Next Article in Journal
Copper Depletion Strongly Enhances Ferroptosis via Mitochondrial Perturbation and Reduction in Antioxidative Mechanisms
Next Article in Special Issue
Review of the Mechanisms by Which Transcription Factors and Exogenous Substances Regulate ROS Metabolism under Abiotic Stress
Previous Article in Journal
Myeloperoxidase as a Potential Biomarker of Acute-Myocardial-Infarction-Induced Depression and Suppression of the Innate Immune System
Previous Article in Special Issue
The Investigation of the Impact of Toxicity of Metals on Oxygen-Evolving Complex in Spinacia oleracea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Regulation and Evolution of Redox Homeostasis in Photosynthetic Machinery

1
Hubei Collaborative Innovation Center for Grain Industry, College of Agriculture, Yangtze University, Jingzhou 414000, China
2
College of Agriculture and Biotechnology, Zhejiang University, Hangzhou 310058, China
3
Department of Biotechnology, University of Okara, Okara, Punjab 56300, Pakistan
4
School of Science, Western Sydney University, Penrith, NSW 2751, Australia
5
Hawkesbury Institute for the Environment, Western Sydney University, Penrith, NSW 2751, Australia
*
Authors to whom correspondence should be addressed.
Antioxidants 2022, 11(11), 2085; https://doi.org/10.3390/antiox11112085
Submission received: 12 September 2022 / Revised: 14 October 2022 / Accepted: 20 October 2022 / Published: 22 October 2022

Abstract

:
The recent advances in plant biology have significantly improved our understanding of reactive oxygen species (ROS) as signaling molecules in the redox regulation of complex cellular processes. In plants, free radicals and non-radicals are prevalent intra- and inter-cellular ROS, catalyzing complex metabolic processes such as photosynthesis. Photosynthesis homeostasis is maintained by thiol-based systems and antioxidative enzymes, which belong to some of the evolutionarily conserved protein families. The molecular and biological functions of redox regulation in photosynthesis are usually to balance the electron transport chain, photosystem II, photosystem I, mesophyll and bundle sheath signaling, and photo-protection regulating plant growth and productivity. Here, we review the recent progress of ROS signaling in photosynthesis. We present a comprehensive comparative bioinformatic analysis of redox regulation in evolutionary distinct photosynthetic cells. Gene expression, phylogenies, sequence alignments, and 3D protein structures in representative algal and plant species revealed conserved key features including functional domains catalyzing oxidation and reduction reactions. We then discuss the antioxidant-related ROS signaling and important pathways for achieving homeostasis of photosynthesis. Finally, we highlight the importance of plant responses to stress cues and genetic manipulation of disturbed redox status for balanced and enhanced photosynthetic efficiency and plant productivity.

1. Introduction

Oxygenic photosynthesis originated in cyanobacteria and subsequently the evolutionary pressure for higher redox potentials (electron source and oxygen (O2)) resulted in the evolution of reaction centers or photosystems. The findings that O2 evolving reaction center 2 (photosystem II; PSII) was originated from reaction center 1 (photosystem I; PSI) through a series of evolutionary events in algae and plants [1]. Molecular evolutionary events of photosynthesis were largely associated with genome duplication, gene fusion, and splitting, and lateral gene transfer, which drove the metabolic fluxes and photosynthetic components among diverse species of algae and plants [2]. Further, the evolution of aerobic respiration and novel biosynthesis pathways provide insights into the modern biology of complex multicellular organisms [3].
A fraction of captured light energy is utilized in converting water (H2O) and carbon dioxide (CO2) to glucose and O2 as end products [4]. The hypothesis that the first O2 molecules were produced as a by-product of photosynthesis is still debatable. Earlier studies proposed the appearance of O2 molecules as a result of the photo-dissociation of CO2 and H2O by ultraviolet (UV) radiations [5,6]. However, it has universally been accepted that cyanobacteria were the first organisms to evolve photosynthetic machinery on Earth, releasing O2 on a large scale into the atmosphere. Chloroplasts, the core photosynthetic organelle, have evolved with plastids from photosynthetic prokaryotes, precisely, ancestors of cyanobacteria into plants via endosymbiotic evolution more than 1 billion years ago [7]. Subsequently, the continuous oxidation of H2O to O2 has remarkably changed the redox status of the Earth, oceanic, and atmosphere.
Reactive oxygen species (ROS) occurred soon after the first O2 molecules originated on Earth by the ancestors of cyanobacteria about 3.8 billion years ago [8]. Since then, these signaling molecules have been produced constantly through aerobic metabolism. It has also been proposed that ROS initiated from atmospheric oxygen soon after its release from biological systems [9,10]. Green plants (Viridiplantae) and algae produce oxygen radicals and their derivatives during aerobic photosynthesis and respiration [11]. These include free radicals [e.g., superoxide anion (O2•−), hydroxyl radical (OH), hydroperoxyl radical (HO2) and alkoxy radical (RO)] and non-radical, e.g., singlet oxygen (1O2) and hydrogen peroxide [(H2O2)] molecules [12,13]. Early reports showed that ROS are toxic signaling molecules accumulated in plant cells to disturb cellular homeostasis. Overaccumulation of ROS disrupts cell metabolism which may lead to DNA damage, genome instability, and programmed cell death [14,15]. In addition, ROS are known to regulate photorespiration, growth, and stress response in plants [16,17,18]. However, our understanding regarding ROS signaling and molecular functions in cell compartments and organelles, has been limited due to technological difficulties over the last few decades [11,19,20].
As an important metabolic process, photosynthesis regulates ROS production in algae and plants. The excessive ROS as a by-product of oxygenic processes in chloroplast, mitochondria, peroxisomes, and nuclei [11,15,20,21,22,23,24], disturb photosynthetic electron transport (PET), PSII, PSI, as well as photorespiration and gene expression levels [25,26,27,28]. Photosynthetic cells have diverse layers of defense to cope with oxidative stress. The prevailing concepts show that redox signaling is carried out within antioxidative systems at the cellular level [29], and many of those associated proteins are evolutionarily conserved in plants and algae [30]. The multilayered defense systems fine-tune the ROS balance during PET, and excessive ROS diffusion and reactivity are also balanced at the organellar level throughout the life cycle [10,20,30], which is also known as redox homeostasis [31].
Here, we review the current knowledge of ROS homeostasis and oxygenic photosynthesis. We discuss ROS toxicity and signaling in the important organelles of a plant cell. We then analyze the molecular evolution of protein families associated with photosynthesis and redox regulation in plants and algae. In addition, we compare the evolution of ROS signaling in photosynthetic machinery in chloroplasts of distinct C3, C4, Crassulacean acid metabolism (CAM) plants and early divergent lineages of plants and algae. Finally, we highlight a few key mechanisms of how plants are adapted to high light intensities. We thus propose efficient plant biological modification of ROS homeostasis and photosynthesis to achieve better crop productivity for global food security. Also, we direct the reader to some excellent recent reviews that have focused on ROS toxicity, signaling, or photosynthesis [28,32,33,34,35,36,37].

2. ROS Signaling Is a Double-Edged Sword in Plant Photosynthesis

ROS affects almost all aspects of plant life. ROS toxicity in different species is commonly regarded as oxidative stress, potentially causing a level of damage that led to intracellular and intercellular lethality. The O2, OH, H2O2 and 1O2 reactivity may vary across or within cells, due to the site of generation, and the nature of the biomolecules. Among them, O2 and 1O2 have longer half-lives ranging from 1–4 μs with a migration distance of 30 nm in mitochondria, chloroplast, and nuclei, conversely, H2O2 is closer to 1 ms with a migration distance of 1 μm reacting with DNA and sensitive cysteine (Cys) and methionine (Met) residues. Additionally, OH’s half-life is approximately 1 ns with a migration distance of 1 nm in the Fenton reaction [37]. Furthermore, OH’s very short migration distance allows this ROS molecule to be extremely reactive with DNA, RNA, lipids, and proteins. The expression levels of ROS also vary across the subcellular organelles suggesting that the regulation of ROS is dynamic and that compartmentalization is utilized to reduce ROS lethality (Table S1). ROS signaling can be mediated by mitochondrial nicotinamide adenine dinucleotide phosphate (NADPH) oxidases, specifically, the respiratory burst oxidase homologs (RBOHs) induce ROS production in the apoplast [38,39] (Figure 1). The redox signaling complex largely consists of RBOHs, superoxide dismutases (SODs; different metal cofactors), catalases (CATs), peroxidase (PODs), glutathione peroxidase (GPXs), iron-dependent mechanisms and a network of thio- and glutaredoxins [40,41,42]. Knockout or knockdown of any of them results in modified redox signaling in plants [43].
In general, some physiological functions of plants are regulated via the interaction of ROS with Cys and Met residues of key proteins [44,45]. The oxidative posttranslational modifications (oxiPTM) oxidize the residues at physiological pH, thus altering the structure and functions of proteins [44,45]. Therefore, the passive ROS diffusion through aquaporins and the direct interaction occurring in chloroplasts and nuclei should be tightly controlled by antioxidant systems to neutralize ROS and oxidation to mitigate damage [46,47]. For example, H2O2 profiling analyses showed that it catalyzes the biochemical reactions at lower concentration (e.g., average H2O2 concentration, 10 μM in cellular compartments especially peroxisomes and apoplast mediate normal cell division) but creates oxidative damage at higher concentrations in plants (Supplementary Figure S1) [48,49]. The H2O2 scavenging is accomplished by coordinated activities of antioxidative enzymes [50] such as SODs, which processes O2 to H2O2 and are ubiquitously expressed in apoplast, cytosol, peroxisomes, chloroplasts, mitochondria and nuclei [51,52].
Photosynthesis in chloroplasts is directly linked to cellular redox regulation in plants [16]. Chloroplast-associated 1O2 is produced from chloroplast triplet state (3Chl) by interacting with O2 molecules (ground state 3O2), specifically in PSII under varying irradiance [49,53]. Elevated 1O2 levels in the PSII reaction center cause photo-inhibition resulting in oxidative damage [54] and metabolic breakdown, leading to oxidation of the D1 protein, which specifically dephosphorylates the PSII reaction center [55]. The irreversible photo-inhibition drives the expression of 1O2 related genes, which activate the antioxidative system, and ultimately induce resistance against high light (HL) and other stresses in plants [54,56,57]. Moreover, PSI oxidation is known as the Mehler reaction—photo-reduction of O2 [58]. This reaction is catalyzed by thylakoid and stromal-associated SODs, producing O2•− as the first product and then dismutase O2 to H2O2. H2O2 may cause oxidation to Calvin-Benson Cycle (CBC) components regulated by the thioredoxin system [59], 2-Cys peroxiredoxin and ascorbate peroxidase (APX) [60]. Further reduction of H2O2 to H2O is catalyzed by the integration of APX, thiol related enzymes such as thioredoxin (TRX), peroxiredoxin (PRX) and NADPH. However, the enzymatic system is tightly regulated in cellular compartments where H2O2 retains the potential to move out of chloroplast and mitochondria interacting with transcription factors for retrograde signaling [56,61,62,63,64], which regulate the gene expression in the nucleus. In Arabidopsis, the NAC domain containing protein 17 (ANAC017) interacts with enhanced H2O2 levels and modulates the gene expression [65]. The ANAC017 activity is inhibited by radical-induced cell death 1 (RCD1), which mediates ROS-related retrograde signaling in mitochondria has also been anticipated [66]. Examining the dynamics of chloroplast demonstrate that this organelle can actively sense environmental cues, regulating the nucleus-chloroplast communication and gene expression. Chloroplast associated ROS retrograde signaling is largely dependent on 1O2, produced as a by-product of PSII reactions. The production of 1O2 facilitates the chloroplast to nuclear communication, which ultimately modulates the gene expression, responds to stress, and programmed cell death [67]. Taken together, these studies provide a mechanism for how PSII and PSI mediate redox reactions are the keys to balance photosynthesis.
Both PSII and PSI are sensitive to light intensities and other abiotic stresses. Redox homeostasis is disturbed when the rate of damage is higher than repair, causing photoinhibition [68]. The excessive energy under HL can be dissipated as heat via non-photochemical quenching (NPQ), which subsequently adjust the chloroplasts composition and metabolism. The presence of an antioxidant system in chloroplast helps in nucleus-chloroplast communication that drives the gene expression. In addition, peroxisomes accumulate H2O2 under photorespiration [49], which is removed by peroxisome associated CATs [69]. Recent findings indicate that antioxidants ascorbate (Asc) and glutathione (GSH) regulate gene expression under stresses [70,71], suggesting their dynamic nature and high sensitivity to stress cues. In summary the mechanisms are now well-defined, reveal that moderate levels of ROS are essential for cell proliferation, photosynthesis mechanisms, and maintaining redox homeostasis at a basal level within chloroplasts. However, ROS signaling components are less studied in relation to the evolution of the key protein families from algae to plants.

3. Molecular Evolution of Redox Regulatory Network

Modern chloroplasts arise from photosynthetic prokaryotes. Green plant chloroplast genomes contain small proportions of genes than their ancestors [7], indicating that endosymbiosis resulted in the loss and relocation of genetic information. During the transition from aquatic life to terrestrial habitats, green plants have evolved some new pathways for redox regulation and defense systems. Trx-based redox regulation has been observed in all groups of life, indicating a significant role under diverse redox environments [72,73].
The alterations in the environment cause large evolutionary pressure, leading to the emergence of new functions to existing genes and the formation of new genes [74,75]. Comparative genomic analyses revealed the regulatory roles of conserved regions of key gene families for diverse cellular functions such as photosynthesis in eukaryotes [76]. Thus, the evolutionary histories of redox homeostasis and photosynthesis can be drawn through comparative genetic analysis, gene expression profiles, phylogenies, conserved domain analysis, and prediction of 3D protein structures. In our previous comparative molecular evolution studies, we have revealed conserved features of over 100 gene families in green algae and land plants [77,78,79,80,81,82,83,84]. Here, the functional regulatory networks among distinct species were analyzed through comparative genetic analysis of sequences in redox signaling of photosynthesis from evolutionarily important lineages of plants and algae (Figure 2 and Figure 3, Supplementary Figures S2–S5).

