Next Article in Journal
Barcoded Pyrosequencing Reveals a Shift in the Bacterial Community in the Rhizosphere and Rhizoplane of Rehmannia glutinosa under Consecutive Monoculture
Next Article in Special Issue
The Complete Chloroplast Genome Sequence of Tree of Heaven (Ailanthus altissima (Mill.) (Sapindales: Simaroubaceae), an Important Pantropical Tree
Previous Article in Journal
Epidermal Growth Factor, through Alleviating Oxidative Stress, Protect IPEC-J2 Cells from Lipopolysaccharides-Induced Apoptosis
Previous Article in Special Issue
Chloroplast Protein Turnover: The Influence of Extraplastidic Processes, Including Autophagy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metabolic Reprogramming in Chloroplasts under Heat Stress in Plants

1
The National Key Laboratory of Plant Molecular Genetics, Institute of Plant Physiology & Ecology, Chinese Academy of Sciences, 300 Fenglin Road, Shanghai 200032, China
2
CAS Center for Excellence in Molecular Plant Sciences, Institute of Plant Physiology & Ecology, Chinese Academy of Sciences, 300 Fenglin Road, Shanghai 200032, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2018, 19(3), 849; https://doi.org/10.3390/ijms19030849
Submission received: 29 January 2018 / Revised: 5 March 2018 / Accepted: 5 March 2018 / Published: 14 March 2018
(This article belongs to the Special Issue Chloroplast)

Abstract

:
Increases in ambient temperatures have been a severe threat to crop production in many countries around the world under climate change. Chloroplasts serve as metabolic centers and play a key role in physiological adaptive processes to heat stress. In addition to expressing heat shock proteins that protect proteins from heat-induced damage, metabolic reprogramming occurs during adaptive physiological processes in chloroplasts. Heat stress leads to inhibition of plant photosynthetic activity by damaging key components functioning in a variety of metabolic processes, with concomitant reductions in biomass production and crop yield. In this review article, we will focus on events through extensive and transient metabolic reprogramming in response to heat stress, which included chlorophyll breakdown, generation of reactive oxygen species (ROS), antioxidant defense, protein turnover, and metabolic alterations with carbon assimilation. Such diverse metabolic reprogramming in chloroplasts is required for systemic acquired acclimation to heat stress in plants.

1. Introduction

Photosynthesis is thought to be the most important photo-chemical reaction, during which sunlight is trapped and converted into biological energy in plants. In general, leaves function as a highly specialized organ that are basically appointed in the photosynthetic process in higher plants. Leaf photosynthesis is substantially affected, often lethally, by high temperature stress, usually 10–15 °C above an optimum temperature for plant growth as plants are not capable of moving to more favorable environments [1,2,3]. Housed in chloroplasts, the photosynthetic apparatus is susceptible to be damaged by heat stress and the chloroplasts have been demonstrated to play an essential role in activation of cellular heat stress signaling [4,5,6]. Given that Photosystem II (PSII) is the most susceptible target within the chloroplast thylakoid membrane protein complexes for heat stress, heat stress commonly causes severe thermal damages to PSII, dramatically affecting photosynthetic electron transfer and ATP synthesis [1,2,3,7,8]. Under heat stress, heat stress-induced damages lead to alterations of photochemical reactions in thylakoid lamellae of chloroplast, reflected as a significant reduction in the ratio of variable fluorescence to maximum fluorescence (Fv/Fm) [1,2,9,10,11]. Exposure to high temperature stress causes oxidative stress in plants, particularly leading to the dissociation of oxygen evolving complex (OEC) in PSII, which further results in inhibition of the electron transportation from OEC to the acceptor side of PSII [1,2,12,13]. Heat stress causes cleavage of the reaction center-binding protein D1 of PSII in spinach thylakoids and induces dissociation of a manganese (Mn)-stabilizing 33-kDa proteins from PSII reaction center complex [14]. As a photosynthetic carbon fixation cycle, the Calvin–Benson cycle is responsible for the fixation of CO2 into carbohydrates, as well as assimilation, transport, and utilization of photoassimilates as the organic products of photosynthesis. In addition to the disruption of OEC in PSII, heat stress also results in dysfunction in the system of carbon assimilation metabolism in the stroma of chloroplast [2,8]. It has been observed that the disruption of electron transport and inactivation of the oxygen evolving enzymes of PSII dramatically inhibit the rate of ribulose-1,5-bisphosphate (RuBP) regeneration [10,15]. Heat stress-induced inhibition in the activity of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) mainly results from inactivation of Rubisco activase that is extremely sensitive to heat stress because the enzyme Rubisco of higher plants is heat-stable [8,15]. In addition to the early effects on photochemical reactions and carbon assimilation, alterations in the microscopic ultrastructures of chloroplast and the integrity of thylakoid membranes were reported to be severely disrupted, including membrane destacking and reorganization when subjected to heat stress [2,16,17,18,19].
In photosynthetic organisms, chloroplasts respond to a variety of environmental stresses, including heat stress, with adjustments for major metabolic processes to optimize carbon fixation and growth requirements [2,3,6,8,20]. As one of the subcellular energy organelles, chloroplast in plant cells conducts major metabolic reprogramming processes including chlorophyll breakdown, generation and scavenging of reactive oxygen species (ROS), protein turnover, and metabolic alterations of carbon assimilation in response to heat stress (Figure 1). Here we review studies that have investigated the effects of heat stress on photosynthesis and its associated metabolic adaptations for optimizing plant growth and development under stress conditions. We will assess the role of chloroplast from an organellar perspective to begin building insights into better understanding the importance and significance of metabolic reprogramming within this organelle during high temperature stresses.

2. Chlorophyll Breakdown under Heat Stress

Chlorophyll (Chl) functions in harvesting light energy and driving electron transfer during the initial and indispensable processes of photosynthesis, and is a pigment consisting of two moieties: a chlorin ring containing a magnesium ion at its center and a long hydrophobic phytol chain, joined by an ester bond. It should be noted that when the photosynthetic apparatus is overexcited, and oxygen receives the absorbed energy from Chl, Chl acts as a harmful molecule, negatively affecting plant cells including most porphyrins [21,22]. Importantly, the breakdown of Chl is a physiological process as a prerequisite for protecting plant cells from hazardous effects of phototoxic pigments in association with recycling nitrogen sources from Chl-binding proteins in chloroplasts when leaves are senescing [22,23,24,25].
Recent studies have provided advances in better understanding of the pathway of Chl catabolism in higher plants. In brief, the removal of the phytol residue and the central Mg by chlorophyllase occurs at the initial reaction in the Chl breakdown pathway. Chlorophyllase genes, termed CLHs, were reported to be involved in Chl breakdown [26], but recent studies questioned the involvement of CLHs in Chl degradation in vivo during leaf senescence [23,27]. In 2009, functional analysis of pheophytinase (PPH) supports its role in porphyrin-phytol hydrolysis involved in senescence-related chlorophyll breakdown in vivo [28]. Several reports have revealed that the resulting pheophorbide (pheide) a is converted into a primary fluorescent chlorophyll catabolite (pFCC) in the next-step reactions, and two enzymes are involved in the reactions, including pheide a oxygenase (PAO) and red chl catabolite reductase (RCCR) [29,30,31,32]. PAO encodes a Rieske-type iron-sulfur oxygenase that is bound to chloroplast envelop, also named as ACCELERATED CELL DEATH (ACD1) [30]. PAO functions as a key enzyme that catalyzing the cleavage of the porphyrin ring in Chl breakdown pathway, and the red chlorophyll catabolite resulted from reactions can be further catalyzed into pFCC by RCCR. Next, a primary active transport system is responsible for exporting the resulting primary fluorescent catabolite pFCCs from the plastid and importing into the vacuole [33,34]. Based on the genetic analysis of the stay-green mutants, NON-YELLOW COLORING1 (NYC1) and NYC1-LIKE (NOL) were cloned and found to selectively retain photosystem II (PSII) light-harvesting complex subunits [35,36]. The first half of chlorophyll b can be catalyzed into chlorophyll a by chlorophyll b reductase that is composed of two subunits NYC1 and NOL [36]. In addition, leaves of the loss-of-function mutant pao showed a stay-green phenotype under dark-induction conditions and a light-dependent lesion mimic phenotype was also observed because of the increase in levels of phototoxic pheide a [31,37,38]. Interestingly, the Bf993 mutant of Festuca pratensis is classified into the third type of stay-green mutants in which the stay-green gene named as senescence-induced degradation (SID) is defective [39,40,41]. SGR (STAY-GREEN), designated as the orthologous gene of SID, has been cloned and characterized in a lot of plant species, including Arabidopsis [42], rice [43,44,45], pea [45,46], bell pepper [47,48], and tomato [47]. Based on the accumulated evidence, the direct interacting relationship has been characterized to exist between SGR and a subset of the protein components of the light harvesting chlorophyll a/b-protein complex II (LHCPII), suggesting that SGR is likely to play a role in making pigment-protein complexes unstable as a prerequisite for the enzymes in chlorophyll breakdown pathway to reach their substrate when leaves are in senescing processes [22,44,49].
A visible sign of leaf senescence and fruit ripening, loss of green color is resulted from massive Chl breakdown into nonphototoxic breakdown products in combination with carotenoid retention or anthocyanin accumulation [22,23,50]. Under normal growth conditions, chlorophyll maintains at a steady level due to the balance of its biosynthesis and degradation without a visible change in chlorophyll content [23,25]. In contrast, chlorophylls undergo turnover or breakdown when the partial or complete dismantling of the photosynthetic machinery occurs in response to environmental stresses, including heat stress. Heat stress symptoms in plants are typically characterized by leaf senescence or chlorosis due to a decline in chlorophyll content [2,51,52]. Heat-induced leaf chlorosis has been observed in a variety of plant species, including Arabidopsis [4,5,53], Soybean (Glycine max L. Merr.) [54], sorghum (Sorghum bicolor) [55], wheat (Triticum aestivum) [56,57] and creeping bentgrass (Agrostis stolonifera) [58]. However, the underlying mechanisms by which leaf senescence is regulated during heat stress remain elusive. Further studies are needed for answering the question of whether heat-induced chlorophyll loss is caused by heat-induced inhibition of chlorophyll synthesis and/or heat-enhanced chlorophyll degradation in plant leaves.
In Arabidopsis, chlorophyll a (Chl a) content was gradually reduced and chlorophyllase (Chlase) activity substantially increased during the high temperature treatment [53]. Upon heat treatment, no significant difference was detected in the enzyme activity of a key chlorophyll-synthesizing enzyme, porphobilinogen deaminase across all the lines of bentgrass (Agrostis spp.). However, the activities of chlorophyll-degrading enzymes, including chlorophyllase and chlorophyll-degrading peroxidase, increased significantly after heat stress whereas pheophytinase activity was unchanged [52]. Interestingly, lower activities of chlorophyll-degrading enzymes were detected in heat-tolerant transgenic lines in which the expression of isopentenyl transferase (ipt) gene is driven by a senescence-activated promoter (SAG12) or heat shock promoter (HSP18.2) in modulation of cytokinin biosynthesis when compared with the WT under heat stress [52]. The authors suggested that the enhanced degradation of chlorophyll under heat stress could result in a severe loss of the chlorophyll content in heat-challenged bentgrass. Studies on genetic variations contrasting in heat tolerance in the strength of leaf senescence under heat stress in hybrids of colonial (Agrostis capillaris) x creeping bentgrass (Agrostis stolonifera) supported that heat-induced loss of chlorophyll in leaves was caused by the rapid breakdown of chlorophyll, as manifested by the high-level activation of genes encoding chlorophyllase and pheophytinase, and the activity of pheophytinase (PPH) [59]. Recently, the map-based cloning of a semidominant, heat-sensitive, missense allele (cld1-1) led to identify a putative hydrolase, named as CHLOROPHYLL DEPHYTYLASE1 (CLD1) that is capable of dephytylating chlorophyll [60]. Their findings suggest that CLD1 is conserved in oxygenic photosynthetic organisms and plays a key role as the long-sought enzyme in making the phytol chain removed from chlorophyll in its degradation process at the steady-state level in chloroplasts.
Unlike studies on hormonal regulation of leaf senescence, less attention has been paid to effects of hormones on heat-induced chlorophyll loss. Exogenous application of a synthetic form of cytokinin, zeatin riboside (ZR) helped maintain higher leaf chlorophyll content creeping bentgrass exposed to heat stress by slowing down the action of protease and by induction or upregulation of heat-shock proteins [61]. The ethylene-inhibiting compound 1-methylcyclopropene (1-MCP) treatment can delay leaf senescence in cotton plants under high temperature by reducing lipid peroxidation, membrane leakage, soluble sugar content, and increasing chlorophyll content [62]. It was reported that the novel ethylene antagonist, 3-cyclopropyl-1-enyl-propanoic acid sodium salt (CPAS), increases grain yield in wheat by delaying leaf senescence under extreme weather conditions [63].
Plant chlorophyll retention-staygreen is considered a valuable trait under heat stress. Accumulating data indicate that understanding the physiological and molecular mechanisms of “STAY-GREEN” trait or delayed leaf senescence is required for regulating photosynthetic capability and may be also a key to break the plateau of productivity associated with adaptation to high temperature [64]. Staygreen traits are associated with heat tolerance in bread wheat and quantitative trait loci (QTL) for staygreen and related traits were identified across the genome co-located with agronomic and physiological traits associated to plant performance under heat stress, confirming that the staygreen phenotype is a useful trait for productivity enhancement in hot-irrigated environments [65]. In soybean, high temperatures and drought stress can lead to chlorophyll retention in mature seeds, termed as “green seed problem”, which is usually related to lower oil and bad seed quality, thus inducing a yield loss of soybean seeds. A “mild” stay-green phenotype was observed in a susceptible soybean cultivar when subjected to combined abiotic stresses of heat and drought stress and also the transcript levels of the STAY-GREEN 1 and STAY-GREEN 2 (D1, D2), PHEOPHORBIDASE 2 (PPH2) and NON-YELLOW COLORING 1 (NYC1) genes were downregulated in soybean seeds, indicating that the high-level transcriptional activation of these genes mentioned above in fully mature seeds is critical for a tolerant cultivar to cope with stresses and conduct a rapid and complete turnover of chlorophyll [66].
It is well-established that the major responses of crop plants to heat stress are classified into several aspects including the enhancement of leaf senescence, reduction of photosynthesis, deactivation of photosynthetic enzymes, and generation of oxidative damages to the chloroplasts. With respect to crop yield, heat stress also reduces grain number and size by affecting grain setting, assimilate translocation and duration and growth rate of grains [57]. In wheat, delayed senescence, or stay-green, contributes to a long grain-filling period and stable yield under heat stress [67]. Waxy maize (Zea mays L. sinensis Kulesh) is frequently exposed to high temperatures during grain filling in southern China. The heat-sensitive waxy maize variety exhibited a significant increase in the translocation amount and rate of assimilates pre-pollination and the accelerated leaf senescence phenotype under heat stress [68]. Short heat waves during grain filling can reduce grain size and consequently yield in wheat (Triticum aestivum L.). The four susceptible varieties of wheat showed greater heat-triggered reductions in final grain weight, grain filling duration and chlorophyll contents in flag leaves under heat stress, suggesting that grain size effects of heat may be driven by premature senescence [69,70].

