reactions are evolutionarily highly conserved among oxygenic photosynthetic organisms from cyanobacteria to higher plants. Because of dangerous chemistry of the water splitting reactions, oxygenic photosynthesis produces reactive oxygen species (ROS) and other radicals that potentially could destroy the photosynthetic machinery. To avoid permanent damage, all oxygenic photosynthetic organisms are equipped with an array of various photoprotective and regulatory mechanisms. Accumulating evidence on these regulatory mechanisms has revealed vast evolutionary differences between organisms performing oxygenic photosynthesis.Photosynthetic organisms have a capacity to adjust to different light intensities and to changes in the availability of electron sinks, which depends largely on metabolic cues. When light or metabolic conditions change, photosystems can dissipate excess energy as heat in nonphotochemical energy dissipation processes in the light-harvesting antenna systems (for review, see Horton et al., 1996; Müller et al., 2001). Cyanobacteria have phycobilisomes (PBs) as light-harvesting antenna, which also participate in state transitions (for review, see van Thor et al., 1998; Mullineaux and Emlyn-Jones, 2005) and nonphotochemical quenching (NPQ) of excitation energy (for review, see Bailey and Grossman, 2008). Both of these processes are involved in short-term regulation of light-harvesting processes and concomitantly function as photoprotective mechanisms. These nonphotochemical energy quenching mechanisms, however, have only limited capacity, and it often occurs that more electrons are excited than can be safely used in photochemistry for reduction of natural metabolic electron acceptors, particularly under stress conditions. In such situations, electrons can be directed, for example, to molecular oxygen resulting in production of ROS. To avoid harmful reactions by ROS that threaten the cell viability, a repertoire of different photoprotection mechanisms have evolved in cyanobacteria as well as in all other oxygenic photosynthetic organisms.Flavodiiron proteins (FDPs), originally called A-type flavoproteins or Flvs (Wasserfallen et al., 1998), were recently demonstrated to have an important role in photoprotection of the photosynthetic machinery (Zhang et al., 2009, 2012; Allahverdiyeva et al., 2011, 2013; Ermakova et al., 2013). FDPs in general are most widespread among strict and facultative anaerobic bacteria. Many of their FDPs have been characterized structurally and functionally, showing homodimeric or homotetrameric forms (Vicente et al., 2008b, 2009). A typical FDP consists of a core composed of a metallo-β-lactamase-like domain and a C-terminal flavodoxin domain. The former domain contains a nonheme diiron center, whereas the latter harbors a FMN moiety. It has been shown that FDPs in anaerobic bacteria are involved in O2 and/or NO detoxification (Vicente et al., 2008a). Completely unique FDPs are, however, found in specific groups of oxygenic photosynthetic organisms. FDPs found in cyanobacteria and some photosynthetic eukaryotes possess an extra C-terminal flavin reductase domain (Zhang et al., 2009). This particular domain composition theoretically allows NAD(P)H oxidation to be coupled with O2 reduction in the same enzyme.Synechocystis sp. PCC 6803 (hereafter Synechocystis), a widely used model organism among cyanobacteria in photosynthesis research, contains four FDPs encoded by the sll1521 (flv1), sll0219 (flv2), sll0550 (flv3), and sll0217 (flv4) genes. In vivo, Flv1 and Flv3 acquire electrons after PSI and deliver them further to molecular oxygen, reducing it to water. We have denominated this process as a Mehler-like reaction (Allahverdiyeva et al., 2013) because the excess of electrons is donated to O2, similarly to the genuine plant-type Mehler reaction, but there is no production of ROS in the FDP-mediated reaction (Helman et al., 2003). Up to 60% of the electrons produced by the oxygen splitting activity of PSII are redirected to Flv1- and Flv3-mediated Mehler-like reactions in severe inorganic carbon starvation conditions (Allahverdiyeva et al., 2011). Flv1 and Flv3 proteins form a very important electron sink that protects PSI against oxidative damage under fluctuating light conditions (Allahverdiyeva et al., 2013).Flv2 and Flv4 can be found only in cyanobacteria and they have been assigned a role in photoprotection of PSII (Zhang et al., 2009, 2012). PSII is historically known to be extremely vulnerable to oxidative damage upon illumination, with the severity of damage being dependent on light intensity and on the availability of electron acceptors. At air-level CO2 concentrations (low CO2 or LC) and/or high light (HL) irradiances, terminal acceptors are consistently limiting the electron flow, making PSII particularly sensitive to these conditions, widely exceeding the repair capacity of damaged PSII