4. ROS Related Gene Families Are Highly Conserved across Land Plants and Green Algae

We found that there are evolutionarily conserved features for ROS signaling and photosynthesis in the examined major green plant lineages (Figure 2, Supplementary Figure S3). Phylogenetic analysis suggests that these families may have evolved from streptophyta. The structure of TrxLs contains conserved WCRKC domain with two cysteine residues which serve as redox switches.
The dominant regulators with oxidation activities in chloroplast are Trx-like, NADPH-dependent Trx- reductase C (NTRC) proteins [88]. These proteins further comprise Trx-like 2 (TrxL2), and atypical Cys His-rich Trx (ACHT) groups based on oxidation factors [89,90,91,92,93,94,95], and catalyze H2O2 reduction through interacting 2-Cys PRX (2CP). Different subfamilies of Trxs may exist, such as Trx-x in Arabidopsis [96], Trx-y in green algae Chlamydomonas reinhardtii and Trx-z in Arabidopsis and Solanum lycopersicum [97,98]. The Trx- gene number in plant species is largely expanded in angiosperms, for example, there are more than twenty Trx isoforms in Arabidopsis [99]. The oxidation factors family proteins can be further classified into Trx-like f-, m-, x-, y-, or z- with unknown functions except for TRX-f, which specifically carried out the oxidation of target photosynthesis proteins [93,100]. In addition, five isoforms of ACHT were reported in Arabidopsis [101], concomitantly ACHT1 and ACHT2 revealed an association with targeted oxidation.
Recent in vivo experiments confirmed the roles of Trx and ACHT in oxidation processes in the chloroplast [102]. Yokochi et al. [102] discovered that the Trx-f and TrxL2.1 serve as oxidation factors of CBC enzymes, ribulose-1,5-bisphosphate carboxylase/oxygenase activase (RCA) and gamma subunit of ATP synthase (CF1-γ), respectively ACHT1 and ACHT2 play a redundant role in oxidizing fructose-1,6-bisphosphatase (FBPase) while phosphoribulokinase (PRK) regulation under Trx- system is still elusive. Comparative genomic studies suggest that TrxL2.1 and TrxL2.2 are conserved in photosynthetic organisms including flowering plants, mosses, and streptophyte algae (Figure 2 and Figure 3), but not in chlorophyte algae such as Chlamydomonas reinhardtii [88]. Both TrxL2.1 and TrxL2.2 share functional residues and biochemical features but differ in expression patterns [90]. In Arabidopsis, TrxL2.1 and ACHT2 highly express in leaves [102,103]. The trxl2.1 and acht knockout mutants displayed similar phenotypes and physiology (chlorophyll content and photosynthetic parameters) compared to those in the wild type under low light. In contrast, ntrc mutants showed reduced ROS levels with reduced growth and pale green phenotypes. Interestingly, ntrc mutants with background Trxl2.1 largely recovered the wild-type phenotype, suggesting that 2CP retains the potential to accept electrons from both TrxL2.1 and ACHT, and the NTRC oxidation system. In addition, it was revealed that reduced growth of Arabidopsis ntrc mutants was associated with overoxidation of CBC proteins rather than imbalanced redox regulation or reduced H2O2 scavenging. A remarkable positive association was observed between ACHT levels and NPQ, which reveals overexpressed ACHT led to high NPQ [102]. These studies suggest that TRX and ACHT are key players in redox regulation in CBC. Future experiments will uncover the underlying mechanism of remaining redox regulators for photosynthesis.

5. Key Photosynthesis Related Gene Families Are Evolved from Streptophyta Algae

We found that CBC enzymes share common features among all tested evolutionary lineages (Figure 2, Supplementary Figure S4). Phylogenetic analysis suggests that these families may have evolved within streptophyta. For instance, the structure of GAPDH holds a conserved functional NAD binding domain that is used as a coenzyme.
The CBC enzymes including RCA, glyceraldehyde-3-phosphate dehydrogenase (GAPDH), CF1-γ, FBPase, SBPase, and PRK [59,73] are highly sensitive to irradiance with high activities under light conditions. RCA is involved in the activation of Rubisco and belongs to the AAA+ family [104], which contains the α and β isoforms. These two isoforms may be generated from alternate splicing or encoded by different genes [105] in green algae and land plants with the exception few species including Chlamydomonas have only a short RCAβ isoform [106]. Both RCA isoforms differ with the presence of two conserved C-terminal Cys residues, where reduction is catalyzed by Trxf and ATP/ADP ratio under light conditions. GAPDH is further grouped into gapA and gapB, sharing similar composition with a distinction C-terminal extension residing 30 amino acids (aa) in gapB. These extended aa contain Trx- target Cys residues regulating light and dark reactions. gapA and gapB are present in streptophyta and green algae [107,108], with the exception that gapB was found absent among most oxygenic phototrophs, containing only gapA copies or additional gapC [109].
FBPase is required for dephosphorylation of fructose-1,6-bisphosphate (F1,6P) to fructose-6-phosphate (F6P). Light activation of FBPase and PRK was first discovered in Chlorella [110] and subsequently observed in higher plants [111,112]. PRKs show striking features along the evolutionary lineages, for example, homodimeric in eukaryotes, heterodimeric in cyanobacteria, and octameric in non-photosynthetic prokaryotes [113]. Considering the redox regulatory components, FBPase and PRK activities are largely dependent on Trxf and Trx-m, respectively, which carry out disulfide reduction and make them activated [114]. As an oxidizing enzyme, SBPase catalyzes sedoheptulose-1,7-bisphosphate (S1,7P) to sedoheptulose-7-phosphate (S7P), which has been discovered in many photosynthetic organisms except cyanobacteria. However, cyanobacteria contain a bifunctional SBPase with similar activities [115]. The findings that share sequence similarity with SBPase and Trxf indicate that both enzymes may have a common evolutionary origin [116].
C4 and CAM photosynthesis metabolism is regulated by NADP+-dependent malate dehydrogenase (NADP-MDH). NADPH serves as an electron donor during C4 CO2 fixation by reducing oxaloacetate into malate during the day and at night in CAM plants [117]. The molecular mechanism associated with NADP-MDH was first observed in the C4 plant, Sorghum bicolor, and subsequently in C3 and CAM plants [118]. NAPH-MDH is highly dependent on light-responsive Trxs, possibly Trxf or Trxm types evolved from Chlamydomonas reinhardtii [119]. In plants, NADP-MDH reside Cys extensions both at the C- and N- terminals, but green algae contains only C-terminal residues [119]. Although enzymes associated attributes have been reported, functional validations in a range of evolutionarily important green plants are yet to be conducted in future research work.

6. Antioxidant Related Gene Families Are Conserved across Land Plants and Green Algae

In our previous publications, we reported an early evolution of antioxidative enzymes from Chlorophyte algae [63,78]. SOD, CAT, POD, GPX, GR, glutathione S-transferase (GST), ascorbate peroxidase (APX), monohydroascorbate reductase (MDHAR) and dehydroascorbate reductase (DHAR), Asc, GSH, and antioxidant, phenolic compounds, alkaloids, flavonoids, carotenoids, free amino acid, and alpha tocopherols [40,41,42] play crucial roles in catalyzing oxidation-reduction reactions to regulate redox homeostasis in plants. The first reported antioxidant enzyme was SOD, which removes ROS, and is thought to have evolved even before the origin of eubacteria and archaea [30]. In cyanobacteria, several antioxidants containing protein and non-protein molecules have evolved from Synechococcus PCC [120]. The constant release of O2•− anions from the photosynthetic machinery and respiratory electron transport chain [120], and on other cellular surfaces [121] are scavenged by SODs and superoxide reductase (SORs). Similarly, PODs and CATs potentially reduce the stability of H2O2, RO and OH radicals [122]. Although O2•- are produced as by-products of aerobic life, they also promote iron acquisition, cell signaling, and growth [121,123]. O2•− appears to function differently upon over-accumulation in plant cells, destabilizing DNA by inducing Fenton reactants [124] and depletion of 4Fe-4S clusters in proteins [125], which are required for photosynthesis and amino acid biosynthesis, respectively. Taken together, the results provide a clue that redox homeostasis needs to be maintained to regulate both the molecular and biological functions of organisms. This study highly supports the existence of SODs in the genomes of Eukarya, e.g., land plants and green algae (Supplementary Figure S5A,B), archaea, and bacteria [126].

7. Redox Regulation during Photosynthesis

The light-driven electron transport chain (ETC) mediates its function by recruiting H2O, generates reducing power NADPH in the thylakoid membrane, and initiates the proton (H+) gradient that leads to ATP synthesis [3]. In chloroplasts, redox status is regulated by Cys residues of photosynthetic enzymes, which are highly conserved and retain the potential to switch ON or switch OFF signaling [127], indicating some fine-tuned mechanisms are involved during photosynthesis. The dominant ROS that reduces photosynthetic efficiency are 1O2, H2O2 and O2•−, in PSI [128]. Plant chloroplasts have evolved different strategies to minimize photo-oxidative stress. One such strategy includes the maintenance of CBC enzymes that contain redox switches as Cys residues [59,129]. Other strategies involve non-cyclic electron flow (NCEF, reduced photochemical energy; PQ and ETC), and cyclic electron flow (CEF), photorespiration, balanced PSI/PSII ratio, O2 photo-reduction, and protein quality have been adopted to increase photosynthetic efficiency [130,131,132,133]. Lowering PQ and ETC initiates NPQ which results in subsequent reductions in the excitation energy of chlorophyll a, which is then dissipated as heat [133,134,135]. On the other hand, CEF produces higher ATP relative to NADPH; thus O2•− and H2O2 become the acceptors of energy in PSI [136]. CEF seems to be important because it reduces 1O2 synthesis at PSII and increases the proton gradient in thylakoid membrane and activates NPQ processes [137] (Supplementary Figure S6). Whether the cyclic or non-cyclic mechanism activates to protect the photosynthetic machinery is highly dependent on photosynthetic cell and/or organelle types and environmental cues.

8. Regulation in Calvin Benson Cycle

Here, we propose a simplified model of the mechanisms of conserved enzymes in the operation of photosynthesis (Figure 4A). Light and dark reactions are catalyzed by unique regulatory pathways involving enzymes with conserved functional Cys residues [59,129]. Among them, Trx- and NTRC systems play a significant role in maintaining redox homeostasis in the chloroplast [73]. As well, Trx- system encoding multicomponent may be localized in chloroplast, mitochondria, nucleus, and cytoplasm [99], thus plant chloroplasts have widely adapted Trx- system that carried out CO2 fixation via disulfide reduction of CBC enzymes, including RCA, CF1-γ, FBPase, SBPase and PRK [59,73]. Increasing evidence suggest that regulatory proteins become higher in expression under light, while lower in the dark reactions [138,139]. The reducing powers from ETC are transferred to TRX through ferredoxin (fd) and ferredoxin-TRX reductase (FTR) [140]. Further, the activated TRX targets the corresponding photosynthetic enzymes, and reduce the disulfide bond. Also, it was found that NTRC power transfer efficiency is higher than TRX proteins; thus, NTRC may function as main regulator donating an electron to 2CP. In plants, the affected stability of NTRC led to phenotypic abnormalities, decreased chlorophyll content, and higher NPQ [141,142,143,144,145,146,147].
Under dark reactions in the chloroplast stroma, the redox-related proteins are reoxidized and reoxidation is needed to balance the NADPH status in the chloroplast, especially under light fluctuations. This phenomenon is accomplished by a plastic malate valve, consisting of oxaloacetate/malate transporter 1 (OMT1) [148] and malate dehydrogenases (MDHs), which carry out stromal OAA reduction to malate [149]. Coexisting MDHs isoforms have a differential preference for coenzymes, such as NADPH-MDH has a strong preference for NADP [H] [150] while NAD-MDH chooses NAD [H] [151]. Interestingly, their co-existence has caused much confusion among scientists for several years. The opinion is that chloroplasts use NAD, merely converting it into NADP by NAD kinase 2 (NADK2) under light, while NADH is not further phosphorylated [152]. Earlier studies showed that FNR’s have strong affiliations for NADP [H] [153,154,155]. Similarly, the chloroplast stromal NADPH conversion into NADH interferes with balanced NADP [H] synthesis and utilization and thus affects photosynthesis, precisely CBC [156]. A recent study revealed that NADP-MDH plays a crucial role in chloroplast redox homeostasis under varying light intensities [127]. This pathway activates chloroplast NADP+ dependent MDH via Trxs, subsequently, the generation of NADP+ and increased reducing power is transported and discharge pressure from the chloroplast. The loss function of MDH in Arabidopsis leads to stunted growth in response to short days or fluctuating light while in the dark the phenotypic differences are still unchanged. MDH-mediated redox regulation plays a crucial role in response to prolonged dark period and varying light intensities [127]. Considering the importance of oxidation, the identification of uncharacterized players will further explore the associated pathways for redox homeostasis in photosynthesis.
The core photosynthetic enzymes have also been found in microorganisms [157,158]. The question arises, how redox switches are regulated by nonplastidial type homologs of CBC enzymes? It should be noted that these homologs do not hold redox switches while remaining activated without reduction via redox transmitters. Dark conditions are essential for maintaining redox status, such as FBPase restricts excessive ATP utilization during the reaction catalyzed by phosphofructokinase [73]. In addition, ATP synthase restricts a useless reverse reaction and ATP hydrolysis [159,160] and MDH limits the reducing molecules’ export to chloroplasts through the malate valve [149,161,162]. The structure of MDH contains Cys residues both at N- and C- terminals indicating that dark mediated MDH regulation is tightly controlled [161,163]. While FBPase and MDH have essential functions in chloroplast metabolic homeostasis under varied light intensity and metabolic states [139,164]. the detailed mechanism of redox regulation for plant metabolism is still elusive.