3. Generation and Homeostasis of ROS in Chloroplasts under Heat Stress

As a significant source of reactive oxygen species (ROS) in plant cells, the chloroplast produces a variety of ROS such as hydrogen peroxide (H2O2), superoxide, hydroxyl radicals (OH), and 1O2 during photosynthesis [71]. The transfer of excitation energy in the PSII antenna complex and the electron transport in the PSII reaction center can be inhibited by a variety of abiotic stresses, resulting in the formation of ROS in algae and higher plants [72]. Given that ROS generation is an unavoidable consequence of aerobic metabolism, plants have evolved a large array of ROS-scavenging mechanisms [71]. ROS such as 1O2 is formed by the excitation energy transfer, whereas superoxide anion radical (O2•−), H2O2 and OH are formed by the electron transport [73]. In chloroplasts, ROS are mainly generated in the reaction centers of PSI and PSII in the chloroplast thylakoid membranes [71]. Photoreduction of oxygen to superoxide occurs in PSI and in PSII, oxygen of the ground (triplet) state of oxygen (3O2) is excited to the excited singlet state of oxygen (1O2) by the P680 reaction center chlorophyll [71]. Under extreme conditions, these ROS synthesis rates can increase, leading to an oxidative stress in both organelle and whole-cell functions [72,74]. It is well established that the chloroplast is extremely sensitive to high temperature stress during photosynthesis [1,2,9,10]. Accumulated data indicate that oxidative bursts of superoxide and/or hydrogen peroxide can be induced rapidly in response to a variety of environmental stresses, including also heat stress in plants [75,76,77]. It was reported that ROS, produced in PSI, PSII as well as in the Calvin-Benson cycle, can cause irreversible oxidative damage to cells when plants were subjected to heat stress [71,72,78]. Under high temperature conditions, large amounts of ROS were generated in tobacco cells for initiating signaling events involved in PCD, which is consistent with the role of applying the antioxidants ascorbate or superoxide dismutase (SOD) to the cultures in supporting the survival of cells [79]. In the leaves of tobacco (Nicotiana tabacum) defective in ndhC-ndhK-ndhJ (Delta ndhCKJ), hydrogen peroxide was rapidly produced in response to heat treatment, implying a role of the NAD(P)H dehydrogenase-dependent pathway in repressing generation of reactive oxygen species in chloroplasts [80]. In exposure to moderate heat treatment conditions, the oxidative damages of the reaction center-binding D1 protein of photosystem II increased and showed a tight positive relationship with the accumulated levels of 1O2 and OH in spinach PSII membranes, implying that inhibition of a water-oxidizing manganese complex led to a rapid production of ROS through lipid peroxidation under heat stress [81]. In Arabidopsis, large amounts of chlorophyllide a caused a surge of phototoxic singlet oxygen in the chlorophyll synthase mutant (chlg-1) under heat stress, suggesting that chlorophyll synthase acts in maintenance of ROS homeostasis in response to heat stress [82].
It is well known that ROS burst can trigger the oxidative damage to pigments, proteins, and lipids in the thylakoid membrane [71,73,83]. ROS are primarily agents of damage, but this view is questioned by data proving their beneficial role and particularly their signaling function. The chloroplast harbors ROS-producing centers (triplet chlorophyll, ETC in PSI and PSII) and a diversified ROS-scavenging network (antioxidants, SOD, APX-glutathione cycle, and a thioredoxin system) to keep the equilibrium between ROS production and scavenging [71]. The non-enzymatic and enzymatic ROS scavenging systems are engaged in preventing harmful effects of ROS on the thylakoid membrane components to keep ROS level in chloroplasts under control [71,73,74,83,84,85,86,87]. The efficient enzymatic scavenging systems are composed of several key enzymes, including superoxide dismutase (SOD), catalase (CAT), ascorbate peroxidase (APX), glutathione reductase (GR), monodehydroascorbate reductase (MDHAR), dehydroascorbate reductase (DHAR), glutathione peroxidase (GPX) and glutathione-S-transferase (GST) and non-enzymatic systems contain antioxidants such as ascorbic acid (ASH), glutathione (GSH), phenolic compounds, alkaloids, non-protein amino acids and alpha-tocopherols [21,71,73,83,87,88,89,90]. These antioxidant defense systems work in concert to maintain homeostasis of ROS, protecting plant cells from oxidative damage by scavenging of ROS.
In addition to triggering ROS burst, heat stress also affects the scavenging systems in plants. Heat induced degradation of chloroplast Cu/Zn superoxide dismutase as shown by reduced protein levels and isozyme-specific SOD activity [91]. Loss of Cu/Zn SOD and induction of catalase activity would explain the altered balance between hydrogen peroxide and superoxide under stress. The authors proposed that degradation of PSII could thus be caused by the loss of components of chloroplast antioxidant defense systems and subsequent decreased function of PSII [91]. In a recent study aiming at identifying heat protective mechanisms promoted by CO2 in coffee crop, the results likely favored that the maintenance of reactive oxygen species (ROS) at controlled levels contributed to mitigate of PSII photoinhibition under the high temperature stress [92]. Exogenous application of spermidine protected rice seedlings from heat-induced damage as marked by lower levels of malondialdehyde (MDA), H2O2, and proline content coupled with increased levels of Ascorbate (AsA), GSH, AsA and GSH redox status [93]. The authors conclude that heat exposure provoked an oxidative burden while enhancement of the antioxidative and glyoxalase systems by spermidine rendered rice seedlings more tolerant to heat stress. Under the later high temperature stress, heat priming contributed to a better redox homeostasis, as exemplified by the higher activities of superoxide dismutase (SOD) in chloroplasts and glutathione reductase (GR), and of peroxidase (POD) in mitochondria, which led to the lower superoxide radical production rate and malondialdehyde concentration in both chloroplasts and mitochondria [94]. These results suggested that heat priming effectively improved thermo-tolerance of wheat seedlings subjected to a later high temperature stress, which could be largely ascribed to the enhanced anti-oxidation at the subcellular level. In Kentucky bluegrass (Poa pratensis), higher activities of superoxide dismutase (SOD), catalase (CAT), peroxidase (POD), ascorbate peroxidase (APX), and glutathione reductase were detected in plants of heat-tolerant “Midnight” after long-term heat stress (21 and 28 d) in comparison with plants of heat-sensitive “Brilliant” [95]. Meanwhile, transcript levels of chloroplastic Cu/Zn SOD, Fe SOD, CAT, POD and cytosolic (cyt) APX were significantly higher in “Midnight” than in “Brilliant” under long-term heat stress. These results supported the hypothesis that enzymatic ROS scavenging systems could play predominant roles in antioxidant protection against oxidative damages from long-term heat stress. In tomato, studies on the role of a tomato (Lycopersicon esculentum) chloroplast-targeted DnaJ protein (LeCDJ1) showed that the sense transgenic tomato plants were more tolerant to heat stress due to the higher activities of ascorbate peroxidase (APX) and superoxide dismutase (SOD) [96]. The high temperature stress in southern China is one of the major factors leading to loss of the yield and quality of wucai (Brassica campestris L.). Comparative investigations on two cultivars of wucai (heat-sensitive and heat-tolerant) showed that greater severity of damage to the photosynthetic apparatus and membrane system was observed in heat-sensitive cultivar probably due to a high-level accumulation of ROS and malondialdehyde (MDA) [97]. In line with the studies mentioned above, plants of heat-tolerant Cucurbit species exhibited comparatively little oxidative damage, with the lowest hydrogen peroxide (H2O2), superoxide (O2−) and malondialdehyde (MDA) compared with the thermolabile and moderately heat-tolerant interspecific inbred line [98]. The enzyme activities of superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT) and peroxidase (POD) were found to be increased with heat stress in tolerant genotypes and the significant inductions of FeSOD, MnSOD, APX2, CAT1 and CAT3 isoforms in tolerant genotypes suggested their participation in heat tolerance [98].
Generally, the chloroplast possesses a variety of constitutively expressed antioxidant defense mechanisms to scavenge the various types of ROS generated during heat stress in preventing and minimizing oxidative damage to biological macromolecules (Figure 2). Therefore, the capacity of antioxidant defense is critical for heat stress adaptation and its strength is correlated with acquisition of thermotolerance with respect to buffering the effect of heat stress on the metabolic system [3,74,99,100]. On the other hand, ROS are produced in chloroplasts can function as plastid signals to inform the nucleus to activate the expression of genes encoding antioxidant enzyme and to adjust the stress-responsive machinery for more efficient adaptation to environmental stresses [6,47,72,74,100,101].