centers (Aro et al., 1993).Flv2 and Flv4 proteins are encoded in an operon including a small Sll0218 protein. Importantly, the flv4-2 operon is strongly induced in LC and HL conditions (Zhang et al., 2009). Flv4 and Flv2 proteins form a heterodimer that localizes in cytoplasm but also has a high affinity to membrane in the presence of cations (Zhang et al., 2012). Sll0218, the 19-kD protein encoded by the flv4-2 operon, locates in the thylakoid membrane and forms a high molecular mass complex in association with unknown partners. In the model proposed by Zhang et al. (2012), Sll0218 stabilizes PSII dimers and facilitates the opening of a novel electron transfer pathway through the Flv2/Flv4 heterodimer, which associates with the thylakoid membrane in light. The Flv2/Flv4 complex is also important for proper energy transfer from PBs to PSII as evidenced by a high emission peak at 685 nm in the 77K fluorescence spectra. This effect is caused by uncoupled PB terminal emitters, as deduced from detailed examination of the deconvoluted emission spectra (Zhang et al., 2012). This strongly suggested a distorted energy transfer from PB terminal emitters to the PSII reaction centers in flv4-2 operon deletion mutants. However, the photoprotection mechanism induced by the flv4-2 operon is not yet clearly understood. Here, with an overexpression approach, we provide evidence that Flv2/Flv4 acts as an important electron sink at the PSII acceptor side, allowing the maintenance of the plastoquinone (PQ) pool in an oxidized state and preventing the production of singlet oxygen in PSII. Furthermore, regular PBs are required for the Flv2/Flv4-related mechanism to be expressed. Genome mining of sequenced cyanobacteria strains provided evidence for the loss of the flv4-2 operon in the genomes of cyanobacteria that have acquired a stress-inducible D1 copy.  相似文献   

15.
16.
17.
The mitochondrial electron transport chain includes an alternative oxidase (AOX) that is hypothesized to aid photosynthetic metabolism, perhaps by acting as an additional electron sink for photogenerated reductant or by dampening the generation of reactive oxygen species. Gas exchange, chlorophyll fluorescence, photosystem I (PSI) absorbance, and biochemical and protein analyses were used to compare respiration and photosynthesis of Nicotiana tabacum ‘Petit Havana SR1’ wild-type plants with that of transgenic AOX knockdown (RNA interference) and overexpression lines, under both well-watered and moderate drought-stressed conditions. During drought, AOX knockdown lines displayed a lower rate of respiration in the light than the wild type, as confirmed by two independent methods. Furthermore, CO2 and light response curves indicated a nonstomatal limitation of photosynthesis in the knockdowns during drought, relative to the wild type. Also relative to the wild type, the knockdowns under drought maintained PSI and PSII in a more reduced redox state, showed greater regulated nonphotochemical energy quenching by PSII, and displayed a higher relative rate of cyclic electron transport around PSI. The origin of these differences may lie in the chloroplast ATP synthase amount, which declined dramatically in the knockdowns in response to drought. None of these effects were seen in plants overexpressing AOX. The results show that AOX is necessary to maintain mitochondrial respiration during moderate drought. In its absence, respiration rate slows and the lack of this electron sink feeds back on the photosynthetic apparatus, resulting in a loss of chloroplast ATP synthase that then limits photosynthetic capacity.The plant mitochondrial electron transport chain (ETC) is bifurcated such that electrons in the ubiquinone pool partition between the cytochrome (cyt) pathway (consisting of Complex III, cyt c, and Complex IV) and alternative oxidase (AOX; Finnegan et al., 2004; Millar et al., 2011; Vanlerberghe, 2013). AOX directly couples ubiquinol oxidation with O2 reduction to water. This reduces the energy yield of respiration because, unlike Complexes III and IV, AOX is not proton pumping. Hence, AOX is an electron sink, the capacity of which is little encumbered by rates of ATP turnover. In this way, AOX might be well suited to prevent cellular over-reduction. Supporting this, transgenic Nicotiana tabacum leaves with suppressed amounts of AOX have increased concentrations of mitochondrial-localized superoxide radical (O2) and nitric oxide, the products that can arise when an over-reduced ETC results in electron leakage to O2 or nitrite (Cvetkovska and Vanlerberghe, 2012, 2013).In angiosperms, AOX is encoded by a small gene family (Considine et al., 2002). In Arabidopsis (Arabidopsis thaliana), mutation or knockdown of the stress-responsive AOX1a gene family member dramatically reduces AOX protein and the capacity of the AOX respiration pathway to consume O2. Several studies have shown that this