9. Regulation in Kranz Anatomy

C4 photosynthesis appears in two types of cells [mesophyll cell (MC) and bundle sheath cell (BSC)], known as Kranz anatomy (Figure 4B). It utilizes an additional two ATPs and results in a higher photosynthetic rate, lower photo-respiratory flux [165] and reduced CO2 compensation point [166]. The transition of photorespiration between both cell types upregulates uncoupling protein (UCP) and bundle sheath mitochondria alternative oxidase (mAOX), which serve as a valve to scavenge NADH via glycine carboxylase mediated mechanism [166]. The activated carboxylation in C4 bundle sheath export NADH to the respiration electron chain, and AOX maintain the ROS balance [166]. Earlier studies showed no significant alterations in the redox systems of C4 and C3 (comprising only MC). The differences indicate the use of alternative decarboxylation enzymes in distinct species [165,167].
In general, C4 species employ CEF around PSI, export electron from Fd to plastoquinone (PQ) and generates an H+ gradient in the thylakoid membrane. That the outcome is a loss of PSII activity, and subsequent accumulation of PSI subunits in chloroplasts [168,169], which may involve the 5-PGR5-like photosynthetic phenotype 1 (PGR5-PGRL1) pathway [170,171]. This pathway diverts the excessive flux via plastid terminal oxidase, namely PTOX or IMMUTANS, and discharges pressure from chloroplast [172]. These differences indicate that some aspects of redox homeostasis in C4 photosynthetic machinery may have resulted from divergent evolutions from C3 to C4 in angiosperms. In addition, the carboxylation pathway may differ with the prevalence of CEF between MC and BSC and the CEF/NCEF ratio in plant species [173,174]. Looking forward, studying the regulation of redox status between photosynthetic cycles and redox distribution between MC and BSC and other associated cycles will uncover the regulation underlying redox homeostasis in C4 plants.

10. Regulation in Crassulacean Acid Metabolism (CAM)

Crassulacean acid metabolism (CAM) functions by closing stomata during the daytime and fixing CO2 at nighttime [175,176] (Figure 4C). CO2 fixation occurs in malic acid at night, which is subsequently stored in the vacuole of MC. Similar to C4 photosynthesis, NADP-MDH dependent oxidation linked with Trx systems has been observed in CAM plants [117]. In C3 angiosperms, over-reduction of acceptor side of PSI and chloroplast stroma mediated “match valve” system transfer reducing equivalents to the cytosol, which generates NADP+ under light fluctuations [163,177]. It is still largely unknown whether CAM plants have a water-water cycle as C3 plants for adjusting PS1 redox status or restricted CO2 fixation under malic acid exhaustion. Recently, studies based on redox regulation in CAM photosynthesis of Bryophyllum pinnatum [178], Dendrobium officinale [179,180], and Vanilla planifolia [181] demonstrate that CEF and water-water cycle could be an important regulator as adjusting PSI redox homeostasis under light fluctuations. However, research work on redox homeostasis and photosynthesis in CAM plants requires acceleration to better compare the research advances in C3 and C4 plants.

11. Regulation of Redox in Early Divergent Plants and Green Algae

Green plants acquired the potential of oxygenic photosynthesis, which produces ATP and NADPH for CO2 fixation and subsequent pathways. Dominant Trx-based redox regulation is important in all groups of life under diverse redox environments. The transition of plants from an aquatic environment to land may have enabled the evolutionary innovation of new pathways or new components of pathways for oxidative regulation and defense systems. Considering the redox regulation in oxygenic photosynthetic organisms, the FTR mechanism appeared dominant [157] (Figure 4D). It has been hypothesized that plant NTRC functional pathways resemble cyanobacteria FTR and are directly associated with metabolism in response to light [182,183]. However, future research work is required to test these hypotheses experimentally.

12. Metabolism of Redox Regulators in Maintaining Redox Homeostasis for Photosynthesis

The detoxification of ROS in chloroplast and associated organelles during photosynthesis is carried out via complex pathways. For instance, the ascorbate pathway plays crucial roles in diverse photosynthetic processes with ascorbate-dependent and independent routes. Ascorbate is produced in mitochondria and utilized in the chloroplast, whereby the transportation to chloroplast membranes from cytosol occurs via ascorbate transporters and stabilizes the thylakoid membrane against oxidation by reducing tocopheroxyl radicals to non-oxidizing form [184]. In the ascorbate-dependent pathway, oxidation of H2O occurs due to damage to PSII [173]. Later, the integration and interaction of antioxidative enzymes reduce H2O2 to water. On the other hand, the ascorbate-independent cycle is regulated via Fd/Fd-dependent Trx reductase (FTR)/TRX or NADPH/NADPH Trx C (NTRC) and PRX [185]. It was revealed that more than one type of ascorbate transporters may exist, which facilitates the ascorbate diffusion from cytosol to chloroplast stroma [186,187]. One such group includes nucleobase- ascorbate transporters (NATs) or nucleobase- cation symporter 2 (NCS2) comprising ascorbate transportation and DNA bases [188]. They have been found in all kingdoms of life ranging from unicellular bacteria to multicellular animals [188,189]. All members in the NAT family share similar features regarding amino acid sequences and conserved trans-membrane proteins [190,191]. Functional characterization of NAT proteins in plants figures out their vital roles in plant growth and development. Taking an example, leaf permease 1 (Lpe1) was first identified as a NAT family member in plants [192], which helps in the transport of xanthine and uric acid but not ascorbic acid [191]. Loss function of Lpe1 phenotype shows defective chloroplasts and lost plasma membrane integrity [192]. Other group includes phosphate transporter proteins, such as Arabidopsis AtPHT4.4 is a member of the phosphate transporter 4 family that facilitates ascorbate movement into chloroplast envelope membranes [193]. AtPHT4;4 serves as a co-transporter with Na+/Pi and Cl dependent activity facilitating the Δψ to become a driving force. The loss function of AtPHT4;4 reduces ascorbate levels and disrupts the xanthophyll cycle, which is ultimately responsible for removing excessive photosynthetic energy as heat [193]. Remarkably, the loss of AtPHT4.4 does not affect shoot phenotype and shows HL tolerance when its antioxidative properties are most required. In addition, AtPHT4.4 is localized in palisade cells rather than MC, which calls for further investigations.
The cellular functions and comparison of antioxidants have been investigated in model and non-model C3 and C4 plants, such as, Helianthus annuus (C3) and Sorghum bicolor (C4) [194], Triticum aestivum (C3) and Zea mays (C4) [195,196], Cleome spinosa (C3) and Cleome gynandra (C4) [197], and Flaveria sp. [198]. Flaveria sp. had higher redox scavenging for CATs under C3 while APXs and GRs were found up-regulated in C4 photosynthesis [198]. CATs activities were inhibited in a C4 species Flaveria bidentis, which were later confirmed at the protein level [198], suggesting the change in photorespiration may result in altered antioxidant activity. Similarly, APXs-associated scavenging was found to be lower in C4 when compared to C3 species [197,199]. Taking the example of Flaveria bidentis, these changes occur from young leaves (C3) to mature plants (C4) due to chloroplasts’ dimorphic nature [200]. Furthermore, APX-mediated H2O2 scavenging was down-regulated four-fold, while PRXs were up-regulated five-fold [199]. Their results provide a potential mechanism for antioxidant activity in both types of photosynthesis, which may be activated vice versa depending on chloroplast responses. Interestingly, both APX and PRX work in coordination; for instance, any genetic change in one will be compensated by the other enzyme to release ROS from chloroplast [201,202]. Considering the role of antioxidative enzymes in redox regulation, we anticipate that future efforts will uncover the uncharacterized genes at the cell/organelle level and further explore their scavenging functions.

13. Manipulating ROS Signaling to Enhance Plant Photosynthetic Efficiency and Crop Yield

Plants being sessile organisms are routinely exposed to environmental fluctuations (Figure 5). Photosynthetic machinery is sensitive to changing light intensities, and photo-damage may occur in response to HL intensities. Photosynthate production (i.e., crop yield) is highly dependent on light energy conversion into chemical energy during photosynthesis [203]. Photoinhibition or damaged PSII affects the D1 protein turnover, which takes part in PSII repairing, encoded by a chloroplastic psbA gene [204]. The activated PSII repair process first degrades the damaged D1 proteins and then synthesizes D1 precursors (pre-D1) and D1 to reactivate PSII [205]. Recently, a bioengineering strategy was employed to install a D1 pathway in the nucleus that enhance the D1 turnover, ultimately leading to enhanced photosynthesis and plant yield under both control and high-temperature conditions [204]. Genetic engineering of nuclear D1 with plastid peptide sequence of RuBisco under the control of heat shock promoter AtHsfA2 promotes PSII activities with tolerance against heat stress in Arabidopsis thaliana, Nicotiana tabacum, and Oryza sativa. The de novo synthesis of D1 precursors both in chloroplast and nucleus mitigates the deficient D1 demand required for repairing PSII under photo-damage [204].
In addition, plants have adopted photo-protective mechanisms, such as NPQ, which prevent the over-excitation of photosynthetic light-harvesting antenna complexes [134]. Although the excitation of NPQ chlorophyll fluorescence may happen rapidly, the relaxation phase is gradual [206], thus exposing the PSII to repeated, varying light intensities and low quantum yield. The slow recovery and decreased CO2 fixation under quenched and non-quenched states, allow the photosynthetic apparatus acclimatization to occur via thylakoid-associated protein PsbS and the xanthophyll cycle [207,208]. Under light fluctuations, the xanthophyll reactions are catalyzed by the reversible inter-converted pigment molecules, zeaxanthin and violaxanthin, which are mediated by zeaxanthin epoxidase (ZEP) and violaxanthin de-epoxidase (VDE) respectively. It was revealed that ZEP accumulation and NPQ installation show a similar mechanism in Oryza sativa, Hordeum vulgare, and Spinacia oleracea under varying irradiance absorbance [209]. In addition, the plants retain the potential to establish NPQ under HL and more rapid re-establishment of photosystem complexes under decreased light has also been anticipated [210,211] indicating balanced or increased plant biomass under varying light intensities. Recently, Kromdijk et al. [212] and Garcia-Molina et al. [213] employed bioengineering of xanthophyll cycle components and PsbS protein in Nicotiana tabaccum and Arabidopsis thaliana, respectively. The findings that overexpression of VDE, ZEP, and PsbS enhanced photo-protection in response to NPQ in both species, and promoting biomass in tobacco, demonstrating that NPQ is essential for plant growth but may interfere with decarboxylation capacity, excessive energy distribution, or retrograde signaling. Moreover, the overexpressed lines in tobacco revealed higher capacities for CO2 uptake, accounting for the accumulated biomass [212].
The core component of the xanthophyll cycle and PsbS protein with NPQ activities have been found conserved among flowering plants and green algae [214]. In the microalgae, Chlamydomonas reinhardtii, random mutations in Light Response Signaling protein 1 (LRS1), a homolog of plant phospholatory protein COP1 were found to be associated with HL adaptation [215]. Similarly, in the cyanobacteria Synechocystis sp., the functional characterization of representative mutant proteins revealed higher adaptation under HL cues. In contrast, a significant negative correlation was found between the overexpression of antioxidative systems and plant biomass accumulation. Such as APXs hinder the accumulation of H2O2 accumulation in the nucleus inhibiting the HL-responsive gene expression [62]. Therefore, we are still in the early stages of understanding the role of bioengineering for regulatory elements and antioxidants for increasing crop productivity.

14. Concluding Remarks and Future Perspectives

ROS toxicity affects almost all aspects of plant metabolism and plants have evolved ROS scavenging both at cellular-based and antioxidant-based systems. ROS such as H2O2 and O2 have important functions in chloroplast, mitochondria, and nuclei. Excessive ROS is either passively diffused via aquaporins or scavenged by antioxidative enzymes and antioxidants. Although several antioxidative enzymes and antioxidants have been identified in plants, a clear model for ROS toxicity at all levels and the participation of each antioxidant in response to multiple external and internal stimuli has yet to be described. Furthermore, there may exist other antioxidant members or isoenzymes of complex photosynthesis ROS regulation awaiting discovery in plants. Photosynthesis metabolism is regulated by evolutionary conserved systems, such as TRX and NTRC. Transcriptome-wide identification of redox regulators of photosynthesis has been analyzed in green algae and higher plants. Recent studies suggest that Trx and NTRC systems play important roles in maintaining photosynthetic efficiency. Therefore, the existing question of how redox systems interact with photosynthesis machinery is becoming clearer than before. However, some redox regulators and systems are still uncharacterized in plants. Considering that redox regulatory pathways have been established, we anticipate that future studies will explore these uncharacterized components at transcriptome levels and functional validation in photosynthesis. Such as, several isoforms and groups related to Trx- system, such as, m-, h-, z-, etc. have been identified but a few provide clear insights into the association with CBC enzyme and others’ participation in reduction processes either directly or indirectly has yet to be evaluated. In addition, redox homeostasis was shown to be regulated by evolutionarily conserved enzymes with antioxidative properties catalyzing photosynthetic efficiency and stress response.
Distinct photosynthetic machinery has evolved similar and diverse features for regulating ROS levels within chloroplasts. In Calvin cycles, Trx and NTRC interaction with photosynthesis enzymes are clearer than C4 photosynthesis, which involves chloroplasts of mesophyll and bundle sheath cells. To further enhance understanding of these types of photosynthesis, key steps will involve the identification of redox regulatory genes, Cys residues, and their interaction with C4 and CAM photosynthetic enzymes. Recently, a few studies have characterized Trx-based regulators in the model plant Arabidopsis and their associated cellular and biological functions, while it remains elusive in other plants. In addition, antioxidant-mediated pathways regulating plant growth, development, and stress response in all types of photosynthesis will be helpful in future crop breeding programs for the enhancement of photosynthesis.
Considerable bioengineering and fast-forward genetic approaches have been utilized to maintain protein efficiency in model plants and some staple crops, but fewer studies were conducted in key early divergent evolutionarily important green plant species and many crops. Therefore, future research focus should be placed on those plant species to improve our understanding of the molecular evolution of redox homeostasis of in photosynthetic machinery and to increase crop yield potential for future food security.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/antiox11112085/s1, Figure S1: Barley plant growth under low and high H2O2; Figure S2: Three-dimensional (3D) protein structures of GAPDH and TrxL2.2 in representative plant and alga species; Figure S3: Phylogenies of redox related genes in plants and algae; Figure S4: Phylogenies of photosynthesis related genes in plants and algae; Figure S5: Phylogenies of antioxidant related genes in plants and algae; Figure S6: Schematic representation of cyclic and non-cyclic electron flow. Table S1: ROS, half-life, migration distance, their production sites and known functions in Plants.