4. Turnover of PSII Core Subunits and PSII Protection under Heat Stress

The deleterious effects of high temperatures on proteins in chloroplasts include protein denaturation and aggregation [18,102,103]. ROS generation is one of the earliest responses of plant cells in response to heat stress and chloroplasts are the main targets of ROS-linked damage [71,102]. Given that chloroplasts are a major site of protein degradation, turnover of the damaged proteins is critical for plants to adapt to heat stress through the process of acclimation. As a multisubunit thylakoid membrane pigment-protein complex, PSII is vulnerable to light and heat damages that inhibit light-driven oxidation of water and reduction of plastoquinone [2,71]. PSII produces ROS, responsible for the frequent damage and turnover of this megacomplex that occur under physiological and stress conditions [71]. Accumulating data suggest that although more than 40 proteins are known to associate with PSII, the damage mainly targeted to one of its core proteins, the D1 protein under light and heat stress conditions [18,102,104,105]. Importantly, D1 protein has been demonstrated as a very susceptible target of 1O2; on the other hand, it appears to function as a major scavenger of 1O2 due to its close localization to the site of 1O2 formation in the reaction center of PSII [106]. In addition to the D1 protein, β-carotene, plastoquinol, and a-tocopherol have also shown to play a role in scavenging 1O2 and protect PSII against photo-oxidative damage [107,108]. Under heat stress conditions, two processes have been characterized in the thylakoids: dephosphorylation of the D1 protein in the stroma thylakoids, and aggregation of the phosphorylated D1 protein in the grana [102]. Heat stress also induced the release of the extrinsic Photosystem II Subunit O (PsbO), P and Q proteins from Photosystem II, which affected D1 degradation and aggregation significantly [102]. In spinach thylakoids, cleavage of the D1 protein was detected in response to moderate heat stress (40 °C, 30 min), producing an N-terminal 23-kDa fragment, a C-terminal 9-kDa fragment, and aggregation of the D1 protein [102,103]. ROS are known to specifically modify PSII proteins. Using high-resolution tandem mass spectrometry, oxidative modifications were identified on 36 amino acid residues on the lumenal side of PSII, in the core PSII proteins D1, D2, and CP43 of the cyanobacterium Synechocystis sp. PCC 6803, providing the compelling evidence to date of physiologically relevant oxidized residues in PSII [109]. Taken together, the oxidative damage of the D1 protein is caused by reactive oxygen species, mostly singlet oxygen, and also by endogenous cationic radicals generated by the photochemical reactions of photosystem II [71,73]. Under heat stress, the damage to the D1 protein by moderate heat stress is due to reactive oxygen species produced by lipid peroxidation near photosystem II [73,110]. Moreover, the damage of the D1 protein has been shown to be directly proportional to light intensity [111,112] or strength of heat stress [102]. D1 protein is well characterized as a protein of high turnover rate due to its rapid degradation when oxidized after its interaction with 1O2 and its replacement by newly synthesized D1 polypeptides [105]. It was reported that the D1 protein was shown to have the half-life of 2.4 h, the fourth fastest turn-over rate of barley (Hordeum vulgare) proteins when plants were growing under normal growth light intensity (500 µmol · m−2 · s−2) [113]. Meanwhile, higher degradation rates of the D2, CP43 and PsbH subunits were also detected when compared with the other PSII subunits [113,114,115]. The maintenance of PSII activity is critical, but also a challenge for oxygenic photosynthetic organisms to survive under normal growth or stress conditions.
Thus, replacing the photo- or heat-damaged D1 with a newly synthesized copy is essential for maintaining PSII activity [105,116]. In chloroplasts, the damaged D1 is degraded by a concerted action of particular filamentation temperature sensitive H (FtsH) protease and Deg (for degradation of periplasmic proteins) isoforms during its rapid and specific turnover and replaced with a de novo synthesized one in a system which is termed the PSII repair cycle [105]. It has been a subject of extensive studies about the basic concept for replacement of the damaged D1 protein by a newly-synthetized copy in the PSII repair cycle [105,116,117,118,119]. Repair of damaged D1 protein in PSII includes five steps: (1) migration of damaged PSII core complex to the stroma thylakoid, (2) partial PSII disassembly of the PSII core monomer, (3) access of protease degrading damaged D1, (4) concomitant D1 synthesis, and (5) reassembly of PSII into grana thylakoid [105,116,120,121]. Deg /HtrA (for high temperature requirement A) proteases, a family of serine-type ATP-independent proteases, have been shown in higher plants to be involved in the degradation of the Photosystem II reaction center protein D1 [105,116,122]. In Arabidopsis, five DEGs (Deg1, Deg2, Deg5, Deg7, and Deg8) have shown to be peripherally attached to the thylakoid membrane of chloroplasts: Deg1, Deg5, and Deg8 are localized on the lumenal side, and Deg2 and Deg7 are localized on the stromal side [123,124,125]. DEG5 and DEG8 may have synergistic function in degradation of D1 protein in the repair cycle of PSII under heat stress based on functional analysis of deg5, deg8 and the double mutant deg5 deg8 of Arabidopsis thaliana [126]. Under photoinhibitory conditions, cooperative degradation of D1 by Deg and FtsH has been demonstrated in vivo, in which Deg cleavage assists FtsH processive degradation [127]. In Arabidopsis, FtsH11-encoded protease have shown to play a direct role in thermotolerance, a function previously reported for bacterial and yeast FtsH proteases [128]. It should be noted that photosynthetic capability and PSII quantum yield are greatly reduced in the leaves of FtsH11 mutants when exposed to the moderately high temperature whereas under high light conditions, FtsH11 mutants and wild-type plants showed no significant difference in photosynthesis capacity [128]. In general, several possible mechanisms have been proposed for activation of these proteases, which depend on oligomerization of the monomer subunits [129]. In line with the hypothesis that hexamers of the FtsH proteases are probably localized near the Photosystem II complexes at the grana, degradation of the D1 protein could take place in the grana rather than in the stroma thylakoids to circumvent long-distance migration of both the Photosystem II complexes containing the photodamaged D1 protein and the proteases [129]. Under high light conditions, the lumen-exposed loops of the D1 protein at specific sites were cleaved by all of three lumenal serine proteases, Deg1, Deg5 and Deg8, during PSII repair cycle [130,131,132,133]. It was reported that Deg5 and Deg8 interact to form an active protease complex under high light [133]. Interestingly, Deg1 is activated when the Deg1 monomers are transformed into a proteolytically active hexamer at acidic pH upon protonation of a histidine amino acid residue [134]. In addition to functioning as a protease, Deg1 also plays a novel role as a chaperone/assembly factor of PSII [135]. In addition, Deg1 has been shown to be responsible for the proteolytic activity against the PsbO protein in vitro [136]. During the PSII repair cycle, only the damaged reaction center protein D1 and occasionally also the D2, CP43 and PsbH subunits are replaced while the other protein components of the complex are recycled, indicating that many aspects of PSII repair cycle and de novo biogenesis are partially overlapping [113,114,115,137]. Originally, the vulnerability of PSII to high light or heat stress was taken as an inherent fault of the photosynthetic machinery. However, recent studies strongly support that the constant, yet highly regulated, photodamage and repair cycle of PSII are of a strong physiological basis. Collectively, the photodamage of PSII is likely to act as a PSI protection mechanism instead of being considered solely as an undesired consequence of the highly oxidizing chemistry of the water splitting PSII [105].
The synthesis of heat shock proteins (HSPs) is characterized as a major response of all organisms responding to heat stress. The HSPs act as chaperones by assisting in protein folding and preventing irreversible protein aggregation [138,139]. A chloroplast-localized sHSP, HSP21, has been identified in diverse higher plant species, including both dicots and monocots [139,140] and its precursor polypeptide is ~5 kD larger than the mature protein [141,142]. HSP21 is thought to protect photosynthetic electron transport, specifically that of Photosystem II, during heat stress [143,144,145,146,147] and oxidative stress [145,148]. In addition, HSP21 has been demonstrated to play a dual role in protecting PSII from oxidative stress and promoting color changes during fruit maturation whereas no protective effects for the transgene were detected on PSII thermotolerance [149]. Importantly, studies using the transgenic tomato plants overexpressing HSP21 have shown that this protein associates with proteins of Photosystem II and does not reactivate heat-denatured Photosystem II, but instead protects this complex from damage during heat stress [150]. Interestingly, around two thirds of chloroplast HSP21 proteins are translocated into the thylakoid membranes in response to heat treatment in plants, suggesting that the association with membranes should be considered to fully understand the role of sHsps in physiological adaptation processes under stress conditions [151]. Despite extensive studies on HSP21, the specific roles of HSP21 in protecting PSII from heat stress remain elusive. Recently, HSP21 has been demonstrated to protect PSII from heat stress-induced damages by directly binding to D1 and D2 proteins, the core subunits of PSII. Importantly, heat-responsive transcriptional activation of HSP21 is regulated by the chloroplast retrograde signaling pathway in which GUN5 acts as a determinant upstream signaling component in Arabidopsis [5]. Based on these findings, an auto-adaptation loop working module has emerged in which the GUN5-dependent plastid signal(s) is triggered in response to heat stress and in turn communicated into the nucleus to activate the heat-responsive expression of HSP21 for optimizing particular demands of chloroplasts in making photosynthetic complexes stable during adaptation to heat stress in plants [5,6].

5. Effects of Heat Stress on Metabolic Flux through the Calvin-Benson-Bassham Cycle

In the Calvin-Benson-Bassham Cycle, Ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) plays a critical role in catalyzing the carboxylation of the 5-carbon sugar ribulose-1,5-bisphosphate (RuBP) when atmospheric CO2 during is fixed during photosynthesis. Rubisco activase (RCA) regulates the activity of Rubisco by facilitating the dissociation of inhibitory sugar phosphates from the active site of Rubisco in an ATP-dependent manner [152]. Extensive evidence supports the conclusion that reduction of plant photosynthesis arises primarily from thermal inactivation of Rubisco activity due to the inhibition of RCA under moderately elevated temperatures [8,152,153,154,155]. In addition to Rubisco activation, electron transport activity, ATP synthesis, and RuBP regeneration are also inhibited by moderately heat stress [156,157,158]. As the temperature increases further above the thermal optimum, the physical integrity of electron transport components of the photosynthetic apparatus can be severely damaged, leading to the increased limitation in photosynthesis [15]. It has been the subject of extensive investigations of elucidating the biochemical basis for the decrease in Rubisco activation state under heat stress [8,15,153,154,159]. Initially, studies on thermal stability of purified RCA showed that heat treatment only slightly inhibited the activities of this enzyme [160,161] and later experiments confirmed that heat stress caused thermal denaturation of activase in wheat and cotton leaves [162]. In line with the thermal denaturation of RCA under heat stress, Feller et al. (1998) suggested that RCA exhibited the exceptional thermal lability in vivo and the thermal stabilities of activases were different in plants from contrasting thermal environments [15,159]. Thus, loss of activase activity during heat stress is caused by an exceptional sensitivity of the protein to thermal denaturation and is responsible, in part, for deactivation of Rubisco [163]. It has been assumed that the stability of RCA could be influenced by heat-induced changes either in redox state [11,164,165] or the concentrations of ions, nucleotides, or other chloroplast constituents in plants [154]. On the other hand, the thermotolerance of Rubisco activase has been proposed to be responsible for restricting the distribution of certain plant species [154] as demonstrated by the response of Rubisco activase activity to temperature for cotton, a warm-season species, and Camelina sativa, a cool-season species. With respect to the effects of high growth temperature on the relative contribution of diffusive and biochemical limitations to photosynthesis, our knowledge is limited although there is abundant evidence that photosynthesis can acclimate to temperature [166,167]. Accumulating data suggest that the biochemical mechanisms about the decrease in Rubisco activation can be attributed to: (1) more rapid de-activation of Rubisco caused by a faster rate of dead-end product formation; and (2) slower re-activation of Rubisco by activase [168]. In a word, the resulting consequence is that RCA becomes less effective in keeping Rubisco catalytically competent as temperature increases.
Inhibition of net photosynthesis by heat stress has been attributed to an inability of Rubisco activase to maintain Rubisco in an active form because of the low thermal stability of Rubisco's chaperone, activase. These results support a role for RCA in limiting photosynthesis at high temperature when the temperature exceeds the optimum range for plants. In cotton (Gossypium hirsutum L.), activase gene expression is influenced by post-transcriptional mechanisms that may contribute to acclimation of photosynthesis during extended periods of heat stress [169]. In wheat, northern blot analysis showed maximum accumulation of TaRCA1 transcript in thermotolerant cv. during mealy-ripe stage, as compared to thermosusceptible ones [170]. To test the hypothesis that thermostable RCA can improve photosynthesis under elevated temperatures, gene shuffling technology was used to generate several Arabidopsis thaliana RCA1 (short isoform) variants exhibiting improved thermostability [171]. In line with the findings mentioned above, transgenic Arabidopsis lines expressing a thermostable chimeric activase showed higher rates of photosynthesis than the wild type after a short exposure to higher temperatures and they also recovered better, when they were returned to the normal temperature [172]. The results showed that photosynthesis and growth were improved under moderate heat stress in transgenic Arabidopsis expressing these thermotolerant RCA isoforms, providing evidence that manipulation of activase properties can improve C3 photosynthesis. In addition, the transcriptional level of wheat RCA (45–46 kDa) positively correlated with the yield of plants under heat-stress conditions in a very significant and linear manner [173]. At present, accumulating data indicate that RCA could affect plant productivity in relation to its endogenous levels under temperature stress conditions. Critically, RCA as the molecular chaperone plays a key role in constant engagement and remodeling of Rubisco to maintain metabolic flux through the Calvin-Benson-Bassham cycle as Rubisco is characterized as a dead-end inhibited complex in higher plants. In plants of the crassulacean acid metabolism (CAM), it has been assumed that possessing thermostable RCA is necessary for these plants to support the metabolic flux of Calvin-Benson-Bassham cycle when closure of stomata is a limitation factor during the day [174]. It is interesting that the CAM Rca isoforms (Agave tequilana) were found to be approximately 10 °C more thermostable when compared with the C3 isoforms of Rca isolated from rice (Oryza sativa) [174]. Interestingly, sequence analysis and immuno-blotting identified the beta-subunit of chaperonin-60 (cpn60 beta), the chloroplast GroEL homologue, as a protein that was bound to Rubisco activase from leaf extracts prepared from heat-stressed, but not control plants [175]. Rubisco requires RCA, an AAA+ ATPase that reactivates Rubisco by remodeling the conformation of inhibitor-bound sites. RCA is regulated by the ratio of ADP:ATP, with the precise response potentiated by redox regulation of the alpha-isoform [176]. Given that RCA uses the energy from ATP hydrolysis to restore catalytic competence to Rubisco, manipulation of RCA by redox regulation of the a-isoform might provide a strategy for enhancing photosynthetic performance in Arabidopsis [177]. In rice, heat stress significantly induced the expression of RCA large isoform (RCAL) as determined by both mRNA and protein levels and correlative analysis indicated that and RCA small isoform (RCAS) protein content was very tightly correlated to Rubisco initial activity and net photosynthetic rate under both heat stress and normal conditions [178]. In two Populus species adapted to contrasting thermal environments, the difference in the primary sequence of Rubisco activases between the species is more significant in the regions conferring ATPase activity and Rubisco recognition, suggesting that the genotypic distinctive characterizations in Rubisco activase are likely to underlie the specificities with respect to the heat-sensitive strength of Rubisco activase and photosynthesis under moderate high temperature conditions [179]. Recent studies on the effects of heat and drought on three major cereal crops, including rice, wheat, and maize, indicate that reductions in Rubisco activation might be not dependent on the amount of Rubisco and RCA, but could be resulted from the inhibition of RCA activity, as evidenced by the mutual reduction and positive relationship existed between the activation state of Rubisco and the rate of electron transport [153]. Critically, Rubisco activase acts as a key player in photosynthesis under heat stress conditions (non-stomatal limitation) [180]. When exposed to a moderate heat stress, Rca can be inhibited reversibly, but is irreversibly inhibited under a higher temperature and/or longer exposure due to heat stress-induced insolubilization and degradation of the Rca protein [180].