loss of AOX capacity in Arabidopsis aox1a plants affected processes such as growth, carbon and energy metabolism, and/or the cellular network of reactive oxygen species (ROS) scavengers (Fiorani et al., 2005; Umbach et al., 2005; Watanabe et al., 2008; Giraud et al., 2008; Skirycz et al., 2010). However, in studies in which respiration was measured, it was consistently reported that the lack of AOX capacity had no significant impact on the respiration rate in the dark (RD; Umbach et al., 2005; Giraud et al., 2008; Strodtkötter et al., 2009; Florez-Sarasa et al., 2011; Yoshida et al., 2011b; Gandin et al., 2012). The exceptions are two reports that RD was actually higher in aox1a than in the wild type under some conditions (Watanabe et al., 2008; Vishwakarma et al., 2014). To our knowledge, how the lack of AOX affects respiration rate in the light (RL) is not reported in Arabidopsis or other species.Numerous studies have established the importance of mitochondrial metabolism in the light to optimize photosynthesis (Hoefnagel et al., 1998; Raghavendra and Padmasree, 2003). In recent years, the potential importance of specifically AOX respiration during photosynthesis has been examined using the Arabidopsis aox1a plants (Giraud et al., 2008; Strodtkötter et al., 2009; Zhang et al., 2010; Florez-Sarasa et al., 2011; Yoshida et al., 2011a, 2011b). In general, these studies reported small perturbations of photosynthesis in standard-grown aox1a plants, including slightly lower rates of CO2 uptake or O2 release (Gandin et al., 2012; Vishwakarma et al., 2014), slightly higher rates of cyclic electron transport (CET; Yoshida et al., 2011b), and slightly increased susceptibility to photoinhibition after a high light treatment (Florez-Sarasa et al., 2011). Generally, these studies concluded that aox1a plants exhibit a biochemical limitation of photosynthesis, in line with the hypothesis that AOX serves as a sink for excess photogenerated reducing power, with the reductant likely reaching the mitochondrion via the malate valve (Noguchi and Yoshida, 2008; Taniguchi and Miyake, 2012). Similar to these Arabidopsis studies, we recently reported that well-watered N. tabacum AOX knockdowns grown at moderate irradiance display a slight reduced rate of photosynthesis (approximately 10%–15%) when measured at high irradiance. However, we established that the lower photosynthetic rate was the result of a stomatal rather than biochemical limitation of photosynthesis, and provided evidence that this stomatal limitation resulted from disrupted nitric oxide homeostasis within the guard cells of AOX knockdown plants (Cvetkovska et al., 2014).Drought is a common abiotic stress that can substantially curtail photosynthesis because stomatal closure, meant to conserve water, also restricts CO2 availability to the Calvin cycle. Besides this well established stomatal limitation of photosynthesis, there may also be water deficit-sensitive biochemical components that contribute to the reduction of photosynthesis during drought. However, the nature of this biochemical limitation and the degree to which it contributes to the curtailment of photosynthesis during drought remain areas of active debate (Flexas et al., 2004; Lawlor and Tezara, 2009; Pinheiro and Chaves, 2011). Additional factors, such as patchy stomatal closure (Sharkey and Seemann, 1989; Gunasekera and Berkowitz, 1992) or changes in the conductance to CO2 of mesophyll cells (Perez-Martin et al., 2009), can further complicate analyses of photosynthesis during drought.Metabolism can experience energy imbalances, when there is a mismatch between rates of synthesis and rates of utilization of ATP and/or NADPH, and the importance of mechanisms to minimize such imbalances has been emphasized (Cruz et al., 2005; Kramer and Evans, 2011; Vanlerberghe, 2013). For example, such imbalances may occur in the chloroplast when the use of ATP and NADPH by the Calvin cycle does not keep pace with the harvesting of light energy (Hüner et al., 2012). This can result in excess excitation energy that can damage photosynthetic components, perhaps through the generation of ROS (Asada, 2006; Noctor et al., 2014). Such a scenario has been hypothesized to underlie the development of the biochemical limitations of photosynthesis reported during drought (Lawlor and Tezara, 2009).In this study, we find that N. tabacum AOX knockdowns show a compromised rate of mitochondrial respiration in the light during moderate drought. This corresponds with a strong nonstomatal limitation of photosynthesis in these plants relative to the wild type, and we describe a biochemical basis for this photosynthetic limitation. The results indicate that AOX is a necessary electron sink to support photosynthesis during drought, a condition when the major photosynthetic electron sink, the Calvin cycle, is becoming limited by CO2 availability.  相似文献   

18.