Author Contributions

A.R., F.Z. and Z.-H.C. conceived the research; A.R., G.C., W.J. and Q.Z. conducted evolutionary analysis; F.D., M.M. and F.Z. conducted literature review; B.R., F.D. and F.Z. prepared the figures. A.R. and Z.-H.C. wrote the paper with the support of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

The research has been supported by the Major International (Regional) Joint Research Project from NSFC-ASRT (32061143044), National Natural Science Foundation of China (32001456), the Australian Research Council (FT210100366, DE1401011143).

Acknowledgments

The authors acknowledge that Figure 1, Figure 4, Figure 5 and Figure S6 were created using biorender.com (accessed on 19 May 2022).

Conflicts of Interest

Authors declare that there is no conflict of interest.

References

  1. Olson, J.M. Evolution of photosynthetic reaction centers. Biosystems 1981, 14, 89–94. [Google Scholar] [CrossRef]
  2. Hohmann-Marriott, M.F.; Blankenship, R.E. Evolution of Photosynthesis. Annu. Rev. Plant Biol. 2011, 62, 515–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Fischer, W.W.; Hemp, J.; Johnson, J.E. Evolution of Oxygenic Photosynthesis. Annu. Rev. Earth Planet. Sci. 2016, 44, 647–683. [Google Scholar] [CrossRef]
  4. Zhu, X.G.; Long, S.P.; Ort, D.R. Improving photosynthetic efficiency for greater yield. Annu. Rev. Plant Biol. 2010, 61, 235–261. [Google Scholar] [CrossRef] [Green Version]
  5. Lu, Z.; Chang, Y.C.; Yin, Q.-Z.; Ng, C.; Jackson, W.M.J.S. Evidence for direct molecular oxygen production in CO2 photodissociation. Science 2014, 346, 61–64. [Google Scholar] [CrossRef] [Green Version]
  6. Meadows, V.S.J.A. Reflections on O2 as a biosignature in exoplanetary atmospheres. Astrobiology 2017, 17, 1022–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Timmis, J.N.; Ayliffe, M.A.; Huang, C.Y.; Martin, W. Endosymbiotic gene transfer: Organelle genomes forge eukaryotic chromosomes. Nat. Rev. Genet. 2004, 5, 123–135. [Google Scholar] [CrossRef]
  8. Rosing, M.T.; Frei, R.J.E.; Letters, P.S. U-rich Archaean sea-floor sediments from Greenland–indications of >3700 Ma oxygenic photosynthesis. Earth Planet. Sci. Lett. 2004, 217, 237–244. [Google Scholar] [CrossRef]
  9. Anbar, A.D.J.S. Elements and evolution. Science 2008, 322, 1481–1483. [Google Scholar] [CrossRef] [PubMed]
  10. Boyd, E.S.; Thomas, K.M.; Dai, Y.; Boyd, J.M.; Outten, F.W.J.B. Interplay between oxygen and Fe–S cluster biogenesis: Insights from the Suf pathway. Biochemistry 2014, 53, 5834–5847. [Google Scholar] [CrossRef] [PubMed]
  11. Halliwell, B.; Gutteridge, J.M.C. Free Radicals in Biology and Medicine; Oxford University Press: Oxford, UK, 2015. [Google Scholar]
  12. Mehla, N.; Sindhi, V.; Josula, D.; Bisht, P.; Wani, S.H. An Introduction to Antioxidants and Their Roles in Plant Stress Tolerance. In Reactive Oxygen Species and Antioxidant Systems in Plants: Role and Regulation under Abiotic Stress; Khan, M.I.R., Khan, N.A., Eds.; Springer: Singapore, 2017; pp. 1–23. [Google Scholar]
  13. Hasanuzzaman, M.; Bhuyan, M.; Anee, T.I.; Parvin, K.; Nahar, K.; Mahmud, J.A.; Fujita, M. Regulation of Ascorbate-Glutathione Pathway in Mitigating Oxidative Damage in Plants under Abiotic Stress. Antioxidants 2019, 8, 384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Raja, V.; Majeed, U.; Kang, H.; Andrabi, K.I.; John, R. Abiotic stress: Interplay between ROS, hormones and MAPKs. Environ. Exp. Bot. 2017, 137, 142–157. [Google Scholar] [CrossRef]
  15. Foyer, C.H.; Noctor, G. Redox signaling in plants. Antioxid. Redox Signal. 2013, 18, 2087–2090. [Google Scholar] [CrossRef]
  16. Dietz, K.-J.; Turkan, I.; Krieger-Liszkay, A. Redox- and Reactive Oxygen Species-Dependent Signaling into and out of the Photosynthesizing Chloroplast. Plant Physiol. 2016, 171, 1541–1550. [Google Scholar] [CrossRef] [Green Version]
  17. Huang, S.; Van Aken, O.; Schwarzländer, M.; Belt, K.; Millar, A.H. The Roles of Mitochondrial Reactive Oxygen Species in Cellular Signaling and Stress Response in Plants. Plant Physiol. 2016, 171, 1551–1559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Sandalio, L.M.; Romero-Puertas, M.C. Peroxisomes sense and respond to environmental cues by regulating ROS and RNS signalling networks. Ann. Bot. 2015, 116, 475–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Choudhury, F.K.; Rivero, R.M.; Blumwald, E.; Mittler, R. Reactive oxygen species, abiotic stress and stress combination. Plant J. 2017, 90, 856–867. [Google Scholar] [CrossRef]
  20. 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]
  21. Singh, A.; Kumar, A.; Yadav, S.; Singh, I.K. Reactive oxygen species-mediated signaling during abiotic stress. Plant Gene 2019, 18, 100173. [Google Scholar] [CrossRef]
  22. König, J.; Muthuramalingam, M.; Dietz, K.-J. Mechanisms and dynamics in the thiol/disulfide redox regulatory network: Transmitters. Sens. Targets 2012, 15, 261–268. [Google Scholar]
  23. Mignolet-Spruyt, L.; Xu, E.; Idänheimo, N.; Hoeberichts, F.A.; Mühlenbock, P.; Brosché, M.; Van Breusegem, F.; Kangasjärvi, J. Spreading the news: Subcellular and organellar reactive oxygen species production and signalling. J. Exp. Bot. 2016, 67, 3831–3844. [Google Scholar] [CrossRef] [PubMed]
  24. Vaahtera, L.; Brosché, M.; Wrzaczek, M.; Kangasjärvi, J. Specificity in ROS signaling and transcript signatures. Antioxid. Redox Signal. 2014, 21, 1422–1441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Fichman, Y.; Mittler, R. Rapid systemic signaling during abiotic and biotic stresses: Is the ROS wave master of all trades? Plant J. 2020, 102, 887–896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Waszczak, C.; Carmody, M.; Kangasjärvi, J. Reactive oxygen species in plant signaling. Annu. Rev. Plant Biol. 2018, 69, 209–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Willems, P.; Mhamdi, A.; Stael, S.; Storme, V.; Kerchev, P.; Noctor, G.; Gevaert, K.; Van Breusegem, F.J.P.P. The ROS wheel: Refining ROS transcriptional footprints. Plant Physiol. 2016, 171, 1720–1733. [Google Scholar] [CrossRef] [Green Version]
  28. Kerchev, P.I.; Van Breusegem, F. Improving oxidative stress resilience in plants. Plant J. 2022, 109, 359–372. [Google Scholar] [CrossRef] [PubMed]
  29. Noctor, G.; Reichheld, J.P.; Foyer, C.H. ROS-related redox regulation and signaling in plants. Semin. Cell Dev. Biol. 2018, 80, 3–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Miller, A.-F. Superoxide dismutases: Ancient enzymes and new insights. FEBS Lett. 2012, 586, 585–595. [Google Scholar] [CrossRef] [Green Version]
  31. 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] [Green Version]
  32. Castro, B.; Citterico, M.; Kimura, S.; Stevens, D.M.; Wrzaczek, M.; Coaker, G. Stress-induced reactive oxygen species compartmentalization, perception and signalling. Nat. Plants 2021, 7, 403–412. [Google Scholar] [CrossRef]
  33. Llorente, B.; Segretin, M.E.; Giannini, E.; Lobais, C.; Juárez, M.E.; Paulsen, I.T.; Blanco, N.E. Homecoming: Rewinding the reductive evolution of the chloroplast genome for increasing crop yields. Nat. Commun. 2021, 12, 6734. [Google Scholar] [CrossRef] [PubMed]
  34. Ermakova, M.; Danila, F.R.; Furbank, R.T.; von Caemmerer, S. On the road to C4 rice: Advances and perspectives. Plant J. 2020, 101, 940–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Liu, M.; Yu, H.; Ouyang, B.; Shi, C.; Demidchik, V.; Hao, Z.; Yu, M.; Shabala, S. NADPH oxidases and the evolution of plant salinity tolerance. Plant Cell Environ. 2020, 43, 2957–2968. [Google Scholar] [CrossRef] [PubMed]
  36. Geigenberger, P.; Smirnoff, N.; Van Breusegem, F.; Dietz, K.-J.; Noctor, G. Plant redox biology—On the move. Plant Physiol. 2021, 186, 1–3. [Google Scholar] [CrossRef]
  37. Mittler, R. ROS Are Good. Trends Plant Sci. 2017, 22, 11–19. [Google Scholar] [CrossRef] [Green Version]
  38. Suzuki, N.; Miller, G.; Morales, J.; Shulaev, V.; Torres, M.A.; Mittler, R. Respiratory burst oxidases: The engines of ROS signaling. Curr. Opin. Plant Biol. 2011, 14, 691–699. [Google Scholar] [CrossRef]
  39. Laurindo, F.R.; Araujo, T.L.; Abrahao, T.B. Nox NADPH oxidases and the endoplasmic reticulum. Antioxid. Redox Signal. 2014, 20, 2755–2775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. 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]
  41. Sachdev, S.; Ansari, S.A.; Ansari, M.I.; Fujita, M.; Hasanuzzaman, M. Abiotic stress and reactive oxygen species: Generation, signaling, and defense mechanisms. Antioxidants 2021, 10, 277. [Google Scholar] [CrossRef]
  42. Hasanuzzaman, M.; Bhuyan, M.; Zulfiqar, F.; Raza, A.; Mohsin, S.M.; Mahmud, J.A.; Fujita, M.; Fotopoulos, V. Reactive oxygen species and antioxidant defense in plants under abiotic stress: Revisiting the crucial role of a universal defense regulator. Antioxidants 2020, 9, 681. [Google Scholar] [CrossRef]
  43. Vanderauwera, S.; Suzuki, N.; Miller, G.; van de Cotte, B.; Morsa, S.; Ravanat, J.L.; Hegie, A.; Triantaphylidès, C.; Shulaev, V.; Van Montagu, M.C.; et al. Extranuclear protection of chromosomal DNA from oxidative stress. Proc. Natl. Acad. Sci. USA 2011, 108, 1711–1716. [Google Scholar] [CrossRef] [PubMed]
  44. He, H.; Van Breusegem, F.; Mhamdi, A. Redox-dependent control of nuclear transcription in plants. J. Exp. Bot. 2018, 69, 3359–3372. [Google Scholar] [CrossRef] [PubMed]
  45. Willems, P.; Van Breusegem, F.; Huang, J. Contemporary proteomic strategies for cysteine redoxome profiling. Plant Physiol. 2021, 186, 110–124. [Google Scholar] [CrossRef] [PubMed]
  46. Mullineaux, P.M.; Exposito-Rodriguez, M.; Laissue, P.P.; Smirnoff, N.; Park, E. Spatial chloroplast-to-nucleus signalling involving plastid-nuclear complexes and stromules. Philos. Trans. R. Soc. B Biol. Sci. 2020, 375, 20190405. [Google Scholar] [CrossRef] [PubMed]
  47. Tian, S.; Wang, X.; Li, P.; Wang, H.; Ji, H.; Xie, J.; Qiu, Q.; Shen, D.; Dong, H. Plant Aquaporin AtPIP1;4 Links Apoplastic H2O2 Induction to Disease Immunity Pathways. Plant Physiol. 2016, 171, 1635–1650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Cheeseman, J.M. Hydrogen peroxide concentrations in leaves under natural conditions. J. Exp. Bot. 2006, 57, 2435–2444. [Google Scholar] [CrossRef] [Green Version]
  49. Foyer, C.H.; Noctor, G. Stress-triggered redox signalling: What’s in pROSpect? Plant Cell Environ. 2016, 39, 951–964. [Google Scholar] [CrossRef] [Green Version]
  50. Foyer, C.H.; Noctor, G. Ascorbate and Glutathione: The Heart of the Redox Hub. Plant Physiol. 2011, 155, 2–18. [Google Scholar] [CrossRef] [Green Version]
  51. Alscher, R.G.; Erturk, N.; Heath, L.S. Role of superoxide dismutases (SODs) in controlling oxidative stress in plants. J. Exp. Bot. 2002, 53, 1331–1341. [Google Scholar] [CrossRef]
  52. Dvořák, P.; Krasylenko, Y.; Ovečka, M.; Basheer, J.; Zapletalová, V.; Šamaj, J.; Takac, T. FSD1: Developmentally-regulated plastidial, nuclear and cytoplasmic enzyme with anti-oxidative and osmoprotective role. Plant Cell Environ. 2020. [Google Scholar] [CrossRef]
  53. Fischer, B.B.; Hideg, É.; Krieger-Liszkay, A. Production, Detection, and Signaling of Singlet Oxygen in Photosynthetic Organisms. Antioxid. Redox Signal. 2013, 18, 2145–2162. [Google Scholar] [CrossRef] [PubMed]
  54. Gollan, P.J.; Tikkanen, M.; Aro, E.M. Photosynthetic light reactions: Integral to chloroplast retrograde signalling. Curr. Opin. Plant Biol. 2015, 27, 180–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Krieger-Liszkay, A.; Fufezan, C.; Trebst, A. Singlet oxygen production in photosystem II and related protection mechanism. Photosynth. Res. 2008, 98, 551–564. [Google Scholar] [CrossRef] [PubMed]
  56. Bobik, K.; Burch-Smith, T.M. Chloroplast signaling within, between and beyond cells. Front. Plant Sci. 2015, 6, 781. [Google Scholar] [CrossRef] [Green Version]