6. Conclusions and Perspectives

In many regions of the world, high temperature stress is one of the most important constraints to plant growth and productivity, especially for crop plants. The mechanism underlying the development of heat-tolerance for important agricultural crops as well as plant responses and adaptation to elevated temperatures needs to be better understood. Extensive studies have shown that metabolic regulation of adaptation processes during heat stress is not only an important developmental process, but also allows for flexibility of physiological responses to heat stress. In photosynthetic organisms, heat stress can affect photosynthesis through altered carbon assimilation metabolism in chloroplasts with remobilizing their starch reserve to release energy, sugars and derived metabolites in order to help mitigate the stress. This is thought to be an essential process for plant fitness with important implications for plant productivity under high temperature stress. One future challenges is to dissect the complex interaction networks between heat stress sensing, signal transduction and activations of key genes involved in metabolic reprogramming in coordination with developmental programmes. Accumulation and modification of metabolites in chloroplasts under heat stress may play a key role in the regulation of adaptation processes at cellular levels in plants, allowing plants to interact with their environment and to activate cellular heat stress responses at the optimal time in order to maintain photosynthesis. This kind of metabolic reprogramming is critical for plants to survive stress periods, and to prevent further damage to the whole plant.
The role of chloroplast in the metabolic regulation of heat stress responses has attracted increasing attention and extensive investigations from an organellar perspective have provided insights into better understanding the hypothesis stated that the heat stress-induced reprogramming, including decline in photosynthesis and alterations in photosynthetic metabolites which, in turn, could act as signal(s) or trigger the initial signal cascades to activate cellular heat stress responses. The present knowledge concerning the interplay between the chloroplast and nucleus in heat stress signal perception and activation of cellular heat stress responses is emerging, but more efforts are needed to reach a detailed overview. It can be predicted that uncovering the molecular mechanisms of heat sensing will pave the way to engineering plants capable of tolerating heat stress. It is well known that the ability of plants adapting to different climate regimes vary dramatically across and within species. Identification and functional analysis of the valuable heat-tolerant genetic resources will bring about a further significant improvement in manipulation of photosynthesis to increase crop yield based on a direct comparative analysis between the different manipulations with all the transgenic and wild-type plants grown and assessed in parallel under filed growth conditions. Thus, in-depth analyses of the interactions between the chloroplast and nucleus in heat stress responses are likely to be in focus during forthcoming years. On the other hand, Rubisco activase and enzymes functioning in the detoxification of reactive oxygen species are thought to be critical targets for breeding heat-tolerant crop plants with high yields under high temperature stress.

Acknowledgments

This study was supported by Chinese Academy of Sciences (Strategic Priority Research Program XDPB0404), the Ministry of Science and Technology of China (National Key R&D Program of China, 2016YFD0100405), and the National Natural Science Foundation of China (31770314 and 31570260).

Author Contributions

Fang-Qing Guo and Qing-Long Wang conceived and designed the outline and contents of this review article; Fang-Qing Guo and Qing-Long Wang wrote the paper with inputs from Juan-Hua Chen and Ning-Yu He.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

Abbreviations

1-MCP1-methylcyclopropene
1O2singlet state of oxygen
3O2triplet state of oxygen
ACD1Accelerated cell death
APXascorbate peroxidase
AsAAscorbate
ASHascorbic acid
ATPadenosine 5′-triphosphate
CAMcrassulacean acid metabolism
CATcatalase
ChlChlorophyll
Chl aChlorophyll a
Chlasechlorophyllase
CLD1CHLOROPHYLL DEPHYTYLASE1
CLHsChlorophyllase genes
CPAS3-cyclopropyl-1-enyl-propanoic acid sodium salt
cpn60 betabeta-subunit of chaperonin-60
Deg/HtrAhigh temperature requirement A
DHARdehydroascorbate reductase
FtsHfilamentation temperature sensitive H
Fv/Fmvariable fluorescence to maximum fluorescence
GPXglutathione peroxidase
GRglutathione reductase
GSHglutathione
GSTglutathione-S-transferase
GUN5genomes uncoupled 5
H2O2hydrogen peroxide
HSPsheat shock proteins
iptisopentenyl transferase
LeCDJ1tomato (Lycopersicon esculentum) chloroplast-targeted DnaJ protein
LHCPIIlight harvesting chlorophyll a/b-protein complex II
MDAmalondialdehyde
MDHARmonodehydroascorbate reductase
NAD(P)Hnicotinamide adenine dinucleotide phosphate (NADP)
NOLNYC1-LIKE
NYC1NON-YELLOW COLORING1
O2•−superoxide anion radical
OECoxygen evolving complex
OHhydroxyl radicals
PAOpheide a oxygenase
pFCCprimary fluorescent chlorophyll catabolite
pheidepheophorbide
PODperoxidase
PPHpheophytinase
PPH2PHEOPHORBIDASE 2
PsbOPHOTOSYSTEM II SUBUNIT O
PSIIPhotosystem II
QTLquantitative trait loci
RCARubisco activase
RCALRCA large isoform
RCASRCA small isoform
RCCRRed chl catabolite reductase
ROSReactive oxygen species
RubiscoRibulose-1,5-bisphosphate carboxylase/oxygenase
RuBPRibulose-1,5-bisphosphate
SGRSTAY-GREEN
SIDsenescence-induced degradation
SODsuperoxide dismutase
ZRzeatin riboside