19.
We have investigated the importance of carotenoids on the accumulation and function of the photosynthetic apparatus using a mutant of the green alga Chlamydomonas reinhardtii lacking carotenoids. The FN68 mutant is deficient in phytoene synthase, the first enzyme of the carotenoid biosynthesis pathway, and therefore is unable to synthesize any carotenes and xanthophylls. We find that FN68 is unable to accumulate the light-harvesting complexes associated with both photosystems as well as the RC subunits of photosystem II. The accumulation of the cytochrome b6f complex is also strongly reduced to a level approximately 10% that of the wild type. However, the residual fraction of assembled cytochrome b6f complexes exhibits single-turnover electron transfer kinetics comparable to those observed in the wild-type strain. Surprisingly, photosystem I is assembled to significant levels in the absence of carotenoids in FN68 and possesses functional properties that are very similar to those of the wild-type complex.Carotenoids (Cars) are fundamental components of the photosynthetic apparatus (Young and Britton, 1993, and refs. therein). The vast majority of Cars are noncovalently bound to either the core or the antenna subunits of PSI or PSII (Siefermann-Harms, 1985; Bassi et al., 1993). The most abundant Car bound to the core subunits of both photosystems is β-carotene, which is found in the vast majority of oxygenic organisms (Siefermann-Harms, 1985; Bassi et al., 1993). The light-harvesting complexes (LHCs) that act as the outer antenna in plants and green algae bind a wider range of oxygenated Cars, known as xanthophylls, the most abundant of which is lutein (Siefermann-Harms, 1985; Bassi et al., 1993; Jennings et al., 1996). The stoichiometry of xanthophylls binding to LHC complexes depends on the particular complexes and often on the illumination conditions during the organism’s growth (Siefermann-Harms, 1985; Demmig-Adams, 1990; Horton et al., 1996). Intriguingly, a molecule of β-carotene (as well as a molecule of chlorophyll [Chl] a) is found also in the cytochrome (Cyt) b6f complex (Kurisu et al., 2003; Stroebel et al., 2003).Cars have multiple functions in the photosynthetic process; they act as light-harvesting pigments (Frank and Cogdell, 1993), enlarging the optical cross section to radiation that is poorly absorbed by Chl. Moreover, Cars play a crucial role in processes such as nonphotochemical quenching that control the efficiency of light harvesting in response to the intensity of the incident radiation (for review, see Demmig-Adams, 1990; Horton et al., 1996; Niyogi, 1999). Probably the most important role of Cars in photosynthesis is the quenching of the excited triplet state of Chl (for review, see Frank and Cogdell, 1993; Giacometti et al., 2007), preventing the formation of highly reactive singlet oxygen, which represents the principal species active under high light stress (Hideg et al., 1994; Krieger-Liszkay, 2005). The importance of Cars is demonstrated by the observation that disruption of their biosynthesis through mutation, or by inhibition of a key enzyme in the pathway, leads to either lethal phenotypes or to rapid photobleaching of the photosynthetic tissue (Claes, 1957; Faludi-Dániel et al., 1968, 1970; Bolychevtseva et al., 1995; Trebst and Depka, 1997).Moreover, it has been shown that the presence of xanthophylls is absolutely necessary for refolding in vitro of LHC I and LHC II antenna complexes (Plumley and Schmidt, 1987; Paulsen et al., 1993; Sandonà et al., 1998). Such Cars, therefore, have a structural role, as well as their involvement in light harvesting, nonphotochemical quenching regulation, and the quenching of the Chl triplet state. Whether Cars also play a key structural role in the formation and stability of the core complexes of both PSI and PSII has not been systematically explored, since assembly of these complexes in vitro is not feasible. Studies in vivo using higher plants are complicated by the fact that Car deficiency is lethal and can be studied only during the early stages of greening and leaf development (Faludi-Dániel et al., 1968, 1970; Inwood et al., 2008). In these studies, it was shown that the accumulation of PSII complexes was greatly impaired in mutants of maize (Zea mays; Faludi-Dániel et al., 1968, 1970; Inwood et al., 2008), while the assembly of PSI appeared to be less sensitive to Car availability. In mutants of the cyanobacterium Synechocystis sp. PCC 6803 lacking the genes for phytoene desaturase or ζ-carotene desaturase, there was a complete loss of PSII assembly, while functional PSI complexes were assembled, albeit with slightly altered electron transfer kinetics with respect to the wild-type