  57. Galvez-Valdivieso, G.; Fryer, M.J.; Lawson, T.; Slattery, K.; Truman, W.; Smirnoff, N.; Asami, T.; Davies, W.J.; Jones, A.M.; Baker, N.R.; et al. The high light response in Arabidopsis involves ABA signaling between vascular and bundle sheath cells. Plant Cell 2009, 21, 2143–2162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Asada, K. The Water-Water Cycle In Chloroplasts: Scavenging of Active Oxygens and Dissipation of Excess Photons. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 601–639. [Google Scholar] [CrossRef] [PubMed]
  59. Michelet, L.; Zaffagnini, M.; Morisse, S.; Sparla, F.; Pérez-Pérez, M.E.; Francia, F.; Danon, A.; Marchand, C.H.; Fermani, S.; Trost, P.; et al. Redox regulation of the Calvin-Benson cycle: Something old, something new. Front. Plant Sci. 2013, 4, 470. [Google Scholar] [CrossRef] [Green Version]
  60. Awad, J.; Stotz, H.U.; Fekete, A.; Krischke, M.; Engert, C.; Havaux, M.; Berger, S.; Mueller, M.J. 2-Cysteine Peroxiredoxins and Thylakoid Ascorbate Peroxidase Create a Water-Water Cycle That Is Essential to Protect the Photosynthetic Apparatus under High Light Stress Conditions. Plant Physiol. 2015, 167, 1592–1603. [Google Scholar] [CrossRef] [Green Version]
  61. Vogel, M.O.; Moore, M.; König, K.; Pecher, P.; Alsharafa, K.; Lee, J.; Dietz, K.-J. Fast Retrograde Signaling in Response to High Light Involves Metabolite Export, Mitogen-Activated Protein Kinase6, and AP2/ERF Transcription Factors in Arabidopsis. Plant Cell 2014, 26, 1151–1165. [Google Scholar] [CrossRef] [Green Version]
  62. Exposito-Rodriguez, M.; Laissue, P.P.; Yvon-Durocher, G.; Smirnoff, N.; Mullineaux, P.M. Photosynthesis-dependent H2O2 transfer from chloroplasts to nuclei provides a high-light signalling mechanism. Nat. Commun. 2017, 8, 49. [Google Scholar] [CrossRef] [Green Version]
  63. Zhao, C.; Wang, Y.; Chan, K.X.; Marchant, D.B.; Franks, P.J.; Randall, D.; Tee, E.E.; Chen, G.; Ramesh, S.; Phua, S.Y.; et al. Evolution of chloroplast retrograde signaling facilitates green plant adaptation to land. Proc. Natl. Acad. Sci. USA 2019, 116, 5015–5020. [Google Scholar] [CrossRef] [PubMed]
  64. Zhao, C.; Haigh, A.M.; Holford, P.; Chen, Z.-H. Roles of Chloroplast Retrograde Signals and Ion Transport in Plant Drought Tolerance. Int. J. Mol. Sci. 2018, 19, 963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ng, S.; Ivanova, A.; Duncan, O.; Law, S.R.; Van Aken, O.; De Clercq, I.; Wang, Y.; Carrie, C.; Xu, L.; Kmiec, B.; et al. A membrane-bound NAC transcription factor, ANAC017, mediates mitochondrial retrograde signaling in Arabidopsis. Plant Cell 2013, 25, 3450–3471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Shapiguzov, A.; Vainonen, J.P.; Hunter, K.; Tossavainen, H.; Tiwari, A.; Järvi, S.; Hellman, M.; Aarabi, F.; Alseekh, S.; Wybouw, B.; et al. Arabidopsis RCD1 coordinates chloroplast and mitochondrial functions through interaction with ANAC transcription factors. eLife 2019, 8, e43284. [Google Scholar] [CrossRef]
  67. op den Camp, R.G.; Przybyla, D.; Ochsenbein, C.; Laloi, C.; Kim, C.; Danon, A.; Wagner, D.; Hideg, E.; Göbel, C.; Feussner, I.; et al. Rapid induction of distinct stress responses after the release of singlet oxygen in Arabidopsis. Plant Cell 2003, 15, 2320–2332. [Google Scholar] [CrossRef] [Green Version]
  68. Vass, I.; Aro, E.-M. Photoinhibition of photosynthetic electron transport, in Primary Processes of Photosynthesis: Priniciples and Apparatus. Part I. Prim. Process. Photosynth. Basic Princ. Appar. 2008, 1, 393–425. [Google Scholar]
  69. Queval, G.; Issakidis-Bourguet, E.; Hoeberichts, F.A.; Vandorpe, M.; Gakière, B.; Vanacker, H.; Miginiac-Maslow, M.; Van Breusegem, F.; Noctor, G. Conditional oxidative stress responses in the Arabidopsis photorespiratory mutant cat2 demonstrate that redox state is a key modulator of daylength-dependent gene expression, and define photoperiod as a crucial factor in the regulation of H2O2-induced cell death. Plant J. 2007, 52, 640–657. [Google Scholar] [CrossRef]
  70. Devireddy, A.R.; Zandalinas, S.I.; Fichman, Y.; Mittler, R. Integration of reactive oxygen species and hormone signaling during abiotic stress. Plant J. 2021, 105, 459–476. [Google Scholar] [CrossRef]
  71. Noctor, G.; Mhamdi, A.; Foyer, C.H. The roles of reactive oxygen metabolism in drought: Not so cut and dried. Plant Physiol. 2014, 164, 1636–1648. [Google Scholar] [CrossRef] [Green Version]
  72. Balsera, M.; Uberegui, E.; Schürmann, P.; Buchanan, B.B. Evolutionary development of redox regulation in chloroplasts. Antioxid. Redox Signal. 2014, 21, 1327–1355. [Google Scholar] [CrossRef]
  73. Gütle, D.D.; Roret, T.; Hecker, A.; Reski, R.; Jacquot, J.-P. Dithiol disulphide exchange in redox regulation of chloroplast enzymes in response to evolutionary and structural constraints. Plant Sci. 2017, 255, 1–11. [Google Scholar] [CrossRef] [PubMed]
  74. Groen, S.C.; Ćalić, I.; Joly-Lopez, Z.; Platts, A.E.; Choi, J.Y.; Natividad, M.; Dorph, K.; Mauck, W.M., 3rd; Bracken, B.; Cabral, C.L.U.; et al. The strength and pattern of natural selection on gene expression in rice. Nature 2020, 578, 572–576. [Google Scholar] [CrossRef] [PubMed]
  75. Kingsolver, J.G.; Hoekstra, H.E.; Hoekstra, J.M.; Berrigan, D.; Vignieri, S.N.; Hill, C.E.; Hoang, A.; Gibert, P.; Beerli, P. The strength of phenotypic selection in natural populations. Am. Nat. 2001, 157, 245–261. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, Z.H.; Chen, G.; Dai, F.; Wang, Y.; Hills, A.; Ruan, Y.L.; Zhang, G.; Franks, P.J.; Nevo, E.; Blatt, M.R. Molecular Evolution of Grass Stomata. Trends Plant Sci. 2017, 22, 124–139. [Google Scholar] [CrossRef] [PubMed]
  77. Deng, F.; Zeng, F.; Chen, G.; Feng, X.; Riaz, A.; Wu, X.; Gao, W.; Wu, F.-b.; Holford, P.; Chen, Z.-H. Metalloid hazards: From plant molecular evolution to mitigation strategies. J. Hazard. Mater. 2020, 409, 124495. [Google Scholar] [CrossRef]
  78. Jiang, W.; Tong, T.; Li, W.; Huang, Z.; Chen, G.; Zeng, F.; Riaz, A.; Amoanimaa-Dede, H.; Pan, R.; Zhang, W.; et al. Molecular Evolution of Plant 14-3-3 Proteins and Function of Hv14-3-3A in Stomatal Regulation and Drought Tolerance. Plant Cell Physiol. 2022. [Google Scholar] [CrossRef]
  79. Adem, G.D.; Chen, G.; Shabala, L.; Chen, Z.-H.; Shabala, S. GORK Channel: A Master Switch of Plant Metabolism? Trends Plant Sci. 2020, 25, 434–445. [Google Scholar] [CrossRef]
  80. Deng, F.; Zeng, F.; Shen, Q.; Abbas, A.; Cheng, J.; Jiang, W.; Chen, G.; Shah, A.N.; Holford, P.; Tanveer, M.; et al. Molecular evolution and functional modification of plant miRNAs with CRISPR. Trends Plant Sci. 2022, 27, 890–907. [Google Scholar] [CrossRef]
  81. Li, Q.; Tong, T.; Jiang, W.; Cheng, J.; Deng, F.; Wu, X.; Chen, Z.H.; Ouyang, Y.; Zeng, F. Highly Conserved Evolution of Aquaporin PIPs and TIPs Confers Their Crucial Contribution to Flowering Process in Plants. Front. Plant Sci. 2021, 12, 761713. [Google Scholar] [CrossRef]
  82. Cai, S.; Huang, Y.; Chen, F.; Zhang, X.; Sessa, E.; Zhao, C.; Marchant, D.B.; Xue, D.; Chen, G.; Dai, F.; et al. Evolution of rapid blue-light response linked to explosive diversification of ferns in angiosperm forests. New Phytol. 2021, 230, 1201–1213. [Google Scholar] [CrossRef]
  83. Tong, T.; Li, Q.; Jiang, W.; Chen, G.; Xue, D.; Deng, F.; Zeng, F.; Chen, Z.H. Molecular Evolution of Calcium Signaling and Transport in Plant Adaptation to Abiotic Stress. Int. J. Mol. Sci. 2021, 22, 12308. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, X.; Jiang, W.; Tong, T.; Chen, G.; Zeng, F.; Jang, S.; Gao, W.; Li, Z.; Mak, M.; Deng, F.; et al. Molecular Interaction and Evolution of Jasmonate Signaling With Transport and Detoxification of Heavy Metals and Metalloids in Plants. Front. Plant Sci. 2021, 12, 665842. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef] [PubMed]
  86. Katoh, K.; Rozewicki, J.; Yamada, K.D. MAFFT online service: Multiple sequence alignment, interactive sequence choice and visualization. Briefings Bioinform. 2019, 20, 1160–1166. [Google Scholar] [CrossRef] [Green Version]
  87. Waterhouse, A.M.; Procter, J.B.; Martin, D.M.A.; Clamp, M.; Barton, G.J. Jalview Version 2—A multiple sequence alignment editor and analysis workbench. Bioinformatics 2009, 25, 1189–1191. [Google Scholar] [CrossRef] [Green Version]
  88. Chibani, K.; Wingsle, G.; Jacquot, J.-P.; Gelhaye, E.; Rouhier, N. Comparative genomic study of the thioredoxin family in photosynthetic organisms with emphasis on Populus trichocarpa. Mol. Plant 2009, 2, 308–322. [Google Scholar] [CrossRef]
  89. Eliyahu, E.; Rog, I.; Inbal, D.; Danon, A. ACHT4-driven oxidation of APS1 attenuates starch synthesis under low light intensity in Arabidopsis plants. Proc. Natl. Acad. Sci. USA 2015, 112, 12876–12881. [Google Scholar] [CrossRef] [Green Version]
  90. Yoshida, K.; Hara, A.; Sugiura, K.; Fukaya, Y.; Hisabori, T. Thioredoxin-like2/2-Cys peroxiredoxin redox cascade supports oxidative thiol modulation in chloroplasts. Proc. Natl. Acad. Sci. USA 2018, 115, E8296–E8304. [Google Scholar] [CrossRef] [Green Version]
  91. Cejudo, F.J.; Ojeda, V.; Delgado-Requerey, V.; González, M.; Pérez-Ruiz, J.M.P. Chloroplast redox regulatory mechanisms in plant adaptation to light and darkness. Front. Plant Sci. 2019, 10, 380. [Google Scholar] [CrossRef] [Green Version]
  92. Yoshida, K.; Uchikoshi, E.; Hara, S.; Hisabori, T. Thioredoxin-like2/2-Cys peroxiredoxin redox cascade acts as oxidative activator of glucose-6-phosphate dehydrogenase in chloroplasts. Biochem. J. 2019, 476, 1781–1790. [Google Scholar] [CrossRef]
  93. Yokochi, Y.; Sugiura, K.; Takemura, K.; Yoshida, K.; Hara, S.; Wakabayashi, K.-I.; Kitao, A.; Hisabori, T. Impact of key residues within chloroplast thioredoxin-f on recognition for reduction and oxidation of target proteins. J. Biol. Chem. 2019, 294, 17437–17450. [Google Scholar] [CrossRef] [PubMed]
  94. Yoshida, K.; Yokochi, Y.; Hisabori, T. New light on chloroplast redox regulation: Molecular mechanism of protein thiol oxidation. Front. Plant Sci. 2019, 10, 1534. [Google Scholar] [CrossRef] [PubMed]
  95. Yoshida, K.; Hisabori, T.J.P.; Physiology, C. Biochemical basis for redox regulation of chloroplast-localized phosphofructokinase from Arabidopsis thaliana. Plant Cell Physiol. 2021, 62, 401–410. [Google Scholar] [CrossRef] [PubMed]
  96. Mestres-Ortega, D.; Meyer, Y. The Arabidopsis thaliana genome encodes at least four thioredoxins m and a new prokaryotic-like thioredoxin. Gene 1999, 240, 307–316. [Google Scholar] [CrossRef]
  97. Arsova, B.; Hoja, U.; Wimmelbacher, M.; Greiner, E.; Ustün, S.; Melzer, M.; Petersen, K.; Lein, W.; Börnke, F. Plastidial thioredoxin z interacts with two fructokinase-like proteins in a thiol-dependent manner: Evidence for an essential role in chloroplast development in Arabidopsis and Nicotiana benthamiana. Plant Cell 2010, 22, 1498–1515. [Google Scholar] [CrossRef] [Green Version]
  98. Schröter, Y.; Steiner, S.; Matthäi, K.; Pfannschmidt, T. Analysis of oligomeric protein complexes in the chloroplast sub-proteome of nucleic acid-binding proteins from mustard reveals potential redox regulators of plastid gene expression. Proteomics 2010, 10, 2191–2204. [Google Scholar] [CrossRef]
  99. Gelhaye, E.; Rouhier, N.; Navrot, N.; Jacquot, J.-P. The plant thioredoxin system. Experientia 2005, 62, 24–35. [Google Scholar] [CrossRef]
  100. Ojeda, V.; Pérez-Ruiz, J.M.; Cejudo, F.J. 2-Cys peroxiredoxins participate in the oxidation of chloroplast enzymes in the dark. Mol. Plant 2018, 11, 1377–1388. [Google Scholar] [CrossRef] [Green Version]
  101. Dangoor, I.; Peled-Zehavi, H.; Levitan, A.; Pasand, O.; Danon, A. A small family of chloroplast atypical thioredoxins. Plant Physiol. 2009, 149, 1240–1250. [Google Scholar] [CrossRef] [Green Version]