References

  1. Wahid, A.; Gelani, S.; Ashraf, M.; Foolad, M.R. Heat tolerance in plants: An overview. Environ. Exp. Bot. 2007, 61, 199–223. [Google Scholar] [CrossRef]
  2. Allakhverdiev, S.I.; Kreslavski, V.D.; Klimov, V.V.; Los, D.A.; Carpentier, R.; Mohanty, P. Heat stress: An overview of molecular responses in photosynthesis. Photosynth. Res. 2008, 98, 541–550. [Google Scholar] [CrossRef] [PubMed]
  3. Berry, J.; Bjorkman, O. Photosynthetic Response and Adaptation to Temperature in Higher-Plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1980, 31, 491–543. [Google Scholar] [CrossRef]
  4. Yu, H.-D.; Yang, X.-F.; Chen, S.-T.; Wang, Y.-T.; Li, J.-K.; Shen, Q.; Liu, X.-L.; Guo, F.-Q. Downregulation of Chloroplast RPS1 Negatively Modulates Nuclear Heat-Responsive Expression of HsfA2 and Its Target Genes in Arabidopsis. PLoS Genet. 2012, 8, e1002669. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, S.-T.; He, N.-Y.; Chen, J.-H.; Guo, F.-Q. Identification of core subunits of photosystem II as action sites of HSP21, which is activated by the GUN5-mediated retrograde pathway in Arabidopsis. Plant J. 2017, 89, 1106–1118. [Google Scholar] [CrossRef] [PubMed]
  6. Sun, A.-Z.; Guo, F.-Q. Chloroplast Retrograde Regulation of Heat Stress Responses in Plants. Front. Plant Sci. 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  7. Havaux, M. Characterization of Thermal-Damage to the Photosynthetic Electron-Transport System in Potato Leaves. Plant Sci. 1993, 94, 19–33. [Google Scholar] [CrossRef]
  8. Sharkey, T.D. Effects of moderate heat stress on photosynthesis: Importance of thylakoid reactions, rubisco deactivation, reactive oxygen species, and thermotolerance provided by isoprene. Plant Cell Environ. 2005, 28, 269–277. [Google Scholar] [CrossRef]
  9. Yamada, M.; Hidaka, T.; Fukamachi, H. Heat tolerance in leaves of tropical fruit crops as measured by chlorophyll fluorescence. Sci. Hortic. 1996, 67, 39–48. [Google Scholar] [CrossRef]
  10. Wise, R.R.; Olson, A.J.; Schrader, S.M.; Sharkey, T.D. Electron transport is the functional limitation of photosynthesis in field-grown Pima cotton plants at high temperature. Plant Cell Environ. 2004, 27, 717–724. [Google Scholar] [CrossRef]
  11. Sharkey, T.D.; Schrader, S.M. High temperature stress. In Physiology and Molecular Biology of Stress Tolerance in Plants; Springer: New York, NY, USA, 2006; pp. 101–129. [Google Scholar]
  12. Havaux, M.; Tardy, F. Temperature-dependent adjustment of the thermal stability of photosystem II in vivo: Possible involvement of xanthophyll-cycle pigments. Planta 1996, 198, 324–333. [Google Scholar] [CrossRef]
  13. Klimov, V.V.; Baranov, S.V.; Allakhverdiev, S.I. Bicarbonate protects the donor side of photosystem II against photoinhibition and thermoinactivation. FEBS Lett. 1997, 418, 243–246. [Google Scholar] [CrossRef]
  14. Yamane, Y.; Kashino, Y.; Koike, H.; Satoh, K. Effects of high temperatures on the photosynthetic systems in spinach: Oxygen-evolving activities, fluorescence characteristics and the denaturation process. Photosynth. Res. 1998, 57, 51–59. [Google Scholar] [CrossRef]
  15. Salvucci, M.E.; Crafts-Brandner, S.J. Relationship between the heat tolerance of photosynthesis and the thermal stability of Rubisco activase in plants from contrasting thermal environments. Plant Physiol. 2004, 134, 1460–1470. [Google Scholar] [CrossRef] [PubMed]
  16. Gounaris, K.; Brain, A.R.R.; Quinn, P.J.; Williams, W.P. Structural Reorganization of Chloroplast Thylakoid Membranes in Response to Heat-Stress. Biochim. Biophys. Acta 1984, 766, 198–208. [Google Scholar] [CrossRef]
  17. Semenova, G.A. Structural reorganization of thylakoid systems in response to heat treatment. Photosynthetica 2004, 42, 521–527. [Google Scholar] [CrossRef]
  18. Yamamoto, Y.; Aminaka, R.; Yoshioka, M.; Khatoon, M.; Komayama, K.; Takenaka, D.; Yamashita, A.; Nijo, N.; Inagawa, K.; Morita, N.; et al. Quality control of photosystem II: Impact of light and heat stresses. Photosynth. Res. 2008, 98, 589–608. [Google Scholar] [CrossRef] [PubMed]
  19. Vani, B.; Saradhi, P.P.; Mohanty, P. Alteration in chloroplast structure and thylakoid membrane composition due to in vivo heat treatment of rice seedlings: Correlation with the functional changes. J. Plant Physiol. 2001, 158, 583–592. [Google Scholar] [CrossRef]
  20. Kmiecik, P.; Leonardelli, M.; Teige, M. Novel connections in plant organellar signalling link different stress responses and signalling pathways. J. Exp. Bot. 2016, 67, 3793–3807. [Google Scholar] [CrossRef] [PubMed]
  21. Apel, K.; Hirt, H. Reactive oxygen species: Metabolism, oxidative stress, and signal transduction. Annu. Rev. Plant Biol. 2004, 55, 373–399. [Google Scholar] [CrossRef] [PubMed]
  22. Hortensteiner, S. Stay-green regulates chlorophyll and chlorophyll-binding protein degradation during senescence. Trends Plant Sci. 2009, 14, 155–162. [Google Scholar] [CrossRef] [PubMed]
  23. Hortensteiner, S. Chlorophyll degradation during senescence. Annu. Rev. Plant Biol. 2006, 57, 55–77. [Google Scholar] [CrossRef] [PubMed]
  24. Hortensteiner, S.; Feller, U. Nitrogen metabolism and remobilization during senescence. J. Exp. Bot. 2002, 53, 927–937. [Google Scholar] [CrossRef] [PubMed]
  25. Ginsburg, S.; Schellenberg, M.; Matile, P. Cleavage of Chlorophyll-Porphyrin—Requirement for Reduced Ferredoxin and Oxygen. Plant Physiol. 1994, 105, 545–554. [Google Scholar] [CrossRef] [PubMed]
  26. Tsuchiya, T.; Ohta, H.; Okawa, K.; Iwamatsu, A.; Shimada, H.; Masuda, T.; Takamiya, K. Cloning of chlorophyllase, the key enzyme in chlorophyll degradation: Finding of a lipase motif and the induction by methyl jasmonate. Proc. Natl. Acad. Sci. USA 1999, 96, 15362–15367. [Google Scholar] [CrossRef] [PubMed]
  27. Schenk, N.; Schelbert, S.; Kanwischer, M.; Goldschmidt, E.E.; Doermann, P.; Hoertensteiner, S. The chlorophyllases AtCLH1 and AtCLH2 are not essential for senescence-related chlorophyll breakdown in Arabidopsis thaliana. FEBS Lett. 2007, 581, 5517–5525. [Google Scholar] [CrossRef] [PubMed]
  28. Schelbert, S.; Aubry, S.; Burla, B.; Agne, B.; Kessler, F.; Krupinska, K.; Hoertensteiner, S. Pheophytin Pheophorbide Hydrolase (Pheophytinase) Is Involved in Chlorophyll Breakdown during Leaf Senescence in Arabidopsis. Plant Cell 2009, 21, 767–785. [Google Scholar] [CrossRef] [PubMed]
  29. Rodoni, S.; Muhlecker, W.; Anderl, M.; Krautler, B.; Moser, D.; Thomas, H.; Matile, P.; Hortensteiner, S. Chlorophyll breakdown in senescent chloroplasts—Cleavage of pheophorbide a in two enzymic steps. Plant Physiol. 1997, 115, 669–676. [Google Scholar] [CrossRef] [PubMed]
  30. Wuthrich, K.L.; Bovet, L.; Hunziker, P.E.; Donnison, I.S.; Hortensteiner, S. Molecular cloning, functional expression and characterisation of RCC reductase involved in chlorophyll catabolism. Plant J. 2000, 21, 189–198. [Google Scholar] [CrossRef] [PubMed]
  31. Pruzinska, A.; Tanner, G.; Anders, I.; Roca, M.; Hortensteiner, S. Chlorophyll breakdown: Pheophorbide a oxygenase is a Rieske-type iron-sulfur protein, encoded by the accelerated cell death 1 gene. Proc. Natl. Acad. Sci. USA 2003, 100, 15259–15264. [Google Scholar] [CrossRef] [PubMed]
  32. Pruzinska, A.; Anders, I.; Aubry, S.; Schenk, N.; Tapernoux-Luthi, E.; Muller, T.; Krautler, B.; Hortensteiner, S. In vivo participation of red chlorophyll catabolite reductase in chlorophyll breakdown. Plant Cell 2007, 19, 369–387. [Google Scholar] [CrossRef] [PubMed]
  33. Hinder, B.; Schellenberg, M.; Rodon, S.; Ginsburg, S.; Vogt, E.; Martinoia, E.; Matile, P.; Hortensteiner, S. How plants dispose of chlorophyll catabolites—Directly energized uptake of tetrapyrrolic breakdown products into isolated vacuoles. J. Biol. Chem. 1996, 271, 27233–27236. [Google Scholar] [CrossRef] [PubMed]
  34. Tommasini, R.; Vogt, E.; Fromenteau, M.; Hortensteiner, S.; Matile, P.; Amrhein, N.; Martinoia, E. An ABC-transporter of Arabidopsis thaliana has both glutathione-conjugate and chlorophyll catabolite transport activity. Plant J. 1998, 13, 773–780. [Google Scholar] [CrossRef] [PubMed]
  35. Kusaba, M.; Ito, H.; Morita, R.; Iida, S.; Sato, Y.; Fujimoto, M.; Kawasaki, S.; Tanaka, R.; Hirochika, H.; Nishimura, M.; et al. Rice NON-YELLOW COLORING1 is involved in light-harvesting complex II and grana degradation during leaf senescence. Plant Cell 2007, 19, 1362–1375. [Google Scholar] [CrossRef] [PubMed]
  36. Sato, Y.; Morita, R.; Katsuma, S.; Nishimura, M.; Tanaka, A.; Kusaba, M. Two short-chain dehydrogenase/reductases, NON-YELLOW COLORING 1 and NYC1-LIKE, are required for chlorophyll b and light-harvesting complex II degradation during senescence in rice. Plant J. 2009, 57, 120–131. [Google Scholar] [CrossRef] [PubMed]
  37. Pruzinska, A.; Tanner, G.; Aubry, S.; Anders, I.; Moser, S.; Muller, T.; Ongania, K.H.; Krautler, B.; Youn, J.Y.; Liljegren, S.J.; et al. Chlorophyll breakdown in senescent Arabidopsis leaves. Characterization of chlorophyll catabolites and of chlorophyll catabolic enzymes involved in the degreening reaction. Plant Physiol. 2005, 139, 52–63. [Google Scholar] [CrossRef] [PubMed]
  38. Tanaka, R.; Hirashima, M.; Satoh, S.; Tanaka, A. The Arabidopsis-accelerated cell death gene ACD1 is involved in oxygenation of pheophorbide a: Inhibition of the pheophorbide a oxygenase activity does not lead to the “Stay-Green” phenotype in Arabidopsis. Plant Cell Physiol. 2003, 44, 1266–1274. [Google Scholar] [CrossRef] [PubMed]
  39. Thomas, H. Sid—A Mendelian Locus Controlling Thylakoid Membrane Disassembly in Senescing Leaves of Festuca-Pratensis. Theor. Appl. Genet. 1987, 73, 551–555. [Google Scholar] [CrossRef] [PubMed]
  40. Armstead, I.; Donnison, I.; Aubry, S.; Harper, J.; Hoertensteiner, S.; James, C.; Mani, J.; Moffet, M.; Ougham, H.; Roberts, L.; et al. From crop to model to crop: Identifying the genetic basis of the staygreen mutation in the Lolium/Festuca forage and amenity grasses. New Phytol. 2006, 172, 592–597. [Google Scholar] [CrossRef] [PubMed]
  41. Armstead, I.; Donnison, I.; Aubry, S.; Harper, J.; Hortensteiner, S.; James, C.; Mani, J.; Moffet, M.; Ougham, H.; Roberts, L.; et al. Cross-species identification of Mendel’s/locus. Science 2007, 315, 73. [Google Scholar] [CrossRef] [PubMed]
  42. Ren, G.; An, K.; Liao, Y.; Zhou, X.; Cao, Y.; Zhao, H.; Ge, X.; Kuai, B. Identification of a novel chloroplast protein AtNYE1 regulating chlorophyll degradation during leaf senescence in Arabidopsis. Plant Physiol. 2007, 144, 1429–1441. [Google Scholar] [CrossRef] [PubMed]
  43. Jiang, H.; Li, M.; Liang, N.; Yan, H.; Wei, Y.; Xu, X.; Liu, J.; Xu, Z.; Chen, F.; Wu, G. Molecular cloning and function analysis of the stay green gene in rice. Plant J. 2007, 52, 197–209. [Google Scholar] [CrossRef] [PubMed]
  44. Park, S.-Y.; Yu, J.-W.; Park, J.-S.; Li, J.; Yoo, S.-C.; Lee, N.-Y.; Lee, S.-K.; Jeong, S.-W.; Seo, H.S.; Koh, H.-J.; et al. The senescence-induced staygreen protein regulates chlorophyll degradation. Plant Cell 2007, 19, 1649–1664. [Google Scholar] [CrossRef] [PubMed]
  45. Sato, Y.; Morita, R.; Nishimura, M.; Yamaguchi, H.; Kusaba, M. Mendel’s green cotyledon gene encodes a positive regulator of the chlorophyll-degrading pathway. Proc. Natl. Acad. Sci. USA 2007, 104, 14169–14174. [Google Scholar] [CrossRef] [PubMed]
  46. Aubry, S.; Mani, J.; Hortensteiner, S. Stay-green protein, defective in Mendel’s green cotyledon mutant, acts independent and upstream of pheophorbide a oxygenase in the chlorophyll catabolic pathway. Plant Mol. Biol. 2008, 67, 243–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pogson, B.J.; Woo, N.S.; Foerster, B.; Small, I.D. Plastid signalling to the nucleus and beyond. Trends Plant Sci. 2008, 13, 602–609. [Google Scholar] [CrossRef] [PubMed]
  48. Borovsky, Y.; Paran, I. Chlorophyll breakdown during pepper fruit ripening in the chlorophyll retainer mutation is impaired at the homolog of the senescence-inducible stay-green gene. Theor. Appl. Genet. 2008, 117, 235–240. [Google Scholar] [CrossRef] [PubMed]