complex (Bautista et al., 2005). In agreement with the higher sensitivity of PSII assembly to Car availability, Trebst and Depka (1997) reported a specific effect on the synthesis of the D1 subunit of PSII RC upon treatment with phytoene desaturase inhibitors. On the other hand, it has recently been reported that in lycopene-β-cyclase mutants of Arabidopsis (Arabidopsis thaliana) that have a decreased amount of β-carotene (bound to the RC) with respect to most of the xanthophyll pool pigments (bound to the LHCs), the level of accumulation of PSI complexes, particularly that of the LHC I complement, was more affected that that of PSII, probably also because of an increased sensitivity to photodamage of mutated PSI RC (Cazzaniga et al., 2012; Fiore et al., 2012).In this investigation, we have studied the accumulation and functionality of the major chromophore-binding complexes of the photosynthetic apparatus, PSI, PSII, and Cyt b6f, in a Car-less mutant of the green alga Chlamydomonas reinhardtii (FN68) that is blocked at the first committed step of Car biosynthesis, namely, phytoene synthesis (McCarthy et al., 2004). Although the mutant is incapable of growing under phototrophic or photomixotrophic conditions, it can grow in complete darkness on a medium supplemented with a carbon source. Here, we show that the PSII core and antenna complexes fail to accumulate in the mutant and that the Cyt b6f complex accumulates to approximately one-tenth of the wild-type level. On the other hand, the PSI reaction center accumulates in FN68 and possesses electron transfer properties that are remarkably similar to those of wild-type PSI. Interestingly, we find that the level of PSI accumulation differs in other phytoene synthase null mutants, suggesting that additional mutations in one or other of these strains affect PSI stability. Nevertheless, our findings demonstrate that Cars are not required for either the assembly or the functionality of PSI in vivo.  相似文献   

20.
According to the “state transitions” theory, the light-harvesting complex II (LHCII) phosphorylation in plant chloroplasts is essential to adjust the relative absorption cross section of photosystem II (PSII) and PSI upon changes in light quality. The role of LHCII phosphorylation upon changes in light intensity is less thoroughly investigated, particularly when changes in light intensity are too fast to allow the phosphorylation/dephosphorylation processes to occur. Here, we demonstrate that the Arabidopsis (Arabidopsis thaliana) stn7 (for state transition7) mutant, devoid of the STN7 kinase and LHCII phosphorylation, shows a growth penalty only under fluctuating white light due to a low amount of PSI. Under constant growth light conditions, stn7 acquires chloroplast redox homeostasis by increasing the relative amount of PSI centers. Thus, in plant chloroplasts, the steady-state LHCII phosphorylation plays a major role in preserving PSI upon rapid fluctuations in white light intensity. Such protection of PSI results from LHCII phosphorylation-dependent equal distribution of excitation energy to both PSII and PSI from the shared LHCII antenna and occurs in cooperation with nonphotochemical quenching and the proton gradient regulation5-dependent control of electron flow, which are likewise strictly regulated by white light intensity. LHCII phosphorylation is concluded to function both as a stabilizer (in time scales of seconds to minutes) and a dynamic regulator (in time scales from tens of minutes to hours and days) of redox homeostasis in chloroplasts, subject to modifications by both environmental and metabolic cues. Exceeding the capacity of LHCII phosphorylation/dephosphorylation to balance the distribution of excitation energy between PSII and PSI results in readjustment of photosystem stoichiometry.Plant acclimation to different quantities and qualities of light has been extensively investigated. The light quality experiments have usually concerned the red/blue and far-red light acclimation strategies, which have been closely related to the state transitions and the phosphorylation of the light-harvesting complex II (LHCII) proteins, Lhcb1 and Lhcb2, by the state transition7 (STN7) kinase (Allen, 2003; Bellafiore et al., 2005; Bonardi et al., 2005; Tikkanen et al., 2006; Rochaix, 2007). Such studies on acclimation to different qualities of light have uncovered key mechanisms required for the maintenance of photosynthetic efficiency in dense populations and canopies (Dietzel et al., 2008). However, the role of LHCII phosphorylation under fluctuations in the quantity of white light has been scarcely investigated. Light conditions in natural environments may be very complex with respect to the quantity of