  102. Yokochi, Y.; Fukushi, Y.; Wakabayashi, K.I.; Yoshida, K.; Hisabori, T. Oxidative regulation of chloroplast enzymes by thioredoxin and thioredoxin-like proteins in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2021, 118, e2114952118. [Google Scholar] [CrossRef]
  103. Belin, C.; Bashandy, T.; Cela, J.; Delorme-Hinoux, V.; Riondet, C.; Reichheld, J. A comprehensive study of thiol reduction gene expression under stress conditions in A rabidopsis thaliana. Plant Cell Environ. 2015, 38, 299–314. [Google Scholar] [CrossRef] [PubMed]
  104. Portis, A.R.; Li, C.; Wang, D.; Salvucci, M.E. Regulation of Rubisco activase and its interaction with Rubisco. J. Exp. Bot. 2007, 59, 1597–1604. [Google Scholar] [CrossRef]
  105. Yin, Z.; Meng, F.; Song, H.; Wang, X.; Xu, X.; Yu, D. Expression Quantitative Trait Loci Analysis of Two Genes Encoding Rubisco Activase in Soybean. Plant Physiol. 2009, 152, 1625–1637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Salvucci, M.E.; Werneke, J.M.; Ogren, W.L.; Portis, A.R., Jr. Purification and Species Distribution of Rubisco Activase. Plant Physiol. 1987, 84, 930–936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Petersen, J.; Teich, R.; Becker, B.; Cerff, R.; Brinkmann, H. The GapA/B Gene Duplication Marks the Origin of Streptophyta (Charophytes and Land Plants). Mol. Biol. Evol. 2006, 23, 1109–1118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Robbens, S.; Petersen, J.; Brinkmann, H.; Rouzé, P.; Van de Peer, Y. Unique regulation of the Calvin cycle in the ultrasmall green alga Ostreococcus. J. Mol. Evol. 2007, 64, 601–604. [Google Scholar] [CrossRef] [PubMed]
  109. Liaud, M.F.; Brandt, U.; Scherzinger, M.; Cerff, R. Evolutionary origin of cryptomonad microalgae: Two novel chloroplast/cytosol-specific GAPDH genes as potential markers of ancestral endosymbiont and host cell components. J. Mol. Evol. 1997, 44 (Suppl. S1), S28–S37. [Google Scholar] [CrossRef]
  110. Pedersen, T.A.; Kirk, M.; Bassham, J.A. Light-Dark Transients in Levels of Intermediate Compounds during Photosynthesis in Air-Adapted Chlorella. Physiol. Plant. 1966, 19, 219–231. [Google Scholar] [CrossRef] [Green Version]
  111. Champigny, M.-L.; Bismuth, E. Role of Photosynthetic Electron Transfer in Light Activation of Calvin Cycle Enzymes. Physiol. Plant. 1976, 36, 95–100. [Google Scholar] [CrossRef]
  112. Avron, M.; Gibbs, M. Properties of phosphoribulokinase of whole chloroplasts. Plant Physiol. 1974, 53, 136–139. [Google Scholar] [CrossRef]
  113. Harrison, D.H.T.; Runquist, J.A.; Holub, A.; Miziorko, H.M. The Crystal Structure of Phosphoribulokinase from Rhodobacter sphaeroides Reveals a Fold Similar to That of Adenylate Kinase. Biochemistry 1998, 37, 5074–5085. [Google Scholar] [CrossRef] [PubMed]
  114. Marri, L.; Zaffagnini, M.; Collin, V.; Issakidis-Bourguet, E.; Lemaire, S.D.; Pupillo, P.; Sparla, F.; Miginiac-Maslow, M.; Trost, P. Prompt and easy activation by specific thioredoxins of calvin cycle enzymes of Arabidopsis thaliana associated in the GAPDH/CP12/PRK supramolecular complex. Mol. Plant 2009, 2, 259–269. [Google Scholar] [CrossRef] [PubMed]
  115. Tamoi, M.; Ishikawa, T.; Takeda, T.; Shigeoka, S. Molecular characterization and resistance to hydrogen peroxide of two fructose-1,6-bisphosphatases from Synechococcus PCC 7942. Arch. Biochem. Biophys. 1996, 334, 27–36. [Google Scholar] [CrossRef]
  116. Raines, C.A.; Lloyd, J.C.; Willingham, N.M.; Potts, S.; Dyer, T.A. cDNA and gene sequences of wheat chloroplast sedoheptulose-1,7-bisphosphatase reveal homology with fructose-1,6-bisphosphatases. Eur. J. Biochem. 1992, 205, 1053–1059. [Google Scholar] [CrossRef] [PubMed]
  117. Foyer, C.H.; Bloom, A.J.; Queval, G.; Noctor, G. Photorespiratory Metabolism: Genes, Mutants, Energetics, and Redox Signaling. Annu. Rev. Plant Biol. 2009, 60, 455–484. [Google Scholar] [CrossRef]
  118. Miginiac-Maslow, M.; Lancelin, J.M. Intrasteric inhibition in redox signalling: Light activation of NADP-malate dehydrogenase. Photosynth. Res. 2002, 72, 1–12. [Google Scholar] [CrossRef]
  119. Lemaire, S.D.; Quesada, A.; Merchan, F.; Corral, J.M.; Igeno, M.I.; Keryer, E.; Issakidis-Bourguet, E.; Hirasawa, M.; Knaff, D.B.; Miginiac-Maslow, M. NADP-Malate Dehydrogenase from Unicellular Green Alga Chlamydomonas reinhardtii. A First Step toward Redox Regulation? Plant Physiol. 2005, 137, 514–521. [Google Scholar] [CrossRef] [Green Version]
  120. Latifi, A.; Ruiz, M.; Zhang, C.-C. Oxidative stress in cyanobacteria. FEMS Microbiol. Rev. 2009, 33, 258–278. [Google Scholar] [CrossRef] [Green Version]
  121. Diaz, J.M.; Plummer, S. Production of extracellular reactive oxygen species by phytoplankton: Past and future directions. J. Plankton Res. 2018, 40, 655–666. [Google Scholar] [CrossRef] [Green Version]
  122. Bernroitner, M.; Zamocky, M.; Furtmüller, P.G.; Peschek, G.A.; Obinger, C. Occurrence, phylogeny, structure, and function of catalases and peroxidases in cyanobacteria. J. Exp. Bot. 2009, 60, 423–440. [Google Scholar] [CrossRef] [Green Version]
  123. Rose, A.J. The influence of extracellular superoxide on iron redox chemistry and bioavailability to aquatic microorganisms. Front. Microbiol. 2012, 3, 124. [Google Scholar] [CrossRef] [Green Version]
  124. Vassiliev, I.R.; Ronan, M.T.; Hauska, G.; Golbeck, J.H. The bound electron acceptors in green sulfur bacteria: Resolution of the g-tensor for the FX iron-sulfur cluster in Chlorobium tepidum. Biophys. J. 2000, 78, 3160–3169. [Google Scholar] [CrossRef]
  125. Wallace, M.A.; Liou, L.-L.; Martins, J.; Clement, M.H.; Bailey, S.; Longo, V.D.; Valentine, J.S.; Gralla, E.B. Superoxide inhibits 4Fe-4S cluster enzymes involved in amino acid biosynthesis: Cross-compartment protection by CuZn-superoxide dismutase. J. Biol. Chem. 2004, 279, 32055–32062. [Google Scholar] [CrossRef] [Green Version]
  126. Sheng, Y.; Abreu, I.A.; Cabelli, D.E.; Maroney, M.J.; Miller, A.-F.; Teixeira, M.; Valentine, J.S. Superoxide dismutases and superoxide reductases. Chem. Rev. 2014, 114, 3854–3918. [Google Scholar] [CrossRef]
  127. Yokochi, Y.; Yoshida, K.; Hahn, F.; Miyagi, A.; Wakabayashi, K.-I.; Kawai-Yamada, M.; Weber, A.P.M.; Hisabori, T. Redox regulation of NADP-malate dehydrogenase is vital for land plants under fluctuating light environment. Proc. Natl. Acad. Sci. USA 2021, 118, e2016903118. [Google Scholar] [CrossRef] [PubMed]
  128. Genty, B.; Harbinson, J. Regulation of Light Utilization for Photosynthetic Electron Transport. In Photosynthesis and the Environment; Baker, N.R., Ed.; Springer: Dordrecht, The Netherlands, 1996; pp. 67–99. [Google Scholar]
  129. Buchanan, B.B. Role of light in the regulation of chloroplast enzymes. Annu. Rev. Plant Physiol. 1980, 31, 341–374. [Google Scholar] [CrossRef]
  130. Chaux, F.; Peltier, G.; Johnson, X. A security network in PSI photoprotection: Regulation of photosynthetic control, NPQ and O2 photoreduction by cyclic electron flow. Front. Plant Sci. 2015, 6, 875. [Google Scholar] [CrossRef] [Green Version]
  131. Sonoike, K. Photoinhibition of photosystem I. Physiol. Plant 2011, 142, 56–64. [Google Scholar] [CrossRef] [PubMed]
  132. Kirilovsky, D.; Kerfeld, C.A. The orange carotenoid protein in photoprotection of photosystem II in cyanobacteria. Biochim. Biophys. Acta 2012, 1817, 158–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Ruban, A.V. Nonphotochemical Chlorophyll Fluorescence Quenching: Mechanism and Effectiveness in Protecting Plants from Photodamage. Plant Physiol. 2016, 170, 1903–1916. [Google Scholar] [CrossRef] [Green Version]
  134. Li, Z.; Wakao, S.; Fischer, B.B.; Niyogi, K.K. Sensing and Responding to Excess Light. Annu. Rev. Plant Biol. 2009, 60, 239–260. [Google Scholar] [CrossRef] [PubMed]
  135. Pinnola, A.; Bassi, R. Molecular mechanisms involved in plant photoprotection. Biochem. Soc. Trans. 2018, 46, 467–482. [Google Scholar] [CrossRef] [PubMed]
  136. Munekage, Y.N.; Genty, B.; Peltier, G. Effect of PGR5 Impairment on Photosynthesis and Growth in Arabidopsis thaliana. Plant Cell Physiol. 2008, 49, 1688–1698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Rumeau, D.; Peltier, G.; Cournac, L. Chlororespiration and cyclic electron flow around PSI during photosynthesis and plant stress response. Plant Cell Environ. 2007, 30, 1041–1051. [Google Scholar] [CrossRef]
  138. Konno, H.; Nakane, T.; Yoshida, M.; Ueoka-Nakanishi, H.; Hara, S.; Hisabori, T.J.P.; Physiology, C. Thiol modulation of the chloroplast ATP synthase is dependent on the energization of thylakoid membranes. Plant Cell Physiol. 2012, 53, 626–634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Yoshida, K.; Matsuoka, Y.; Hara, S.; Konno, H.; Hisabori, T. Distinct redox behaviors of chloroplast thiol enzymes and their relationships with photosynthetic electron transport in Arabidopsis thaliana. Plant Cell Physiol. 2014, 55, 1415–1425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Wolosiuk, R.A.; Buchanan, B.B. Thioredoxin and glutathione regulate photosynthesis in chloroplasts. Nature 1977, 266, 565–567. [Google Scholar] [CrossRef]
  141. Serrato, A.J.; Pérez-Ruiz, J.M.; Spínola, M.a.C.; Cejudo, F.J. A novel NADPH thioredoxin reductase, localized in the chloroplast, which deficiency causes hypersensitivity to abiotic stress in Arabidopsis thaliana. J. Biol. Chem. 2004, 279, 43821–43827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Naranjo, B.; Mignée, C.; Krieger-Liszkay, A.; Hornero-Méndez, D.; Gallardo-Guerrero, L.; Cejudo, F.J.; Lindahl, M. The chloroplast NADPH thioredoxin reductase C, NTRC, controls non-photochemical quenching of light energy and photosynthetic electron transport in Arabidopsis. Plant Cell Environ. 2016, 39, 804–822. [Google Scholar] [CrossRef] [Green Version]
  143. Yoshida, K.; Hisabori, T. Two distinct redox cascades cooperatively regulate chloroplast functions and sustain plant viability. Proc. Natl. Acad. Sci. USA 2016, 113, E3967–E3976. [Google Scholar] [CrossRef] [Green Version]
  144. Pérez-Ruiz, J.M.; Naranjo, B.; Ojeda, V.; Guinea, M.; Cejudo, F.J. NTRC-dependent redox balance of 2-Cys peroxiredoxins is needed for optimal function of the photosynthetic apparatus. Proc. Natl. Acad. Sci. USA 2017, 114, 12069–12074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Thormählen, I.; Meitzel, T.; Groysman, J.; Öchsner, A.B.; von Roepenack-Lahaye, E.; Naranjo, B.; Cejudo, F.J.; Geigenberger, P. Thioredoxin f 1 and NADPH-dependent thioredoxin reductase C have overlapping functions in regulating photosynthetic metabolism and plant growth in response to varying light conditions. Plant Physiol. 2015, 169, 1766–1786. [Google Scholar] [CrossRef] [PubMed]
  146. Nikkanen, L.; Toivola, J.; Rintamäki, E. Crosstalk between chloroplast thioredoxin systems in regulation of photosynthesis. Plant Cell Environ. 2016, 39, 1691–1705. [Google Scholar] [CrossRef] [Green Version]
  147. Nikkanen, L.; Rintamäki, E. Chloroplast thioredoxin systems dynamically regulate photosynthesis in plants. Biochem. J. 2019, 476, 1159–1172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Kinoshita, H.; Nagasaki, J.; Yoshikawa, N.; Yamamoto, A.; Takito, S.; Kawasaki, M.; Sugiyama, T.; Miyake, H.; Weber, A.P.; Taniguchi, M. The chloroplastic 2-oxoglutarate/malate transporter has dual function as the malate valve and in carbon/nitrogen metabolism. Plant J. 2011, 65, 15–26. [Google Scholar] [CrossRef]
  149. Selinski, J.; Scheibe, R. Malate valves: Old shuttles with new perspectives. Plant Biol. 2019, 21 (Suppl. 1), 21–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Scheibe, R. NADP+-malate dehydrogenase in C3-plants: Regulation and role of a light-activated enzyme. Physiol. Plant. 1987, 71, 393–400. [Google Scholar] [CrossRef]