  49. Barry, C.S. The stay-green revolution: Recent progress in deciphering the mechanisms of chlorophyll degradation in higher plants. Plant Sci. 2009, 176, 325–333. [Google Scholar] [CrossRef]
  50. Hoertensteiner, S. Update on the biochemistry of chlorophyll breakdown. Plant Mol. Biol. 2013, 82, 505–517. [Google Scholar] [CrossRef] [PubMed]
  51. Lim, P.O.; Kim, H.J.; Nam, H.G. Leaf senescence. Annu. Rev. Plant Biol. 2007, 58, 115–136. [Google Scholar] [CrossRef] [PubMed]
  52. Rossi, S.; Burgess, P.; Jespersen, D.; Huang, B. Heat-Induced Leaf Senescence Associated with Chlorophyll Metabolism in Bentgrass Lines Differing in Heat Tolerance. Crop Sci. 2017, 57, S169–S178. [Google Scholar] [CrossRef]
  53. Todorov, D.T.; Karanov, E.N.; Smith, A.R.; Hall, M.A. Chlorophyllase activity and chlorophyll content in wild type and eti 5 mutant of Arabidopsis thaliana subjected to low and high temperatures. Biol. Plant. 2003, 46, 633–636. [Google Scholar] [CrossRef]
  54. Djanaguiraman, M.; Prasad, P.V.V.; Boyle, D.L.; Schapaugh, W.T. High-Temperature Stress and Soybean Leaves: Leaf Anatomy and Photosynthesis. Crop Sci. 2011, 51, 2125–2131. [Google Scholar] [CrossRef]
  55. Djanaguiraman, M.; Prasad, P.V.V.; Murugan, M.; Perumal, R.; Reddy, U.K. Physiological differences among sorghum (Sorghum bicolor L. Moench) genotypes under high temperature stress. Environ. Exp. Bot. 2014, 100, 43–54. [Google Scholar] [CrossRef]
  56. Ristic, Z.; Momcilovic, I.; Fu, J.M.; Callegaric, E.; DeRidder, B.P. Chloroplast protein synthesis elongation factor, EF-Tu, reduces thermal aggregation of Rubisco activase. J. Plant Physiol. 2007, 164, 1564–1571. [Google Scholar] [CrossRef] [PubMed]
  57. Akter, N.; Islam, M.R. Heat stress effects and management in wheat. A review. Agron. Sustain. Dev. 2017, 37, 37. [Google Scholar] [CrossRef]
  58. Liu, X.Z.; Huang, B.R. Heat stress injury in relation to membrane lipid peroxidation in creeping bentgrass. Crop Sci. 2000, 40, 503–510. [Google Scholar] [CrossRef]
  59. Jespersen, D.; Zhang, J.; Huang, B. Chlorophyll loss associated with heat-induced senescence in bentgrass. Plant Sci. 2016, 249, 1–12. [Google Scholar] [CrossRef] [PubMed]
  60. Lin, Y.-P.; Wu, M.-C.; Charng, Y.-Y. Identification of a Chlorophyll Dephytylase Involved in Chlorophyll Turnover in Arabidopsis. Plant Cell 2016, 28, 2974–2990. [Google Scholar] [CrossRef] [PubMed]
  61. Veerasamy, M.; He, Y.; Huang, B. Leaf senescence and protein metabolism in creeping bentgrass exposed to heat stress and treated with cytokinins. J. Am. Soc. Hortic. Sci. 2007, 132, 467–472. [Google Scholar]
  62. Chen, Y.; Cothren, J.T.; Chen, D.-H.; Ibrahim, A.M.H.; Lombardini, L. Ethylene-inhibiting compound 1-MCP delays leaf senescence in cotton plants under abiotic stress conditions. J. Integr. Agric. 2015, 14, 1321–1331. [Google Scholar] [CrossRef]
  63. Huberman, M.; Riov, J.; Goldschmidt, E.E.; Apelbaum, A.; Goren, R. The novel ethylene antagonist, 3-cyclopropyl-1-enyl-propanoic acid sodium salt (CPAS), increases grain yield in wheat by delaying leaf senescence. Plant Growth Regul. 2014, 73, 249–255. [Google Scholar] [CrossRef]
  64. Abdelrahman, M.; El-Sayed, M.; Jogaiah, S.; Burritt, D.J.; Lam-Son Phan, T. The “STAY-GREEN” trait and phytohormone signaling networks in plants under heat stress. Plant Cell Rep. 2017, 36, 1009–1025. [Google Scholar] [CrossRef] [PubMed]
  65. Suzuky Pinto, R.; Lopes, M.S.; Collins, N.C.; Reynolds, M.P. Modelling and genetic dissection of staygreen under heat stress. Theor. Appl. Genet. 2016, 129, 2055–2074. [Google Scholar] [CrossRef] [PubMed]
  66. Teixeira, R.N.; Ligterink, W.; Franca-Neto, J.D.B.; Hilhorst, H.W.M.; da Silva, E.A.A. Gene expression profiling of the green seed problem in Soybean. BMC Plant Biol. 2016, 16, 37. [Google Scholar] [CrossRef] [PubMed]
  67. Vijayalakshmi, K.; Fritz, A.K.; Paulsen, G.M.; Bai, G.; Pandravada, S.; Gill, B.S. Modeling and mapping QTL for senescence-related traits in winter wheat under high temperature. Mol. Breed. 2010, 26, 163–175. [Google Scholar] [CrossRef]
  68. Chen, Y.-E.; Zhang, C.-M.; Su, Y.-Q.; Ma, J.; Zhang, Z.-W.; Yuan, M.; Zhang, H.-Y.; Yuan, S. Responses of photosystem II and antioxidative systems to high light and high temperature co-stress in wheat. Environ. Exp. Bot. 2017, 135, 45–55. [Google Scholar] [CrossRef]
  69. Shirdelmoghanloo, H.; Cozzolino, D.; Lohraseb, I.; Collins, N.C. Truncation of grain filling in wheat (Triticum aestivum) triggered by brief heat stress during early grain filling: Association with senescence responses and reductions in stem reserves. Funct. Plant Biol. 2016, 43, 919–930. [Google Scholar] [CrossRef]
  70. Shirdelmoghanloo, H.; Lohraseb, I.; Rabie, H.S.; Brien, C.; Parent, B.; Collins, N.C. Heat susceptibility of grain filling in wheat (Triticum aestivum L.) linked with rapid chlorophyll loss during a 3-day heat treatment. Acta Physiol. Plant. 2016, 38, 208. [Google Scholar] [CrossRef]
  71. Asada, K. Production and scavenging of reactive oxygen species in chloroplasts and their functions. Plant Physiol. 2006, 141, 391–396. [Google Scholar] [CrossRef] [PubMed]
  72. Suzuki, N.; Koussevitzky, S.; Mittler, R.; Miller, G. ROS and redox signalling in the response of plants to abiotic stress. Plant Cell Environ. 2012, 35, 259–270. [Google Scholar] [CrossRef] [PubMed]
  73. Pospisil, P.; Prasad, A. Formation of singlet oxygen and protection against its oxidative damage in Photosystem II under abiotic stress. J. Photochem. Photobiol. B Biol. 2014, 137, 39–48. [Google Scholar] [CrossRef] [PubMed]
  74. Mittler, R.; Vanderauwera, S.; Gollery, M.; Van Breusegem, F. Reactive oxygen gene network of plants. Trends Plant Sci. 2004, 9, 490–498. [Google Scholar] [CrossRef] [PubMed]
  75. Foyer, C.H.; LopezDelgado, H.; Dat, J.F.; Scott, I.M. Hydrogen peroxide- and glutathione-associated mechanisms of acclimatory stress tolerance and signalling. Physiol. Plant. 1997, 100, 241–254. [Google Scholar] [CrossRef]
  76. Dat, J.F.; Lopez-Delgado, H.; Foyer, C.H.; Scott, I.M. Parallel changes in H2O2 and catalase during thermotolerance induced by salicylic acid or heat acclimation in mustard seedlings. Plant Physiol. 1998, 116, 1351–1357. [Google Scholar] [CrossRef] [PubMed]
  77. Vallelian-Bindschedler, L.; Schweizer, P.; Mosinger, E.; Metraux, J.P. Heat-induced resistance in barley to powdery mildew (Blumeria graminis f.sp. hordei) is associated with a burst of active oxygen species. Physiol. Mol. Plant Pathol. 1998, 52, 185–199. [Google Scholar] [CrossRef]
  78. Foyer, C.H.; Noctor, G. Redox Signaling in Plants. Antioxid. Redox Signal. 2013, 18, 2087–2090. [Google Scholar] [CrossRef] [PubMed]
  79. Vacca, R.A.; Valenti, D.; Bobba, A.; Merafina, R.S.; Passarella, S.; Marra, E. Cytochrome c is released in a reactive oxygen species-dependent manner and is degraded via caspase-like proteases in tobacco bright-yellow 2 cells en route to heat shock-induced cell death. Plant Physiol. 2006, 141, 208–219. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, P.; Duan, W.; Takabayashi, A.; Endo, T.; Shikanai, T.; Ye, J.Y.; Mi, H.L. Chloroplastic NAD(P)H dehydrogenase in tobacco leaves functions in alleviation of oxidative damage caused by temperature stress. Plant Physiol. 2006, 141, 465–474. [Google Scholar] [CrossRef] [PubMed]
  81. Yamashita, A.; Nijo, N.; Pospisil, P.; Morita, N.; Takenaka, D.; Aminaka, R.; Yamamoto, Y.; Yamamoto, Y. Quality control of photosystem II—Reactive oxygen species are responsible for the damage to photosystem II under moderate heat stress. J. Biol. Chem. 2008, 283, 28380–28391. [Google Scholar] [CrossRef] [PubMed]
  82. Lin, Y.-P.; Lee, T.-Y.; Tanaka, A.; Charng, Y.-Y. Analysis of an Arabidopsis heat-sensitive mutant reveals that chlorophyll synthase is involved in reutilization of chlorophyllide during chlorophyll turnover. Plant J. 2014, 80, 14–26. [Google Scholar] [CrossRef] [PubMed]
  83. Edreva, A. Generation and scavenging of reactive oxygen species in chloroplasts: A submolecular approach. Agric. Ecosyst. Environ. 2005, 106, 119–133. [Google Scholar] [CrossRef]
  84. Foyer, C.H.; Neukermans, J.; Queval, G.; Noctor, G.; Harbinson, J. Photosynthetic control of electron transport and the regulation of gene expression. J. Exp. Bot. 2012, 63, 1637–1661. [Google Scholar] [CrossRef] [PubMed]
  85. Foyer, C.H.; Noctor, G. Oxidant and antioxidant signalling in plants: A re-evaluation of the concept of oxidative stress in a physiological context. Plant Cell Environ. 2005, 28, 1056–1071. [Google Scholar] [CrossRef]
  86. Foyer, C.H.; Noctor, G. Redox homeostasis and antioxidant signaling: A metabolic interface between stress perception and physiological responses. Plant Cell 2005, 17, 1866–1875. [Google Scholar] [CrossRef] [PubMed]
  87. Foyer, C.H.; Noctor, G. Redox Regulation in Photosynthetic Organisms: Signaling, Acclimation, and Practical Implications. Antioxid. Redox Signal. 2009, 11, 861–905. [Google Scholar] [CrossRef] [PubMed]
  88. Foyer, C.H.; Noctor, G. Ascorbate and Glutathione: The Heart of the Redox Hub. Plant Physiol. 2011, 155, 2–18. [Google Scholar] [CrossRef] [PubMed]
  89. Foyer, C.H.; Noctor, G. Managing the cellular redox hub in photosynthetic organisms. Plant Cell Environ. 2012, 35, 199–201. [Google Scholar] [CrossRef] [PubMed]
  90. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef] [PubMed]
  91. Sainz, M.; Diaz, P.; Monza, J.; Borsani, O. Heat stress results in loss of chloroplast Cu/Zn superoxide dismutase and increased damage to Photosystem II in combined drought-heat stressed Lotus japonicus. Physiol. Plant. 2010, 140, 46–56. [Google Scholar] [CrossRef] [PubMed]
  92. Martins, M.Q.; Rodrigues, W.P.; Fortunato, A.S.; Leitao, A.E.; Rodrigues, A.P.; Pais, I.P.; Martins, L.D.; Silva, M.J.; Reboredo, F.H.; Partelli, F.L.; et al. Protective Response Mechanisms to Heat Stress in Interaction with High [CO2] Conditions in Coffea spp. Front. Plant Sci. 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  93. Mostofa, M.G.; Yoshida, N.; Fujita, M. Spermidine pretreatment enhances heat tolerance in rice seedlings through modulating antioxidative and glyoxalase systems. Plant Growth Regul. 2014, 73, 31–44. [Google Scholar] [CrossRef]
  94. Wang, X.; Cai, J.; Liu, F.; Dai, T.; Cao, W.; Wollenweber, B.; Jiang, D. Multiple heat priming enhances thermo-tolerance to a later high temperature stress via improving subcellular antioxidant activities in wheat seedlings. Plant Physiol. Biochem. 2014, 74, 185–192. [Google Scholar] [CrossRef] [PubMed]
  95. Du, H.; Zhou, P.; Huang, B. Antioxidant enzymatic activities and gene expression associated with heat tolerance in a cool-season perennial grass species. Environ. Exp. Bot. 2013, 87, 159–166. [Google Scholar] [CrossRef]
  96. Kong, F.; Deng, Y.; Wang, G.; Wang, J.; Liang, X.; Meng, Q. LeCDJ1, a chloroplast DnaJ protein, facilitates heat tolerance in transgenic tomatoes. J. Integr. Plant Biol. 2014, 56, 63–74. [Google Scholar] [CrossRef] [PubMed]
  97. Zou, M.; Yuan, L.; Zhu, S.; Liu, S.; Ge, J.; Wang, C. Effects of heat stress on photosynthetic characteristics and chloroplast ultrastructure of a heat-sensitive and heat-tolerant cultivar of wucai (Brassica campestris L.). Acta Physiol. Plant. 2017, 39. [Google Scholar] [CrossRef]
  98. Ara, N.; Nakkanong, K.; Lv, W.; Yang, J.; Hu, Z.; Zhang, M. Antioxidant Enzymatic Activities and Gene Expression Associated with Heat Tolerance in the Stems and Roots of Two Cucurbit Species (“Cucurbita maxima” and “Cucurbita moschata”) and Their Interspecific Inbred Line “Maxchata”. Int. J. Mol. Sci. 2013, 14, 24008–24028. [Google Scholar] [CrossRef] [PubMed]
  99. Miller, G.; Shulaev, V.; Mittler, R. Reactive oxygen signaling and abiotic stress. Physiol. Plant. 2008, 133, 481–489. [Google Scholar] [CrossRef] [PubMed]
  100. Mittler, R.; Vanderauwera, S.; Suzuki, N.; Miller, G.; Tognetti, V.B.; Vandepoele, K.; Gollery, M.; Shulaev, V.; Van Breusegem, F. ROS signaling: The new wave? Trends Plant Sci. 2011, 16, 300–309. [Google Scholar] [CrossRef] [PubMed]