white light, which constantly fluctuates both in short- and long-term durations (Smith, 1982; Külheim et al., 2002). Thus, the acclimation strategies to natural environments must concomitantly meet the challenges of both high- and low-light acclimation. Changing cloudiness, for example, would initiate both the high-light and low-light acclimation signals in the time scale of minutes and hours, whereas the movements of leaves in the wind or the rapid movement of clouds would initiate even more frequent light acclimation signals. The kinetics of reversible LHCII phosphorylation is far too slow to cope with rapid environmental changes.The phosphorylation level of LHCII proteins in the thylakoid membrane is regulated by both the STN7 kinase and the counteracting PPH1/TAP38 phosphatase (Pribil et al., 2010; Shapiguzov et al., 2010). No definite results are available about regulation of the PPH1/TAP38 phosphatase, but the STN7 kinase is strongly under redox regulation (Lemeille et al., 2009) and controls the phosphorylation level of LHCII proteins under varying white light intensities as well as according to chloroplast metabolic cues, as described already decades ago (Fernyhough et al., 1983; Rintamäki et al., 2000; Hou et al., 2003). So far, research on the role of the STN7 kinase and LHCII phosphorylation in the light acclimation of higher plants has heavily focused on reversible LHCII phosphorylation and concomitant state transitions. The state 1-to-state 2 transition, by definition, means the phosphorylation of LHCII proteins, their detachment from PSII in grana membranes, and migration to the stroma membranes to serve in the collection of excitation energy to PSI (Fork and Satoh, 1986; Williams and Allen, 1987; Wollman, 2001; Rochaix, 2007; Kargul and Barber, 2008; Murata, 2009; Lemeille et al., 2010; Minagawa, 2011). Concomitantly, the absorption cross section of PSII decreases and that of PSI increases (Canaani and Malkin, 1984; Malkin et al., 1986; Ruban and Johnson, 2009). Indeed, state transitions have been well documented when different qualities (blue/red and far red) of light, preferentially exciting either PSII or PSI, have been applied.Different from state transitions, the white light intensity-dependent reversible LHCII phosphorylation does not result in differential excitation of the two photosystems (Tikkanen et al., 2010). Instead, both photosystems remain nearly equally excited independently whether the LHCII proteins are heavily phosphorylated or strongly dephosphorylated. Moreover, it is worth noting that the different qualities of light generally used to induce reversible LHCII phosphorylation and state transitions (blue/red and far-red lights) have usually been of very low intensity (for review, see Haldrup et al., 2001), and apparently, minimal protonation of the lumen takes place under such illumination conditions. Yet another difference between induction of LHCII protein phosphorylation by different qualities of light or different quantities of white light concerns the concomitant induction of PSII core protein phosphorylation. In the former case, the level of PSII core protein phosphorylation follows the phosphorylation pattern of LHCII proteins, whereas under different quantities of white light, the phosphorylation behavior of PSII core and LHCII proteins is the opposite (Tikkanen et al., 2008b).To gain a more comprehensive understanding of the physiological role of white light-induced changes in LHCII protein phosphorylation, we have integrated Arabidopsis (Arabidopsis thaliana) LHCII phosphorylation with other light-dependent regulatory modifications of light harvesting and electron transfer in the thylakoid membrane, which include the nonphotochemical quenching of excitation energy (for review, see Niyogi, 1999; Horton and Ruban, 2005; Barros and Kühlbrandt, 2009; de Bianchi et al., 2010; Jahns and Holzwarth, 2012; Ruban et al., 2012) and the photosynthetic control of electron transfer by the cytochrome b6f (Cytb6f) complex (Rumberg and Siggel, 1969; Witt, 1979; Tikhonov et al., 1981; Bendall, 1982; Nishio and Whitmarsh, 1993; Joliot and Johnson, 2011; Suorsa et al., 2012; for review, see Foyer et al., 1990, 2012), both strongly dependent on lumenal protonation.It is demonstrated that the steady-state LHCII phosphorylation is particularly important under rapidly fluctuating light (FL) conditions. This ensures equal energy distribution to both photosystems, prevents the accumulation of electrons in the intersystem electron transfer chain (ETC), eliminates perturbations in chloroplast redox balance, and maintains PSI functionality upon rapid fluctuations in white light intensity.  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