  151. Berkemeyer, M.; Scheibe, R.; Ocheretina, O. A Novel, Non-redox-regulated NAD-dependent Malate Dehydrogenase from Chloroplasts of Arabidopsis thalianaL. J. Biol. Chem. 1998, 273, 27927–27933. [Google Scholar] [CrossRef] [Green Version]
  152. Turner, W.L.; Waller, J.C.; Vanderbeld, B.; Snedden, W.A. Cloning and characterization of two NAD kinases from Arabidopsis. Identification of a calmodulin binding isoform. Plant Physiol. 2004, 135, 1243–1255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Correll, C.C.; Ludwig, M.L.; Bruns, C.M.; Karplus, P.A. Structural prototypes for an extended family of flavoprotein reductases: Comparison of phthalate dioxygenase reductase with ferredoxin reductase and ferredoxin. Protein Sci. A Publ. Protein Soc. 1993, 2, 2112–2133. [Google Scholar] [CrossRef] [Green Version]
  154. Piubelli, L.; Aliverti, A.; Arakaki, A.K.; Carrillo, N.; Ceccarelli, E.A.; Karplus, P.A.; Zanetti, G. Competition between C-terminal tyrosine and nicotinamide modulates pyridine nucleotide affinity and specificity in plant ferredoxin-NADP(+) reductase. J. Biol. Chem. 2000, 275, 10472–10476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Medina, M.; Luquita, A.; Tejero, J.; Hermoso, J.; Mayoral, T.; Sanz-Aparicio, J.; Grever, K.; Gomez-Moreno, C. Probing the determinants of coenzyme specificity in ferredoxin-NADP+ reductase by site-directed mutagenesis. J. Biol. Chem. 2001, 276, 11902–11912. [Google Scholar] [CrossRef]
  156. Barber, J. The intact chloroplast. FEBS Lett. 1976, 97, 171–214. [Google Scholar]
  157. Balsera, M.; Uberegui, E.; Susanti, D.; Schmitz, R.A.; Mukhopadhyay, B.; Schürmann, P.; Buchanan, B.B. Ferredoxin:thioredoxin reductase (FTR) links the regulation of oxygenic photosynthesis to deeply rooted bacteria. Planta 2013, 237, 619–635. [Google Scholar] [CrossRef] [PubMed]
  158. Gütle, D.D.; Roret, T.; Müller, S.J.; Couturier, J.; Lemaire, S.D.; Hecker, A.; Dhalleine, T.; Buchanan, B.B.; Reski, R.; Einsle, O.; et al. Chloroplast FBPase and SBPase are thioredoxin-linked enzymes with similar architecture but different evolutionary histories. Proc. Natl. Acad. Sci. USA 2016, 113, 6779–6784. [Google Scholar] [CrossRef] [Green Version]
  159. Ross, S.A.; Zhang, M.X.; Selman, B.R. Role of the Chlamydomonas reinhardtii Coupling Factor 1 γ-Subunit Cysteine Bridge in the Regulation of ATP Synthase(∗). J. Biol. Chem. 1995, 270, 9813–9818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Wu, G.; Ort, D.R. Mutation in the cysteine bridge domain of the gamma-subunit affects light regulation of the ATP synthase but not photosynthesis or growth in Arabidopsis. Photosynth. Res. 2008, 97, 185–193. [Google Scholar] [CrossRef]
  161. Miginiac-Maslow, M.; Issakidis, E.; Lemaire, M.; Ruelland, E.; Jacquot, J.-P.; Decottignies, P. Light-dependent Activation of NADP-Malate Dehydrogenase: A Complex Process. Funct. Plant Biol. 1997, 24, 529–542. [Google Scholar] [CrossRef]
  162. Hebbelmann, I.; Selinski, J.; Wehmeyer, C.; Goss, T.; Voss, I.; Mulo, P.; Kangasjärvi, S.; Aro, E.M.; Oelze, M.L.; Dietz, K.J.; et al. Multiple strategies to prevent oxidative stress in Arabidopsis plants lacking the malate valve enzyme NADP-malate dehydrogenase. J. Exp. Bot. 2012, 63, 1445–1459. [Google Scholar] [CrossRef] [Green Version]
  163. Scheibe, R. Malate valves to balance cellular energy supply. Physiol. Plant 2004, 120, 21–26. [Google Scholar] [CrossRef] [PubMed]
  164. Knuesting, J.; Scheibe, R. Small Molecules Govern Thiol Redox Switches. Trends Plant Sci. 2018, 23, 769–782. [Google Scholar] [CrossRef]
  165. Heckmann, D.; Schulze, S.; Denton, A.; Gowik, U.; Westhoff, P.; Weber, A.P.; Lercher, M.J. Predicting C4 photosynthesis evolution: Modular, individually adaptive steps on a Mount Fuji fitness landscape. Cell 2013, 153, 1579–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Schlüter, U.; Bräutigam, A.; Gowik, U.; Melzer, M.; Christin, P.-A.; Kurz, S.; Mettler-Altmann, T.; Weber, A.P. Photosynthesis in C3–C4 intermediate Moricandia species. J. Exp. Bot. 2016, 68, 191–206. [Google Scholar] [CrossRef] [PubMed]
  167. Bräutigam, A.; Schliesky, S.; Külahoglu, C.; Osborne, C.P.; Weber, A.P. Towards an integrative model of C4 photosynthetic subtypes: Insights from comparative transcriptome analysis of NAD-ME, NADP-ME, and PEP-CK C4 species. J. Exp. Bot. 2014, 65, 3579–3593. [Google Scholar] [CrossRef] [Green Version]
  168. Bräutigam, A.; Kajala, K.; Wullenweber, J.; Sommer, M.; Gagneul, D.; Weber, K.L.; Carr, K.M.; Gowik, U.; Maß, J.; Lercher, M.J.; et al. An mRNA Blueprint for C4 Photosynthesis Derived from Comparative Transcriptomics of Closely Related C3 and C4 Species. Plant Physiol. 2010, 155, 142–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Nakamura, N.; Iwano, M.; Havaux, M.; Yokota, A.; Munekage, Y.N. Promotion of cyclic electron transport around photosystem I during the evolution of NADP-malic enzyme-type C4 photosynthesis in the genus Flaveria. New Phytol. 2013, 199, 832–842. [Google Scholar] [CrossRef]
  170. Shikanai, T. Central role of cyclic electron transport around photosystem I in the regulation of photosynthesis. Curr. Opin. Biotechnol. 2014, 26, 25–30. [Google Scholar] [CrossRef] [PubMed]
  171. Suorsa, M.; Rossi, F.; Tadini, L.; Labs, M.; Colombo, M.; Jahns, P.; Kater, M.M.; Leister, D.; Finazzi, G.; Aro, E.M.; et al. PGR5-PGRL1-Dependent Cyclic Electron Transport Modulates Linear Electron Transport Rate in Arabidopsis thaliana. Mol. Plant 2016, 9, 271–288. [Google Scholar] [CrossRef] [Green Version]
  172. Krieger-Liszkay, A.; Feilke, K. The Dual Role of the Plastid Terminal Oxidase PTOX: Between a Protective and a Pro-oxidant Function. Front. Plant Sci. 2015, 6, 1147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Asada, K. Production and Scavenging of Reactive Oxygen Species in Chloroplasts and Their Functions. Plant Physiol. 2006, 141, 391–396. [Google Scholar] [CrossRef] [Green Version]
  174. Dietz, K.J. Thiol-Based Peroxidases and Ascorbate Peroxidases: Why Plants Rely on Multiple Peroxidase Systems in the Photosynthesizing Chloroplast? Mol. Cells 2016, 39, 20–25. [Google Scholar] [CrossRef] [Green Version]
  175. Lee, S.D.; Kim, S.J.; Jung, S.I.; Son, K.-C.; Kays, S.J. Diurnal CO2 assimilation patterns in nine species of CAM-type succulent plants. HortScience 2006, 41, 1373–1376. [Google Scholar] [CrossRef] [Green Version]
  176. Cushman, J. CAM Plants. In Encyclopedia of Applied Plant Sciences, 2nd ed.; Thomas, B., Murray, B.G., Murphy, D.J., Eds.; Academic Press: Oxford, UK, 2017; pp. 60–77. [Google Scholar]
  177. Turkan, I.; Uzilday, B.; Dietz, K.J.; Bräutigam, A.; Ozgur, R. Reactive oxygen species and redox regulation in mesophyll and bundle sheath cells of C4 plants. J. Exp. Bot. 2018, 69, 3321–3331. [Google Scholar] [CrossRef] [PubMed]
  178. Yang, Y.-J.; Zhang, S.-B.; Huang, W. Photosynthetic regulation under fluctuating light in young and mature leaves of the CAM plant Bryophyllum pinnatum. Biochim. Biophys. Acta 2019, 1860, 469–477. [Google Scholar] [CrossRef]
  179. Yang, Y.-J.; Tan, S.-L.; Huang, J.-L.; Zhang, S.-B.; Huang, W.J.E.; Botany, E. The water-water cycle facilitates photosynthetic regulation under fluctuating light in the epiphytic orchid Dendrobium officinale. Environ. Exp. Bot. 2020, 180, 104238. [Google Scholar] [CrossRef]
  180. Sun, H.; Shi, Q.; Zhang, S.-B.; Huang, W.J.P. Coordination of cyclic electron flow and water–water cycle facilitates photoprotection under fluctuating light and temperature stress in the epiphytic orchid Dendrobium officinale. Plants 2021, 10, 606. [Google Scholar] [CrossRef]
  181. Wang, H.; Wang, X.-Q.; Zeng, Z.-L.; Yu, H.; Huang, W. Photosynthesis under fluctuating light in the CAM plant Vanilla planifolia. Plant Sci. 2022, 317, 111207. [Google Scholar] [CrossRef] [PubMed]
  182. Jacquot, J.-P.; Eklund, H.; Rouhier, N.; Schürmann, P. Structural and evolutionary aspects of thioredoxin reductases in photosynthetic organisms. Trends Plant Sci. 2009, 14, 336–343. [Google Scholar] [CrossRef] [PubMed]
  183. Schürmann, P.; Buchanan, B.B. The ferredoxin/thioredoxin system of oxygenic photosynthesis. Antioxid. Redox Signal. 2008, 10, 1235–1274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Foyer, C.H. Redox homeostasis: Opening up ascorbate transport. Nat. Plants 2015, 1, 14012. [Google Scholar] [CrossRef] [PubMed]
  185. Dietz, K.-J.; Jacob, S.; Oelze, M.-L.; Laxa, M.; Tognetti, V.; de Miranda, S.M.N.; Baier, M.; Finkemeier, I. The function of peroxiredoxins in plant organelle redox metabolism. J. Exp. Bot. 2006, 57, 1697–1709. [Google Scholar] [CrossRef] [PubMed]
  186. Horemans, N.; Foyer, C.H.; Potters, G.; Asard, H. Ascorbate function and associated transport systems in plants. Plant Physiol. Biochem. 2000, 38, 531–540. [Google Scholar] [CrossRef]
  187. Beck, E.; Burkert, A.; Hofmann, M. Uptake of L-ascorbate by intact spinach chloroplasts. Plant Physiol. 1983, 73, 41–45. [Google Scholar] [CrossRef] [PubMed]
  188. de Koning, H.; Diallinas, G. Nucleobase transporters. Mol. Membr. Biol. 2000, 17, 75–94. [Google Scholar] [PubMed]
  189. Bürzle, M.; Suzuki, Y.; Ackermann, D.; Miyazaki, H.; Maeda, N.; Clémençon, B.; Burrier, R.; Hediger, M.A. The sodium-dependent ascorbic acid transporter family SLC23. Mol. Asp. Med. 2013, 34, 436–454. [Google Scholar] [CrossRef]
  190. Diallinas, G.; Valdez, J.; Sophianopoulou, V.; Rosa, A.; Scazzocchio, C. Chimeric purine transporters of Aspergillus nidulans define a domain critical for function and specificity conserved in bacterial, plant and metazoan homologues. EMBO J. 1998, 17, 3827–3837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Argyrou, E.; Sophianopoulou, V.; Schultes, N.; Diallinas, G. Functional characterization of a maize purine transporter by expression in Aspergillus nidulans. Plant Cell 2001, 13, 953–964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Schultes, N.P.; Brutnell, T.P.; Allen, A.; Dellaporta, S.L.; Nelson, T.; Chen, J. Leaf permease1 gene of maize is required for chloroplast development. Plant Cell 1996, 8, 463–475. [Google Scholar] [PubMed] [Green Version]
  193. Miyaji, T.; Kuromori, T.; Takeuchi, Y.; Yamaji, N.; Yokosho, K.; Shimazawa, A.; Sugimoto, E.; Omote, H.; Ma, J.F.; Shinozaki, K.; et al. AtPHT4;4 is a chloroplast-localized ascorbate transporter in Arabidopsis. Nat. Commun. 2015, 6, 5928. [Google Scholar] [CrossRef] [Green Version]
  194. Zhang, J.; Kirkham, M.B. Antioxidant responses to drought in sunflower and sorghum seedlings. New Phytol. 1996, 132, 361–373. [Google Scholar] [CrossRef]
  195. Stepien, P.; Klobus, G. Antioxidant defense in the leaves of C3 and C4 plants under salinity stress. Physiol. Plant. 2005, 125, 31–40. [Google Scholar] [CrossRef]
  196. Nayyar, H.; Gupta, D. Differential sensitivity of C3 and C4 plants to water deficit stress: Association with oxidative stress and antioxidants. Environ. Exp. Bot. 2006, 58, 106–113. [Google Scholar] [CrossRef]
  197. Uzilday, B.; Turkan, I.; Sekmen, A.H.; Ozgur, R.; Karakaya, H.C. Comparison of ROS formation and antioxidant enzymes in Cleome gynandra (C4) and Cleome spinosa (C3) under drought stress. Plant Sci. 2012, 182, 59–70. [Google Scholar] [CrossRef] [PubMed]
  198. Uzilday, B.; Turkan, I.; Ozgur, R.; Sekmen, A.H. Strategies of ROS regulation and antioxidant defense during transition from C3 to C4 photosynthesis in the genus Flaveria under PEG-induced osmotic stress. J. Plant Physiol. 2014, 171, 65–75. [Google Scholar] [CrossRef] [PubMed]
  199. Uzilday, B.; Ozgur, R.; Yalcinkaya, T.; Turkan, I.; Sekmen, A.H. Changes in redox regulation during transition from C3 to single cell C4 photosynthesis in Bienertia sinuspersici. J. Plant Physiol. 2018, 220, 1–10. [Google Scholar] [CrossRef] [PubMed]
  200. Lara, M.V.; Offermann, S.; Smith, M.; Okita, T.W.; Andreo, C.S.; Edwards, G.E. Leaf development in the single-cell C4 system in Bienertia sinuspersici: Expression of genes and peptide levels for C4 metabolism in relation to chlorenchyma structure under different light conditions. Plant Physiol. 2008, 148, 593–610. [Google Scholar] [CrossRef] [Green Version]
  201. Baier, M.; Noctor, G.; Foyer, C.H.; Dietz, K.J. Antisense suppression of 2-cysteine peroxiredoxin in Arabidopsis specifically enhances the activities and expression of enzymes associated with ascorbate metabolism but not glutathione metabolism. Plant Physiol. 2000, 124, 823–832. [Google Scholar] [CrossRef] [Green Version]
  202. Kangasjärvi, S.; Lepistö, A.; Hännikäinen, K.; Piippo, M.; Luomala, E.M.; Aro, E.M.; Rintamäki, E. Diverse roles for chloroplast stromal and thylakoid-bound ascorbate peroxidases in plant stress responses. Biochem. J. 2008, 412, 275–285. [Google Scholar] [CrossRef] [Green Version]
  203. Aro, E.M.; Virgin, I.; Andersson, B. Photoinhibition of Photosystem II. Inactivation, protein damage and turnover. Biochim. Biophys. Acta 1993, 1143, 113–134. [Google Scholar] [CrossRef]
  204. Chen, J.H.; Chen, S.T.; He, N.Y.; Wang, Q.L.; Zhao, Y.; Gao, W.; Guo, F.Q. Nuclear-encoded synthesis of the D1 subunit of photosystem II increases photosynthetic efficiency and crop yield. Nat. Plants 2020, 6, 570–580. [Google Scholar] [CrossRef]
  205. Theis, J.; Schroda, M. Revisiting the photosystem II repair cycle. Plant Signal. Behav. 2016, 11, e1218587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Johnson, M.; Bueno, M.L.P.; Zia, A.; Horton, P.; Ruban, A.V. The Zeaxanthin-Independent and Zeaxanthin-Dependent qE Components of Nonphotochemical Quenching Involve Common Conformational Changes within the Photosystem II Antenna in Arabidopsis. Plant Physiol. 2008, 149, 1061–1075. [Google Scholar] [CrossRef] [Green Version]
  207. Jahns, P.; Holzwarth, A.R. The role of the xanthophyll cycle and of lutein in photoprotection of photosystem II. Biochim. Biophys. Acta 2012, 1817, 182–193. [Google Scholar] [CrossRef]
  208. Hieber, A.D.; Bugos, R.C.; Yamamoto, H.Y. Plant lipocalins: Violaxanthin de-epoxidase and zeaxanthin epoxidase. Biochim. Biophys. Acta 2000, 1482, 84–91. [Google Scholar] [CrossRef]
  209. Hoang, M.H.; Kim, H.-S.; Zulfugarov, I.S.; Lee, C.-H. Down-Regulation of Zeaxanthin Epoxidation in Vascular Plant Leaves Under Normal and Photooxidative Stress Conditions. J. Plant Biol. 2020, 63, 331–336. [Google Scholar] [CrossRef]
  210. Zhu, X.G.; Ort, D.R.; Whitmarsh, J.; Long, S.P. The slow reversibility of photosystem II thermal energy dissipation on transfer from high to low light may cause large losses in carbon gain by crop canopies: A theoretical analysis. J. Exp. Bot. 2004, 55, 1167–1175. [Google Scholar] [CrossRef]
  211. Murchie, E.H.; Niyogi, K.K. Manipulation of photoprotection to improve plant photosynthesis. Plant Physiol. 2011, 155, 86–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Kromdijk, J.; Głowacka, K.; Leonelli, L.; Gabilly, S.T.; Iwai, M.; Niyogi, K.K.; Long, S.P. Improving photosynthesis and crop productivity by accelerating recovery from photoprotection. Science 2016, 354, 857–861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Garcia-Molina, A.; Leister, D. Accelerated relaxation of photoprotection impairs biomass accumulation in Arabidopsis. Nat. Plants 2020, 6, 9–12. [Google Scholar] [CrossRef]
  214. Goss, R.; Lepetit, B. Biodiversity of NPQ. J. Plant Physiol. 2015, 172, 13–32. [Google Scholar] [CrossRef] [PubMed]
  215. Schierenbeck, L.; Ries, D.; Rogge, K.; Grewe, S.; Weisshaar, B.; Kruse, O. Fast forward genetics to identify mutations causing a high light tolerant phenotype in Chlamydomonas reinhardtii by whole-genome-sequencing. BMC Genom. 2015, 16, 57. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The oxidative stress signaling in plant cell compartments. ROS, such as O2•− and H2O2 are produced by NADPH oxidases specifically RBOHs at the apoplast, accumulating in chloroplast, mitochondria, peroxisomes, and nuclei. Accumulation of H2O2 in the presence of F2+ form hydroxyl radicals and initiates oxidative stress that results in distorting the structure of lipids, proteins, and DNA. Later, the integration of enzymes, such as SODs, APXs, DHAR, and CAT in cell organelles maintains ROS levels and gene expression.
Figure 1. The oxidative stress signaling in plant cell compartments. ROS, such as O2•− and H2O2 are produced by NADPH oxidases specifically RBOHs at the apoplast, accumulating in chloroplast, mitochondria, peroxisomes, and nuclei. Accumulation of H2O2 in the presence of F2+ form hydroxyl radicals and initiates oxidative stress that results in distorting the structure of lipids, proteins, and DNA. Later, the integration of enzymes, such as SODs, APXs, DHAR, and CAT in cell organelles maintains ROS levels and gene expression.
Antioxidants 11 02085 g001
Figure 2. Comparative molecular evolution of key photosynthetic redox regulators and photosynthesis protein families in algae and land plants candidate proteins were extracted using BLASTP (NCBI), and the all-against-all BLASTP search with satisfied E-value of 10−10 and query coverage of 50%. Heat map was generated using TBtools [85] from the data of seven photosynthesis-related proteins and four redox-related families. Black squares indicate the proteins that do not satisfy the selection criteria. ACHTs: atypical Cys His-rich, CF1-y: gamma subunit of ATP synthase; CP12; FBPase: fructose-1,6-bisphosphatase; GADPH: glyceraldehyde-3-phosphate dehydrogenase; NADP-MDH: NADP+ dependent malate dehydrogenase; RCA: ribulose-1,5-bisphosphate carboxylase/oxygenase activase; SBPase: sedoheptulose-1,7-bisphosphatase; TrxLs: Trx-like.
Figure 2. Comparative molecular evolution of key photosynthetic redox regulators and photosynthesis protein families in algae and land plants candidate proteins were extracted using BLASTP (NCBI), and the all-against-all BLASTP search with satisfied E-value of 10−10 and query coverage of 50%. Heat map was generated using TBtools [85] from the data of seven photosynthesis-related proteins and four redox-related families. Black squares indicate the proteins that do not satisfy the selection criteria. ACHTs: atypical Cys His-rich, CF1-y: gamma subunit of ATP synthase; CP12; FBPase: fructose-1,6-bisphosphatase; GADPH: glyceraldehyde-3-phosphate dehydrogenase; NADP-MDH: NADP+ dependent malate dehydrogenase; RCA: ribulose-1,5-bisphosphate carboxylase/oxygenase activase; SBPase: sedoheptulose-1,7-bisphosphatase; TrxLs: Trx-like.
Antioxidants 11 02085 g002
Figure 3. Phylogenetic, conserved domains and 3D protein analysis of GAPDH and TxL2.2 in Plants and algae (A) Phylogenetic tree of GAPDH and TrxL2.2 from representative plant and algae lineages. The alignment was performed using MAFFT v7.409 [86]. A maximum likelihood tree was generated using RAxML-HPC2 on XSEDE. (B) Sequence alignment of GAPDH and TrxL2.2 for conserved NAD binding and Trx-like domains, respectively. The alignment was performed using Jalview [87] with default parameters. (C) 3D protein structures were predicted using a Swiss-prot server in representative algae and plant species. At, Arabidopsis thaliana, Os, Oryza sative; Zm, Zea mays; Amt, Amborella trichopoda; Pa, Picea abies; Af, Azolla filiculoids; Sm, Selaginella moellendorffii; Pp, Physcomitrella patens; Kf, Klebsormidium flaccidum; Mp, Marchantia polymorpha.
Figure 3. Phylogenetic, conserved domains and 3D protein analysis of GAPDH and TxL2.2 in Plants and algae (A) Phylogenetic tree of GAPDH and TrxL2.2 from representative plant and algae lineages. The alignment was performed using MAFFT v7.409 [86]. A maximum likelihood tree was generated using RAxML-HPC2 on XSEDE. (B) Sequence alignment of GAPDH and TrxL2.2 for conserved NAD binding and Trx-like domains, respectively. The alignment was performed using Jalview [87] with default parameters. (C) 3D protein structures were predicted using a Swiss-prot server in representative algae and plant species. At, Arabidopsis thaliana, Os, Oryza sative; Zm, Zea mays; Amt, Amborella trichopoda; Pa, Picea abies; Af, Azolla filiculoids; Sm, Selaginella moellendorffii; Pp, Physcomitrella patens; Kf, Klebsormidium flaccidum; Mp, Marchantia polymorpha.
Antioxidants 11 02085 g003
Figure 4. Redox signaling in evolutionary distinct photosynthetic machinery. Light and dark periods activate and deactivate the target photosynthesis enzymes. (A) Calvin-Benson Cycle (B) Kranz anatomy (C) CAM photosynthesis (D) Early divergent land plants and green algae. Trx-like proteins carry oxidation and reduction reactions to maintain PSI homeostasis. RCA: ribulose-1,5-bisphosphate carboxylase/oxygenase activase; FBPase: fructose-1,6-bisphosphatase; SBPase: sedoheptulose-1,7-bisphosphatase; PRK: phosphoribulokinase; PSI: photosystem I; PQ: plastoquinone, PTOX: plastid terminal oxidase; PGR5-PGRL1: 5-PGR5-like photosynthetic phenotype 1; FNR: ferredoxin; NADPH reductase; Fd: ferredoxin.
Figure 4. Redox signaling in evolutionary distinct photosynthetic machinery. Light and dark periods activate and deactivate the target photosynthesis enzymes. (A) Calvin-Benson Cycle (B) Kranz anatomy (C) CAM photosynthesis (D) Early divergent land plants and green algae. Trx-like proteins carry oxidation and reduction reactions to maintain PSI homeostasis. RCA: ribulose-1,5-bisphosphate carboxylase/oxygenase activase; FBPase: fructose-1,6-bisphosphatase; SBPase: sedoheptulose-1,7-bisphosphatase; PRK: phosphoribulokinase; PSI: photosystem I; PQ: plastoquinone, PTOX: plastid terminal oxidase; PGR5-PGRL1: 5-PGR5-like photosynthetic phenotype 1; FNR: ferredoxin; NADPH reductase; Fd: ferredoxin.
Antioxidants 11 02085 g004
Figure 5. Manipulation of ROS toxicity in response to environmental cues. Excessive ROS production under extreme environment damage chloroplast-related proteins. Using bioengineering approaches, supplementing the nuclear D1 protein pathway into chloroplast and installation of xanthophyll cycle mediating reversible inter-conversion of ZEP and VDE could enhance photosynthetic efficiency. In addition, functional characterization of random mutations could help plants to adapt to undesirable environmental cues efficiently.
Figure 5. Manipulation of ROS toxicity in response to environmental cues. Excessive ROS production under extreme environment damage chloroplast-related proteins. Using bioengineering approaches, supplementing the nuclear D1 protein pathway into chloroplast and installation of xanthophyll cycle mediating reversible inter-conversion of ZEP and VDE could enhance photosynthetic efficiency. In addition, functional characterization of random mutations could help plants to adapt to undesirable environmental cues efficiently.
Antioxidants 11 02085 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Riaz, A.; Deng, F.; Chen, G.; Jiang, W.; Zheng, Q.; Riaz, B.; Mak, M.; Zeng, F.; Chen, Z.-H. Molecular Regulation and Evolution of Redox Homeostasis in Photosynthetic Machinery. Antioxidants 2022, 11, 2085. https://doi.org/10.3390/antiox11112085

AMA Style

Riaz A, Deng F, Chen G, Jiang W, Zheng Q, Riaz B, Mak M, Zeng F, Chen Z-H. Molecular Regulation and Evolution of Redox Homeostasis in Photosynthetic Machinery. Antioxidants. 2022; 11(11):2085. https://doi.org/10.3390/antiox11112085

Chicago/Turabian Style

Riaz, Adeel, Fenglin Deng, Guang Chen, Wei Jiang, Qingfeng Zheng, Bisma Riaz, Michelle Mak, Fanrong Zeng, and Zhong-Hua Chen. 2022. "Molecular Regulation and Evolution of Redox Homeostasis in Photosynthetic Machinery" Antioxidants 11, no. 11: 2085. https://doi.org/10.3390/antiox11112085

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