  101. Xiao, Y.; Wang, J.; Dehesh, K. Review of stress specific organelles-to-nucleus metabolic signal molecules in plants. Plant Sci. 2013, 212, 102–107. [Google Scholar] [CrossRef] [PubMed]
  102. Komayama, K.; Khatoon, M.; Takenaka, D.; Horie, J.; Yamashita, A.; Yoshioka, M.; Nakayama, Y.; Yoshida, M.; Ohira, S.; Morita, N.; et al. Quality control of photosystem II: Cleavage and aggregation of heat-damaged D1 protein in spinach thylakoids. Biochim. Biophys. Acta Bioenerg. 2007, 1767, 838–846. [Google Scholar] [CrossRef] [PubMed]
  103. Yoshioka, M.; Uchida, S.; Mori, H.; Komayama, K.; Ohira, S.; Morita, N.; Nakanishi, T.; Yamamoto, Y. Quality control of photosystem II—Cleavage of reaction center D1 protein in spinach thylakoids by FtsH protease under moderate heat stress. J. Biol. Chem. 2006, 281, 21660–21669. [Google Scholar] [CrossRef] [PubMed]
  104. Khatoon, M.; Inagawa, K.; Pospisil, P.; Yamashita, A.; Yoshioka, M.; Lundin, B.; Horie, J.; Morita, N.; Jajoo, A.; Yamamoto, Y.; et al. Quality Control of Photosystem II thylakoid unstacking is necessary to avoid further damage to the D1 protein and to facilitate D1 degradation under light stress in spinach thylakoids. J. Biol. Chem. 2009, 284, 25343–25352. [Google Scholar] [CrossRef] [PubMed]
  105. Jarvi, S.; Suorsa, M.; Aro, E.M. Photosystem II repair in plant chloroplasts—Regulation, assisting proteins and shared components with photosystem II biogenesis. Biochim. Biophys. Acta Bioenerg. 2015, 1847, 900–909. [Google Scholar] [CrossRef] [PubMed]
  106. Vass, I.; Cser, K. Janus-faced charge recombinations in photosystem II photoinhibition. Trends Plant Sci. 2009, 14, 200–205. [Google Scholar] [CrossRef] [PubMed]
  107. Telfer, A.; Dhami, S.; Bishop, S.M.; Phillips, D.; Barber, J. Beta-Carotene Quenches Singlet Oxygen Formed by Isolated Photosystem-II Reaction Centers. Biochemistry 1994, 33, 14469–14474. [Google Scholar] [CrossRef] [PubMed]
  108. Kruk, J.; Trebst, A. Plastoquinol as a singlet oxygen scavenger in photosystem II. Biochim. Biophys. Acta Bioenerg. 2008, 1777, 154–162. [Google Scholar]
  109. Weisz, D.A.; Gross, M.L.; Pakrasi, H.B. Reactive oxygen species leave a damage trail that reveals water channels in Photosystem II. Sci. Adv. 2017, 3. [Google Scholar] [CrossRef] [PubMed]
  110. Pospisil, P.; Yamamoto, Y. Damage to photosystem II by lipid peroxidation products. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 457–466. [Google Scholar] [CrossRef] [PubMed]
  111. Park, Y.I.; Chow, W.S.; Anderson, J.M. Light Inactivation of Functional Photosystem-Ii in Leaves of Peas Grown in Moderate Light Depends on Photon Exposure. Planta 1995, 196, 401–411. [Google Scholar] [CrossRef]
  112. Tyystjarvi, E.; Aro, E.M. The rate constant of photoinhibition, measured in lincomycin-treated leaves, is directly proportional to light intensity. Proc. Natl. Acad. Sci. USA 1996, 93, 2213–2218. [Google Scholar] [CrossRef] [PubMed]
  113. Nelson, C.J.; Alexova, R.; Jacoby, R.P.; Millar, A.H. Proteins with High Turnover Rate in Barley Leaves Estimated by Proteome Analysis Combined with in Planta Isotope Labeling. Plant Physiol. 2014, 166, 91–108. [Google Scholar] [CrossRef] [PubMed]
  114. Bergantino, E.; Brunetta, A.; Touloupakis, E.; Segalla, A.; Szabo, I.; Giacometti, G.M. Role of the PSII-H subunit in photoprotection—Novel aspects of D1 turnover in Synechocystis 6803. J. Biol. Chem. 2003, 278, 41820–41829. [Google Scholar] [CrossRef] [PubMed]
  115. Rokka, A.; Suorsa, M.; Saleem, A.; Battchikova, N.; Aro, E.M. Synthesis and assembly of thylakoid protein complexes: Multiple assembly steps of photosystem II. Biochem. J. 2005, 388, 159–168. [Google Scholar] [CrossRef] [PubMed]
  116. Kato, Y.; Sakamoto, W. Protein Quality Control in Chloroplasts: A Current Model of D1 Protein Degradation in the Photosystem II Repair Cycle. J. Biochem. 2009, 146, 463–469. [Google Scholar] [CrossRef] [PubMed]
  117. Barber, J.; Andersson, B. Too Much of a Good Thing—Light Can Be Bad for Photosynthesis. Trends Biochem. Sci. 1992, 17, 61–66. [Google Scholar] [CrossRef]
  118. Aro, E.M.; Virgin, I.; Andersson, B. Photoinhibition of Photosystem-2—Inactivation, Protein Damage and Turnover. Biochim. Biophys. Acta 1993, 1143, 113–134. [Google Scholar] [CrossRef]
  119. Adir, N.; Zer, H.; Shochat, S.; Ohad, I. Photoinhibition—A historical perspective. Photosynth. Res. 2003, 76, 343–370. [Google Scholar] [CrossRef] [PubMed]
  120. Aro, E.M.; Suorsa, M.; Rokka, A.; Allahverdiyeva, Y.; Paakkarinen, V.; Saleem, A.; Battchikova, N.; Rintamaki, E. Dynamics of photosystem II: A proteomic approach to thylakoid protein complexes. J. Exp. Bot. 2005, 56, 347–356. [Google Scholar] [CrossRef] [PubMed]
  121. Baena-Gonzalez, E.; Aro, E.M. Biogenesis, assembly and turnover of photosystem II units. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 2002, 357, 1451–1459. [Google Scholar] [CrossRef] [PubMed]
  122. Cheregi, O.; Wagner, R.; Funk, C. Insights into the Cyanobacterial Deg/HtrA Proteases. Front. Plant Sci. 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  123. Huesgen, P.F.; Schuhmann, H.; Adamska, I. Deg/HtrA proteases as components of a network for photosystem II quality control in chloroplasts and cyanobacteria. Res. Microbiol. 2009, 160, 726–732. [Google Scholar] [CrossRef] [PubMed]
  124. Schuhmann, H.; Huesgen, P.F.; Adamska, I. The family of Deg/HtrA proteases in plants. BMC Plant Biol. 2012, 12. [Google Scholar] [CrossRef] [PubMed]
  125. Schuhmann, H.; Adamska, I. Deg proteases and their role in protein quality control and processing in different subcellular compartments of the plant cell. Physiol. Plant. 2012, 145, 224–234. [Google Scholar] [CrossRef] [PubMed]
  126. Sun, X.; Wang, L.; Zhang, L. Involvement of DEG5 and DEG8 proteases in the turnover of the photosystem II reaction center D1 protein under heat stress in Arabidopsis thaliana. Chin. Sci. Bull. 2007, 52, 1742–1745. [Google Scholar] [CrossRef]
  127. Kato, Y.; Sun, X.; Zhang, L.; Sakamoto, W. Cooperative D1 Degradation in the Photosystem II Repair Mediated by Chloroplastic Proteases in Arabidopsis. Plant Physiol. 2012, 159, 1428–1439. [Google Scholar] [CrossRef] [PubMed]
  128. Chen, J.P.; Burke, J.J.; Velten, J.; Xin, Z.U. FtsH11 protease plays a critical role in Arabidopsis thermotolerance. Plant J. 2006, 48, 73–84. [Google Scholar] [CrossRef] [PubMed]
  129. Yoshioka, M.; Yamamoto, Y. Quality control of Photosystem II: Where and how does the degradation of the D1 protein by FtsH proteases start under light stress?—Facts and hypotheses. J. Photochem. Photobiol. B Biol. 2011, 104, 229–235. [Google Scholar] [CrossRef] [PubMed]
  130. Peltier, J.B.; Emanuelsson, O.; Kalume, D.E.; Ytterberg, J.; Friso, G.; Rudella, A.; Liberles, D.A.; Soderberg, L.; Roepstorff, P.; von Heijne, G.; et al. Central functions of the lumenal and peripheral thylakoid proteome of Arabidopsis determined by experimentation and genome-wide prediction. Plant Cell 2002, 14, 211–236. [Google Scholar] [CrossRef] [PubMed]
  131. Schubert, M.; Petersson, U.A.; Haas, B.J.; Funk, C.; Schroder, W.P.; Kieselbach, T. Proteome map of the chloroplast lumen of Arabidopsis thaliana. J. Biol. Chem. 2002, 277, 8354–8365. [Google Scholar] [CrossRef] [PubMed]
  132. Kapri-Pardes, E.; Naveh, L.; Adam, Z. The thylakoid lumen protease Deg1 is involved in the repair of photosystem II from photoinhibition in Arabidopsis. Plant Cell 2007, 19, 1039–1047. [Google Scholar] [CrossRef] [PubMed]
  133. Sun, X.; Peng, L.; Guo, J.; Chi, W.; Ma, J.; Lu, C.; Zhang, L. Formation of DEG5 and DEG8 complexes and their involvement in the degradation of photodamaged photosystem II reaction center D1 protein in Arabidopsis. Plant Cell 2007, 19, 1347–1361. [Google Scholar] [CrossRef] [PubMed]
  134. Kley, J.; Schmidt, B.; Boyanov, B.; Stolt-Bergner, P.C.; Kirk, R.; Ehrmann, M.; Knopf, R.R.; Naveh, L.; Adam, Z.; Clausen, T. Structural adaptation of the plant protease Deg1 to repair photosystem II during light exposure. Nat. Struct. Mol. Biol. 2011, 18, 728–731. [Google Scholar] [CrossRef] [PubMed]
  135. Sun, X.; Ouyang, M.; Guo, J.; Ma, J.; Lu, C.; Adam, Z.; Zhang, L. The thylakoid protease Deg1 is involved in photosystem-II assembly in Arabidopsis thaliana. Plant J. 2010, 62, 240–249. [Google Scholar] [CrossRef] [PubMed]
  136. Chassin, Y.; Kapri-Pardes, E.; Sinvany, G.; Arad, T.; Adam, Z. Expression and characterization of the thylakoid lumen protease DegP1 from Arabidopsis. Plant Physiol. 2002, 130, 857–864. [Google Scholar] [CrossRef] [PubMed]
  137. Aro, E.M.; McCaffery, S.; Anderson, J.M. Photoinhibition and D1 Protein-Degradation in Peas Acclimated to Different Growth Irradiances. Plant Physiol. 1993, 103, 835–843. [Google Scholar] [CrossRef] [PubMed]
  138. Tyedmers, J.; Mogk, A.; Bukau, B. Cellular strategies for controlling protein aggregation. Nat. Rev. Mol. Cell Biol. 2010, 11, 777–788. [Google Scholar] [CrossRef] [PubMed]
  139. Vierling, E. The Roles of Heat-Shock Proteins in Plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1991, 42, 579–620. [Google Scholar] [CrossRef]
  140. Chen, Q.; Vierling, E. Analysis of Conserved Domains Identifies a Unique Structural Feature of a Chloroplast Heat-Shock Protein. Mol. Gen. Genet. 1991, 226, 425–431. [Google Scholar] [CrossRef] [PubMed]
  141. Chen, Q.; Lauzon, L.M.; Derocher, A.E.; Vierling, E. Accumulation, Stability, and Localization of a Major Chloroplast Heat-Shock Protein. J. Cell Biol. 1990, 110, 1873–1883. [Google Scholar] [CrossRef] [PubMed]
  142. Vierling, E.; Harris, L.M.; Chen, Q. The Major Low-Molecular-Weight Heat-Shock Protein in Chloroplasts Shows Antigenic Conservation among Diverse Higher-Plant Species. Mol. Cell. Biol. 1989, 9, 461–468. [Google Scholar] [CrossRef] [PubMed]
  143. Heckathorn, S.A.; Downs, C.A.; Sharkey, T.D.; Coleman, J.S. The small, methionine-rich chloroplast heat-shock protein protects photosystem II electron transport during heat stress. Plant Physiol. 1998, 116, 439–444. [Google Scholar] [CrossRef] [PubMed]
  144. Heckathorn, S.A.; Ryan, S.L.; Baylis, J.A.; Wang, D.F.; Hamilton, E.W.; Cundiff, L.; Luthe, D.S. In vivo evidence from an Agrostis stolonifera selection genotype that chloroplast small heat-shock proteins can protect photosystem II during heat stress. Funct. Plant Biol. 2002, 29, 933–944. [Google Scholar] [CrossRef]
  145. Kim, K.-H.; Alam, I.; Kim, Y.-G.; Sharmin, S.A.; Lee, K.-W.; Lee, S.-H.; Lee, B.-H. Overexpression of a chloroplast-localized small heat shock protein OsHSP26 confers enhanced tolerance against oxidative and heat stresses in tall fescue. Biotechnol. Lett. 2012, 34, 371–377. [Google Scholar] [CrossRef] [PubMed]
  146. Shakeel, S.; Ul Haq, N.; Heckathorn, S.A.; Hamilton, E.W.; Luthe, D.S. Ecotypic variation in chloroplast small heat-shock proteins and related thermotolerance in Chenopodium album. Plant Physiol. Biochem. 2011, 49, 898–908. [Google Scholar] [CrossRef] [PubMed]
  147. Wang, D.F.; Luthe, D.S. Heat sensitivity in a bentgrass variant. Failure to accumulate a chloroplast heat shock protein isoform implicated in heat tolerance. Plant Physiol. 2003, 133, 319–327. [Google Scholar] [CrossRef] [PubMed]
  148. Harndahl, U.; Hall, R.B.; Osteryoung, K.W.; Vierling, E.; Bornman, J.F.; Sundby, C. The chloroplast small heat shock protein undergoes oxidation-dependent conformational changes and may protect plants from oxidative stress. Cell Stress Chaperones 1999, 4, 129–138. [Google Scholar] [CrossRef]
  149. Neta-Sharir, I.; Isaacson, T.; Lurie, S.; Weiss, D. Dual role for tomato heat shock protein 21: Protecting photosystem II from oxidative stress and promoting color changes during fruit maturation. Plant Cell 2005, 17, 1829–1838. [Google Scholar] [CrossRef] [PubMed]
  150. Downs, C.A.; Coleman, J.S.; Heckathorn, S.A. The chloroplast 22-Ku heat-shock protein: A lumenal protein that associates with the oxygen evolving complex and protects photosystem II during heat stress. J. Plant Physiol. 1999, 155, 477–487. [Google Scholar] [CrossRef]
  151. Bernfur, K.; Rutsdottir, G.; Emanuelsson, C. The chloroplast-localized small heat shock protein Hsp21 associates with the thylakoid membranes in heat-stressed plants. Protein Sci. Publ. Protein Soc. 2017, 26, 1773–1784. [Google Scholar] [CrossRef] [PubMed]
  152. Spreitzer, R.J.; Salvucci, M.E. Rubisco: Structure, regulatory interactions, and possibilities for a better enzyme. Annu. Rev. Plant Biol. 2002, 53, 449–475. [Google Scholar] [CrossRef] [PubMed]
  153. Perdomo, J.A.; Capo-Bauca, S.; Carmo-Silva, E.; Galmes, J. Rubisco and Rubisco Activase Play an Important Role in the Biochemical Limitations of Photosynthesis in Rice, Wheat, and Maize under High Temperature and Water Deficit. Front. Plant Sci. 2017, 8. [Google Scholar] [CrossRef] [PubMed]
  154. Sage, R.F.; Way, D.A.; Kubien, D.S. Rubisco, Rubisco activase, and global climate change. J. Exp. Bot. 2008, 59, 1581–1595. [Google Scholar] [CrossRef] [PubMed]
  155. Portis, A.R. Rubisco activase—Rubisco’s catalytic chaperone. Photosynth. Res. 2003, 75, 11–27. [Google Scholar] [CrossRef] [PubMed]
  156. Schrader, S.M.; Wise, R.R.; Wacholtz, W.F.; Ort, D.R.; Sharkey, T.D. Thylakoid membrane responses to moderately high leaf temperature in Pima cotton. Plant Cell Environ. 2004, 27, 725–735. [Google Scholar] [CrossRef]
  157. Yamori, W.; Noguchi, K.; Kashino, Y.; Terashima, I. The role of electron transport in determining the temperature dependence of the photosynthetic rate in spinach leaves grown at contrasting temperatures. Plant Cell Physiol. 2008, 49, 583–591. [Google Scholar] [CrossRef] [PubMed]
  158. Carmo-Silva, A.E.; Salvucci, M.E. The activity of Rubisco’s molecular chaperone, Rubisco activase, in leaf extracts. Photosynth. Res. 2011, 108, 143–155. [Google Scholar] [CrossRef] [PubMed]
  159. Salvucci, M.E.; Crafts-Brandner, S.J. Mechanism for deactivation of Rubisco under moderate heat stress. Physiol. Plant. 2004, 122, 513–519. [Google Scholar] [CrossRef]
  160. Robinson, S.P.; Streusand, V.J.; Chatfield, J.M.; Portis, A.R. Purification and Assay of Rubisco activase from Leaves. Plant Physiol. 1988, 88, 1008–1014. [Google Scholar] [CrossRef] [PubMed]
  161. Holbrook, G.P.; Galasinski, S.C.; Salvucci, M.E. Regulation of 2-Carboxyarabinitol 1-Phosphatase. Plant Physiol. 1991, 97, 894–899. [Google Scholar] [CrossRef] [PubMed]
  162. Feller, U.; Crafts-Brandner, S.J.; Salvucci, M.E. Moderately high temperatures inhibit ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) activase-mediated activation of Rubisco. Plant Physiol. 1998, 116, 539–546. [Google Scholar] [CrossRef] [PubMed]
  163. Salvucci, M.E.; Osteryoung, K.W.; Crafts-Brandner, S.J.; Vierling, E. Exceptional sensitivity of Rubisco activase to thermal denaturation in vitro and in vivo. Plant Physiol. 2001, 127, 1053–1064. [Google Scholar] [CrossRef] [PubMed]
  164. Schrader, S.M.; Kleinbeck, K.R.; Sharkey, T.D. Rapid heating of intact leaves reveals initial effects of stromal oxidation on photosynthesis. Plant Cell Environ. 2007, 30, 671–678. [Google Scholar] [CrossRef] [PubMed]
  165. Zhang, R.; Sharkey, T.D. Photosynthetic electron transport and proton flux under moderate heat stress. Photosynth. Res. 2009, 100, 29–43. [Google Scholar] [CrossRef] [PubMed]
  166. Way, D.A.; Yamori, W. Thermal acclimation of photosynthesis: On the importance of adjusting our definitions and accounting for thermal acclimation of respiration. Photosynth. Res. 2014, 119, 89–100. [Google Scholar] [CrossRef] [PubMed]
  167. Yamori, W.; Hikosaka, K.; Way, D.A. Temperature response of photosynthesis in C-3, C-4, and CAM plants: Temperature acclimation and temperature adaptation. Photosynth. Res. 2014, 119, 101–117. [Google Scholar] [CrossRef] [PubMed]
  168. Salvucci, M.E.; Crafts-Brandner, S.J. Inhibition of photosynthesis by heat stress: The activation state of Rubisco as a limiting factor in photosynthesis. Physiol. Plant. 2004, 120, 179–186. [Google Scholar] [CrossRef] [PubMed]
  169. DeRidder, B.P.; Salvucci, M.E. Modulation of Rubisco activase gene expression during heat stress in cotton (Gossypium hirsutum L.) involves post-transcriptional mechanisms. Plant Sci. 2007, 172, 246–254. [Google Scholar] [CrossRef]
  170. Kumar, R.R.; Goswami, S.; Singh, K.; Dubey, K.; Singh, S.; Sharma, R.; Verma, N.; Kala, Y.K.; Rai, G.K.; Grover, M.; et al. Identification of Putative RuBisCo Activase (TaRca1)-The Catalytic Chaperone Regulating Carbon Assimilatory Pathway in Wheat (Triticum aestivum) under the Heat Stress. Front. Plant Sci. 2016, 7, 986. [Google Scholar] [CrossRef] [PubMed]
  171. Kurek, I.; Chang, T.K.; Bertain, S.M.; Madrigal, A.; Liu, L.; Lassner, M.W.; Zhu, G. Enhanced thermostability of Arabidopsis Rubisco activase improves photosynthesis and growth rates under moderate heat stress. Plant Cell 2007, 19, 3230–3241. [Google Scholar] [CrossRef] [PubMed]
  172. Kumar, A.; Li, C.; Portis, A.R., Jr. Arabidopsis thaliana expressing a thermostable chimeric Rubisco activase exhibits enhanced growth and higher rates of photosynthesis at moderately high temperatures. Photosynth. Res. 2009, 100, 143–153. [Google Scholar] [CrossRef] [PubMed]
  173. Ristic, Z.; Momcilovic, I.; Bukovnik, U.; Prasad, P.V.V.; Fu, J.; DeRidder, B.P.; Elthon, T.E.; Mladenov, N. Rubisco activase and wheat productivity under heat-stress conditions. J. Exp. Bot. 2009, 60, 4003–4014. [Google Scholar] [CrossRef] [PubMed]
  174. Shivhare, D.; Mueller-Cajar, O. In Vitro Characterization of Thermostable CAM Rubisco Activase Reveals a Rubisco Interacting Surface Loop. Plant Physiol. 2017, 174, 1505–1516. [Google Scholar] [CrossRef] [PubMed]
  175. Salvucci, M.E. Association of Rubisco activase with chaperonin-60 beta: A possible mechanism for protecting photosynthesis during heat stress. J. Exp. Bot. 2008, 59, 1923–1933. [Google Scholar] [CrossRef] [PubMed]
  176. Scales, J.C.; Parry, M.A.J.; Salvucci, M.E. A non-radioactive method for measuring Rubisco activase activity in the presence of variable ATP: ADP ratios, including modifications for measuring the activity and activation state of Rubisco. Photosynth. Res. 2014, 119, 355–365. [Google Scholar] [CrossRef] [PubMed]
  177. Carmo-Silva, A.E.; Salvucci, M.E. The Regulatory Properties of Rubisco activase Differ among Species and Affect Photosynthetic Induction during Light Transitions. Plant Physiol. 2013, 161, 1645–1655. [Google Scholar] [CrossRef] [PubMed]
  178. Wang, D.; Li, X.-F.; Zhou, Z.-J.; Feng, X.-P.; Yang, W.-J.; Jiang, D.-A. Two Rubisco activase isoforms may play different roles in photosynthetic heat acclimation in the rice plant. Physiol. Plant. 2010, 139, 55–67. [Google Scholar] [CrossRef] [PubMed]
  179. Hozain, M.D.I.; Salvucci, M.E.; Fokar, M.; Holaday, A.S. The differential response of photosynthesis to high temperature for a boreal and temperate Populus species relates to differences in Rubisco activation and Rubisco activase properties. Tree Physiol. 2010, 30, 32–44. [Google Scholar] [CrossRef] [PubMed]
  180. Feller, U. Drought stress and carbon assimilation in a warming climate: Reversible and irreversible impacts. J. Plant Physiol. 2016, 203, 69–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Extensive and transient metabolic reprogramming in chloroplasts under heat stress. Major events of metabolic reprogramming in response to heat stress include chlorophyll breakdown, generation of reactive oxygen species (ROS), antioxidant defense, protein turnover, and metabolic alterations with carbon assimilation. With respect to the systemic acquired acclimation to heat stress in plants, diverse metabolic reprogramming in chloroplasts is required for optimizing plant growth and development during high temperature stresses.
Figure 1. Extensive and transient metabolic reprogramming in chloroplasts under heat stress. Major events of metabolic reprogramming in response to heat stress include chlorophyll breakdown, generation of reactive oxygen species (ROS), antioxidant defense, protein turnover, and metabolic alterations with carbon assimilation. With respect to the systemic acquired acclimation to heat stress in plants, diverse metabolic reprogramming in chloroplasts is required for optimizing plant growth and development during high temperature stresses.
Ijms 19 00849 g001
Figure 2. A representative scheme of reactive oxygen species (ROS) generation and scavenging in chloroplasts under heat stress. High temperature stress triggers oxidative bursts of superoxide and/or hydrogen peroxide in plants. The transfer of excitation energy in the photosystem II (PSII) antenna complex and the electron transport in the PSII reaction center can be inhibited by heat stress. It has been established that ROS are generated both on the electron acceptor and the electron donor side of PSII under heat stress during which electron transport from the manganese complex to plastoquinone (PQ) is limited. The leakage of electrons to molecular oxygen on the electron acceptor side of PSII forms O2•−, inducing initiation of a cascade reaction leading to the formation of H2O2. A diversified ROS-scavenging network functions in concert in chloroplasts, mainly including antioxidants and APX-glutathione cycle, to keep the equilibrium between ROS production and scavenging. The efficient enzymatic scavenging systems are composed of several key enzymes, including superoxide dismutase (SOD), ascorbate peroxidase (APX), glutathione reductase (GR), monodehydroascorbate reductase (MDHAR), dehydroascorbate reductase (DHAR), glutathione peroxidase (GPX) and glutathione-S-transferase (GST) and non-enzymatic systems contain antioxidants such as ascorbic acid (Asc) and glutathione (GSH).
Figure 2. A representative scheme of reactive oxygen species (ROS) generation and scavenging in chloroplasts under heat stress. High temperature stress triggers oxidative bursts of superoxide and/or hydrogen peroxide in plants. The transfer of excitation energy in the photosystem II (PSII) antenna complex and the electron transport in the PSII reaction center can be inhibited by heat stress. It has been established that ROS are generated both on the electron acceptor and the electron donor side of PSII under heat stress during which electron transport from the manganese complex to plastoquinone (PQ) is limited. The leakage of electrons to molecular oxygen on the electron acceptor side of PSII forms O2•−, inducing initiation of a cascade reaction leading to the formation of H2O2. A diversified ROS-scavenging network functions in concert in chloroplasts, mainly including antioxidants and APX-glutathione cycle, to keep the equilibrium between ROS production and scavenging. The efficient enzymatic scavenging systems are composed of several key enzymes, including superoxide dismutase (SOD), ascorbate peroxidase (APX), glutathione reductase (GR), monodehydroascorbate reductase (MDHAR), dehydroascorbate reductase (DHAR), glutathione peroxidase (GPX) and glutathione-S-transferase (GST) and non-enzymatic systems contain antioxidants such as ascorbic acid (Asc) and glutathione (GSH).
Ijms 19 00849 g002

Share and Cite

MDPI and ACS Style

Wang, Q.-L.; Chen, J.-H.; He, N.-Y.; Guo, F.-Q. Metabolic Reprogramming in Chloroplasts under Heat Stress in Plants. Int. J. Mol. Sci. 2018, 19, 849. https://doi.org/10.3390/ijms19030849

AMA Style

Wang Q-L, Chen J-H, He N-Y, Guo F-Q. Metabolic Reprogramming in Chloroplasts under Heat Stress in Plants. International Journal of Molecular Sciences. 2018; 19(3):849. https://doi.org/10.3390/ijms19030849

Chicago/Turabian Style

Wang, Qing-Long, Juan-Hua Chen, Ning-Yu He, and Fang-Qing Guo. 2018. "Metabolic Reprogramming in Chloroplasts under Heat Stress in Plants" International Journal of Molecular Sciences 19, no. 3: 849. https://doi.org/10.3390/ijms19030849

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop