Next Article in Journal
Taste and Smell Disorders in Cancer Treatment: Results from an Integrative Rapid Systematic Review
Previous Article in Journal
Benefits of the Non-Steroidal Mineralocorticoid Receptor Antagonist Finerenone in Metabolic Syndrome-Related Heart Failure with Preserved Ejection Fraction
Previous Article in Special Issue
Dynamic Changes in Ascorbic Acid Content during Fruit Development and Ripening of Actinidia latifolia (an Ascorbate-Rich Fruit Crop) and the Associated Molecular Mechanisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Functions of Chloroplastic Ascorbate in Vascular Plants and Algae

Laboratory for Molecular Photobioenergetics, Institute of Plant Biology, Biological Research Centre, Temesvári krt 62, H-6726 Szeged, Hungary
Int. J. Mol. Sci. 2023, 24(3), 2537; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032537
Submission received: 21 December 2022 / Revised: 17 January 2023 / Accepted: 24 January 2023 / Published: 28 January 2023
(This article belongs to the Special Issue Sensing, Regulation, and Signalling of Ascorbate in Plants)

Abstract

:
Ascorbate (Asc) is a multifunctional metabolite essential for various cellular processes in plants and animals. The best-known property of Asc is to scavenge reactive oxygen species (ROS), in a highly regulated manner. Besides being an effective antioxidant, Asc also acts as a chaperone for 2-oxoglutarate-dependent dioxygenases that are involved in the hormone metabolism of plants and the synthesis of various secondary metabolites. Asc also essential for the epigenetic regulation of gene expression, signaling and iron transport. Thus, Asc affects plant growth, development, and stress resistance via various mechanisms. In this review, the intricate relationship between Asc and photosynthesis in plants and algae is summarized in the following major points: (i) regulation of Asc biosynthesis by light, (ii) interaction between photosynthetic and mitochondrial electron transport in relation to Asc biosynthesis, (iii) Asc acting as an alternative electron donor of photosystem II, (iv) Asc inactivating the oxygen-evolving complex, (v) the role of Asc in non-photochemical quenching, and (vi) the role of Asc in ROS management in the chloroplast. The review also discusses differences in the regulation of Asc biosynthesis and the effects of Asc on photosynthesis in algae and vascular plants.

1. Introduction

Ascorbate (Asc) is the most abundant water-soluble metabolite in plants and is essential for plant growth and development. Asc is also required in the human diet; therefore, serious efforts are underway to increase the Asc contents of fruits and vegetables (reviewed, e.g., by [1,2]). In vascular plants and green algae, Asc is synthesized via the Smirnoff-Wheeler pathway, in which GDP-D-mannose is converted to L-galactono-1,4-lactone in the cytoplasm by GDP-D-mannose 3′,5′ epimerase (GME), GDP-L-galactose phosphorylase (GGP), L-galactose-1-P phosphatase (GPP), and L-galactose dehydrogenase (GDH) [3,4]. The final step, the conversion of L-galactono-1,4-lactone to Asc, occurs in the mitochondria and is catalyzed by L-galactono-1,4-lactone dehydrogenase (GLDH) at Complex I [4,5] (Figure 1). VTC2, encoding GGP, plays a vital role in the regulation of Asc biosynthesis, both in vascular plants and green algae [6,7]. In vascular plants, not only its expression but also a feedback mechanism on GGP translation by Asc and a small ORF largely determines the rate of Asc biosynthesis, thereby its cellular concentration [8,9,10,11].
Alternative Asc biosynthesis routes, namely, the galacturonate, the L-gulose pathway, and the myoinositol pathway, have been suggested to contribute to Asc biosynthesis (reviewed by [12]). However, various lines of evidence show that they do not contribute significantly to Asc biosynthesis in plant leaves [13].
The best-known role of Asc is to scavenge reactive oxygen species (ROS), in a highly regulated manner in each cellular compartment (reviewed by [14]). It is linked to its capacity to act as a weak reductant and to the non-toxicity of the generated monodehydroascorbate (MDA) and the bicyclic dehydroascorbate (DHA) [15,16]. Asc effectively eliminates superoxide and tocopherol radicals, though it reacts slowly with H2O2. Plants use Asc peroxidases (APX) to eliminate H2O2 more effectively (reviewed by [14,15]). The reaction oxidizes Asc into MDA radicals, which can spontaneously disproportionate into DHA and Asc. The MDA radicals are then recycled back to Asc through MDA reductases (MDARs) or reduced ferredoxin produced at the acceptor side of photosystem I (PSI) [14,17], while DHA is reduced by DHA reductase (DHAR) in glutathione-dependent non-enzymatic or enzymatic reactions [18,19] (Figure 1).
DHA is unstable; therefore, if it is not recycled back to Asc, it is rapidly and irreversibly degraded into oxalate, L-threonate, or tartaric acid, depending on the plant species [20,21]. Nevertheless, the regeneration of DHA and MDA occurs very effectively, as most of the Asc pool is in the reduced state under both non-stress and mild stress conditions. An exception in this respect is the apoplast in which the oxidation of Asc occurs at a relatively high rate due to the activity of Asc oxidase; this peculiarity is most probably related to cell wall loosening and growth [22,23].
Besides being an effective antioxidant, Asc also acts as a chaperone for 2-oxoglutarate-dependent dioxygenases (2-ODDs). Plant 2-ODDs are involved, for instance, in hormone metabolism (ethylene, abscisic acid, gibberellins, indole-3-acetic acid) and the synthesis of various secondary metabolites, such as anthocyanins and glucosinolates [24,25,26]. Asc is also involved in the epigenetic regulation of gene expression [27] via modifying the activities of ten-eleven translocation (TET) and Jumonji C-domain-containing histone demethylases that are both 2-ODDs [27,28,29,30]. Asc has also been proposed to be involved in iron transport and Ca2+ signaling [31,32].
Thus Asc affects plant growth, development, and stress resistance via various mechanisms that are challenging to dissect. Growth defects have been reported for the Asc-deficient vtc2-1 Arabidopsis mutant containing about 20% Asc relative to the wild type, but later it turned out to be due at least partially to cryptic mutations [33]. The more recently identified T-DNA mutant vtc2-4 did not show any alteration in the phenotype under optimal growth conditions in [33], but a mild phenotype was observed in [34]. Moreover, the approx. 80% reduction in Asc level observed in the vtc2 mutants only moderately enhances high light sensitivity [35]; thus, this low level of Asc seems to be sufficient for the housekeeping functions under laboratory conditions. On the other hand, the situation can be different in the field, and it may also be species-dependent because in rice, for instance, Asc content reduction did lead to diminished growth and photosynthesis rate [36].
Several excellent reviews have already been published on the biosynthesis and the various functions of Asc (e.g., [11,14,15,37,38,39,40]). This review focuses on less extensively discussed issues, such as the relationship between photosynthesis and Asc biosynthesis and the effects of Asc on photosynthetic electron transport in vascular plants and algae.

2. Regulation of Ascorbate Biosynthesis by Light

Under regular growth and moderate light conditions (at approx. 100 μmol m−2s−1), the Asc level of Arabidopsis leaves is in the range of 2 to 5 μmol/g fresh weight (FW). The Asc content varies by a factor of two depending on the time of day, with a minimum at night and a maximum in the afternoon [41,42,43]. Under stress conditions, including high light [9,42,44,45], ozone [46,47], salt [48,49], drought stress [50], and nitrogen starvation [51] a two- to three-fold increase in Asc level occurs on a timescale of days, with an apparent upper limit of about 8–10 µmol Asc/g FW. Using stable and inducible expression systems in Arabidopsis, the maximum Asc concentration that can be reached is in the same range of maximum 10 µmol Asc/g FW (e.g., [7]; reviewed by [11,38]). Asc content can be somewhat more effectively increased by L-galactono-1,4-lactone treatment to the range of 15 µmol Asc/g FW [52,53].
These results demonstrate that Asc biosynthesis is highly regulated, to keep Asc in the optimum concentration range during the normal life cycle of plants and under stress conditions. Therefore, understanding the signaling pathways and the control mechanism of Asc biosynthesis are of primary importance to achieve a substantial increase in the Asc content of crops.
One major factor controlling Asc biosynthesis is light. Several Smirnoff-Wheeler pathway genes are induced by illumination. These include GDP-D-mannose pyrophosphorylase (GMP), GGP, GPP, and GLDH [54,55,56]. Transcript levels of these biosynthesis genes also follow a circadian rhythm [42]. It has been shown that 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), an inhibitor of photosystem II (PSII), prevents the increase in Asc pool size in Arabidopsis upon a shift from light-dark cycle to continuous light conditions, along with a decrease in the transcript levels of GMP, GGP, GPP, and GLDH [55].
When Arabidopsis plants were transferred to a medium containing sucrose, the leaf Asc levels decreased along with a decrease in the rate of CO2 fixation. Asc content diminishment upon sucrose feeding did not occur in a sugar-insensitive Arabidopsis abi4/sun6 mutant [55]. Exogenous application of glucose in pea seedlings did not affect Asc content [57], whereas sucrose feeding in detached broccoli inflorescences delayed Asc depletion [58], and in tomato fruit, it increased the Asc content [59].
The Asc content decreases markedly in darkness (at a rate of approx. 2% per hour), while the levels of its degradation products increase [21], and dark-grown plants produce no Asc at all [60]. Interestingly, the downregulation of VTC2 expression by DCMU was slightly reversed in Arabidopsis mutants lacking GENOMES UNCOUPLED 1 (GUN1); [61], the master regulator of chloroplast-to-nucleus retrograde signaling.
Thus, the results suggest that photosynthesis-derived signal(s) participate in the light activation of Asc biosynthesis, and the role of photosynthesis is not solely to provide a carbon source for Asc biosynthesis. On the other hand, soluble carbohydrates may affect the expression of Asc biosynthesis genes and the regeneration of Asc. The molecular mechanisms underlying the regulation of Asc biosynthesis by photosynthesis remain to be explored. It is conceivable that the redox state of the plastoquinone pool plays a regulatory role in Asc biosynthesis. In principle, H2O2 and other ROS produced during photosynthesis could also influence the expression of Asc biosynthesis genes; however, experimental evidence is still lacking on such roles.
Light may also directly affect Asc biosynthesis, degradation, and regeneration. Indeed, light can regulate GMP activity, namely, via CSN5B, a photomorphogenic factor that is part of a CSN complex, negatively regulating photomorphogenesis in Arabidopsis via proteasomal degradation [62,63,64]. Wang et al. [64] demonstrated that GMP is polyubiquitinated and degraded in darkness via interaction with CSN5B. They also found that loss of CSN5B function impaired the effect of light on Asc synthesis in response to continuous light or darkness, showing that CSN5B is a posttranslational regulator in Asc biosynthesis.
A F-box type repressor, known as Asc acid mannose pathway regulator 1 (AMR1), also regulates Asc biosynthesis in a light-dependent manner [65]. DNA knockout lines for AMR1 accumulated two-fold greater foliar Asc than the wild type. AMR1 also negatively affected the expression levels of most Asc biosynthesis genes, including GMP, GME, GGP, GPP, L-GalDH, and GLDH. In addition, AMR1 expression was higher in aging leaves, and lower at medium light than at low light intensity.
A HD-Zip I family transcription factor in tomato, SlHZ24, positively regulates the accumulation of Asc by binding to the promoter of SlGMP3, and it also regulates the expression of GME and GGP. Accordingly, the Asc content fluctuated following the expression of SlHZ24 in a light-dependent manner [66].
GGP also shows a significant response to different light levels due to the regulation by light-responsive cis-elements in its promoter, namely, a G-box motif [67]. Furthermore, light-responsive cis-elements have also been identified in the promoters of GPP and GLDH in rice [68].
It was also proposed that the VTC3 dual protein kinase/protein phosphatase is involved in signal transduction to adjust Asc levels in response to light and temperature changes [69]. VTC3 is probably located in the chloroplast, and it was suggested that the control exerted by VTC3 is post-transcriptional and does not alter the transcript levels of Asc biosynthesis genes in Arabidopsis [69].
The conversion of L-galactono-1,4-lactone to Asc in the mitochondria is light and photosynthesis dependent because photosynthetic inhibitors prevent it [52]. H2O2 produced upon stress effects may inactivate selectively and reversibly Arabidopsis GLDH by oxidizing a cysteine residue (Cys-340) [70]. GLDH is protected from inactivation both by L-galactono-1,4-lactone and Asc; therefore, their availabilities and the level of H2O2 may affect the rate of Asc biosynthesis in vascular plants.
Bryophytes, including Brachytecium velutinum, Marchantia polymorpha, and Physcomitrium (formerly Physcomitrella) patens, possess Asc content slightly less than seed plants, in the range of 0.3 to 2 μmol/g [71,72]. Sodeyama et al. [72] presented evidence on the Smirnoff-Wheeler pathway as a source of Asc, whereas the D-galacturonate pathway did not seem to contribute to Asc biosynthesis in P. patens. Interestingly, two VTC2 paralogs are functional in P. patens, and, in contrast to higher plants, they are both equally responsible for Asc biosynthesis. Furthermore, the light-induced fluctuation of the transcript level of the two VTC2 genes was comparable to AtVTC2. Additionally, DCMU treatment diminished VTC2 expression and Asc content, as observed earlier in Arabidopsis. Thus, VTC2 expression is a crucial control point of Asc biosynthesis in P. patens. Interestingly, knockout mutants with low Asc content exhibited restricted side branch growth in their protonemata; this may be caused by multiple factors related to the various physiological functions of Asc.
Green alga grown under benign conditions contain about 100-fold less Asc than vascular plants, i.e., in the range of 100 to 500 μM [73,74,75]. Asc is synthesized via the Smirnoff-Wheeler pathway in green algae; however, its regulation significantly differs compared to plants. It has been shown that in Chlamydomonas reinhardtii, the expression of the VTC2 gene is induced by H2O2 and 1O2, resulting in a strong increase in Asc content. On the other hand, photosynthesis is not directly required for Asc biosynthesis. Additionally, in contrast to plants, there is no circadian regulation of Asc biosynthesis, and C. reinhardtii lacks negative feedback regulation by Asc in the physiological concentration range. These mechanisms enable a rapid and manifold increase in Asc content upon various stress treatments, including light stress and sulfur deprivation [74,76,77].

3. Interaction between Photosynthetic and Mitochondrial Electron Transport in Relation to Asc Biosynthesis

Photosynthetic electron transport rates (ETR) and NADPH levels are only slightly affected in vtc2 mutants containing about 20% Asc in comparison with wild-type Arabidopsis plants when grown under moderate or low light conditions [35,44,78,79], whereas the photosynthetic ETR of vtc2 mutants is diminished at high light [44]. The vtc2 mutants have a slightly lower stomatal conductance, which may be related to a regulatory effect of Asc [79,80]; this, however, is compensated by a larger stomatal number and increased RuBisCO content. Therefore, the overall CO2 assimilation rate is affected by Asc deficiency only in high light, but not under normal growth conditions in Arabidopsis [79,81].
The last step of Asc biosynthesis, the conversion of L-galactono-lactone into Asc, is catalyzed by GLDH in the mitochondria [82,83]. GLDH is a protein of 58 kDa, located at Complex I in the mitochondrial intermembrane space, tightly tethered to the membrane through protein-protein interactions [84] (Figure 1). Complex I is organized in two arms: the matrix arm transfers electrons from NADH to ubiquinone, and the membrane arm is responsible for proton translocation [85]. In addition to being essential for Asc biosynthesis, GLDH has a non-enzymatic role in the assembly of the membrane arm of Complex I [86]. This function is likely to be independent of the role of GLDH in Asc biosynthesis because the vtc2-1 Arabidopsis mutant accumulates wild-type levels of Complex I [86].
During the oxidation of L-galactono-1,4-lactone to Asc by GLDH, electrons are fed into the mitochondrial electron transport chain via cytochrome c [5,87], a soluble redox-active heme protein that transfers electrons from Complex III to Complex IV [88] (Figure 1). Cytochrome c knockdown mutants of Arabidopsis had a 60% decrease in GLDH activity without affecting the Asc content, showing that low cytochrome c levels are enough under normal growth conditions to sustain Asc biosynthesis [89].
Plant mitochondria also have an alternative oxidase (AOX) pathway, taking electrons directly from the ubiquinone pool without the contribution of the cytochrome c pathway (Figure 1). Bartoli et al. [90] found that AOX-overexpressing Arabidopsis lines accumulated more Asc than wild-type plants, particularly at high light. Higher throughput in the cytochrome c pathway would require a larger pool of electron acceptors for L-galactono-1,4-lactone oxidation; therefore, an enhanced capacity of the AOX pathway may favor Asc biosynthesis by maintaining the cytochrome c pool in a more oxidized state. This is particularly relevant under high light conditions to prevent over-reduction of the mitochondrial electron transport chain and, at the same time, to enhance Asc biosynthesis to protect the cells against the damaging effect of ROS [90]. These results show that integration of L-galactono-1,4-lactone oxidation and mitochondrial electron transport chain activity via cytochrome c could coordinate Asc biosynthesis and respiration [88].
It has also been shown that the conversion of L-galactono-1,4-lactone depends on the photosynthetic electron transport chain because DCMU and dibromothymoquinone could effectively inhibit the Asc content increase upon L-galactono-1,4-lactone treatment [52]. This result indicates the occurrence of a crosstalk between photosynthetic and respiratory electron transport chains (Figure 1).
It has been suggested that Asc may be a signal connecting the metabolisms of chloroplast and mitochondria [37,91] based on observations on transgenic tomato plants antisensed in mitochondrial malate dehydrogenase (mdh). When grown under long-day conditions, these mdh lines had reduced tricarboxylic acid (TCA) cycle activity without affecting respiration, and intriguingly, CO2 assimilation rates and carbohydrates were slightly enhanced compared to wild-type plants, and an approx. fourfold increase in Asc content occurred [92]. This was explained by an upregulated flux through GLDH in the mdh lines and a higher capacity to use L-galactono-lactone as a respiratory substrate, thereby suggesting that GLDH can effectively act as an alternative electron donor in cases where flux through the TCA cycle is impaired [91,92]. However, contradicting results were obtained under short-day conditions [91] and in Arabidopsis mdh mutants [93]. On the other hand, Asc feeding to isolated leaf discs also resulted in increased photosynthesis rates, further suggesting an Asc-mediated link between the energy-generating processes of respiration and photosynthesis.
In summary, these results suggest that the interaction between chloroplasts and mitochondria acts as a vital determinant of the light-dependent regulation of Asc biosynthesis in plants and that Asc may act as a metabolic regulator between the energy systems of the mitochondria and chloroplasts [37]. However, the mechanistic details and the sensing system remain to be explored.

4. Ascorbate Is an Alternative, ‘Emergency’ Donor to Photosystem II

Asc is a weak reductant that has the potential to reduce amino acid radicals, such as tyrosine and tryptophan [94]. This feature makes it capable of donating electrons to TyrZ+ in PSII with inactive oxygen-evolving complexes (OEC). This was initially demonstrated in vitro on TRIS-washed, UV-B-irradiated, and heat-treated isolated thylakoids [95,96,97] and later on heat-treated intact leaves [98,99].
Heat stress results in the removal of the extrinsic proteins and the release of Ca- and Mn-ions from their binding sites, resulting in the inactivation of OEC [100,101]. It has been shown that electron donation from Asc to TyrZ+ occurs in heat-stressed leaves; thus, Asc is a naturally occurring electron donor that can replace water, the terminal electron donor of PSII [99] (Figure 1). With the aid of chlorophyll a fluorescence induced by short (5-ms) light pulses it was shown that the halftime of electron donation from Asc to PSII is in the range of 25 ms in wild-type Arabidopsis leaves and about 55 ms in Asc-deficient vtc2 mutants [78,99]. This alternative electron transport occurs in Arabidopsis, pea, barley, Marchantia polymorpha, Nephrolepis exaltata, C. reinhardtii, etc.; thus, it appears to be ubiquitous in the plant kingdom [99]. The electron transfer rate from Asc to PSII depends on the species and their physiological state, which is most probably related to the availability of Asc in the lumen (probably in the range of a few mM, [102]).
Isolated photosynthetic samples with inactivated OECs are extremely susceptible to illumination. The impaired electron donation from the OEC results in the accumulation of highly oxidizing radicals, including P680+, TyrZ+, and superoxide and hydroxyl radicals [103,104], leading to a rapid inactivation and degradation of PSII reaction centers [105,106]. This type of photodamage is called weak light or donor-side-induced photoinhibition. By using intact leaves of wild-type, Asc-overproducing (miox4) [107], and Asc-deficient Arabidopsis mutants (vtc2-3) [6] subjected to heat stress (40 °C, 15 min), it was demonstrated that the continuous electron flow from Asc to PSII alleviates PSII photoinactivation [78]. Gradual inactivation of PSII charge separation activity occurred on a time scale of tens of minutes, along with extensive protein degradation, including probably the complete disassembly of PSII [78]. Besides the rate of photoinactivation, the recovery rate from the photoinactivated state also depended on leaf Asc content. Thereby, Asc contributes significantly to the ability of plants to withstand heat stress conditions and aids recovery [78,108].
Asc may also act as an alternative electron donor in bundle sheath chloroplasts. These are found in the so-called NADP+ malic enzyme type species carrying out C4-photosynthesis, such as maize and sorghum. The amount of PSII in bundle sheath chloroplasts is small, and their OECs have low activity. It was shown that Asc is an effective electron donor for PSII in bundle sheath chloroplasts in vivo [109,110]. On the other hand, since the number of PSII reaction centers is low, photosynthetic electron transport is moderate, but it is sufficient to maintain PSI cyclic electron flow, ensuring thylakoid membrane energization and ATP synthesis for the Calvin-Benson cycle [109,110]. Moreover, the replacement of water by Asc as a PSII electron donor also ensures low O2 concentration within the bundle sheath chloroplasts, diminishing the risk of competition of O2 with CO2 molecules for the catalytic sites of RuBisCO (reviewed by [111]).
Asc also donates electrons to PSI in isolated thylakoid membranes (see, e.g., [112]) and in DCMU-treated bundle sheath cells isolated from maize leaves [113]. However, in vivo, Asc was a far more effective electron donor for PSII than for PSI [99,109,110].

5. Ascorbate May Impair the Oxygen-Evolving Complex

When considering the physiological roles of Asc, it has to be taken into account that it is a reductant, and therefore, its cellular concentration is to be maintained in a particular range [114]. The basal Asc concentration in green algae is very low compared to higher plants (approx. 60 to 100 µM in C. reinhardtii and 5 mM in Arabidopsis) [74,75,77,115]. It was observed that upon sulfur deprivation of C. reinhardtii, Asc accumulates to the mM range and that, in this range, Asc over-reduces the Mn cluster of OEC [76,77]. The exact mode of action, i.e., as to which S-state is being inactivated, is not understood.
Once the Mn-cluster is over-reduced by Asc, it may continuously provide electrons to TyrZ+. However, the electron donation by Asc to PSII is relatively slow (halftime of approx. 20 to 50 ms, [99]) in comparison with the rate of electron transfer from intact OECs to TyrZ+ (halftime of about 0.1 to 1 ms; [116]); for this reason, Asc cannot entirely prevent the accumulation of TyrZ+ and P680+ upon illumination in sulfur-deprived C. reinhardtii cultures. Furthermore, strongly oxidizing species lead to donor-side induced photoinhibition, resulting in the relatively rapid degradation of PSII reaction center proteins, including PsbA, CP43, PSBO, and possibly others [77].
The loss of PSII activity could be regarded as damage induced by sulfur deprivation. However, upon downregulation of PSII activity, overexcitation and further photodamage are minimized. The metabolic changes downregulating photosynthetic activity and cell proliferation may serve to preserve cellular sulfur content and avoid more substantial damage [77]. On the other hand, the inactivation of OECs contributes to the establishment of hypoxia enabling hydrogenase expression, and H2 production will act as a safety valve for photosynthetic electron transport [117]. Thereby, the damage imposed by sulfur limitation is minimized, and the alga cells may recover if sulfur becomes available again [118].
Intriguingly, Asc inactivates the OEC in green algae when accumulated to the mM range, but the same Asc concentration in vascular plants is physiological [115], which enables a full operation of OEC activity. On the other hand, it was demonstrated earlier that upon the chemical removal of the extrinsic OEC subunits in isolated PSII membranes, bulky reductants, including Asc, could directly reduce the Mn-cluster [119].
There are notable differences between the extrinsic OEC proteins of vascular plants and green algae [120,121]. Namely, in green algae, the Mn-cluster is shielded by a 33 kDa PSBO subunit and two smaller subunits, PSBP and PSBQ, with some structural differences and binding properties in comparison to vascular plants [121,122]. Moreover, in vascular plants, an additional subunit, PSBR, is found in the vicinity of the Mn-cluster [123]. The structural differences and the variance in binding properties of the extrinsic proteins may be key in explaining why Asc reduces the Mn-cluster in algae but not in higher plants when present in the same concentration range. It is also conceivable that, during evolution, as cellular Asc concentration increased [124], the extrinsic proteins have evolved to protect the OEC against the reducing effect of Asc present in the thylakoid lumen.
In Arabidopsis subjected to darkness for 24 h, the Mn-cluster becomes inactivated, being one of the earliest effects of dark-induced senescence on photosynthesis [125]. Remarkably, the extent of OEC inactivation was much weaker in Asc-deficient mutants compared to wild-type plants, suggesting that Asc was responsible for the diminishment of oxygen evolution. In a psbo1 knockout mutant, the compromised OEC activity was further aggravated upon dark treatment, suggesting that the extrinsic proteins protect the OEC against the reducing effect of Asc. In the absence of PSBR, only a slightly disturbed photosynthetic activity was observed under normal growth conditions, whereas a strongly diminished OEC activity was observed in the dark. A double psbo1 vtc2 mutant showed a slightly milder photosynthetic phenotype than that of the single psbo1 mutant [125]. Thus, these results suggest that Asc over-reduces the Mn-complex in prolonged darkness. This is probably enabled by a dark-induced dissociation of the extrinsic OEC subunits, which would otherwise hinder the access of Asc to the Mn-cluster [125] (Figure 1) or, hypothetically, it may be related to the lumen volume decrease in the dark [126].
Another example that Asc may negatively affect certain cellular processes was found by Castro et al. [127]. Exogenous Asc concentrations above 30 mM caused cellular and oxidative damage that were enhanced by high light. The high Asc concentration induced H2O2 accumulation, stomatal closure, and impairment in CO2 assimilation, as well as photosynthetic electron transport. Therefore, these data show that Asc concentration and localization need to be highly controlled, particularly relevant when aiming at generating crops with elevated Asc contents.

6. The Role of Ascorbate in Non-Photochemical Quenching

Excess light may lead to photooxidative stress involving the formation of ROS and light damage (recently reviewed by, e.g., [128,129,130]). Non-photochemical quenching (NPQ) of excitation energy is a complex photoprotection mechanism, including energy-dependent (qE), zeaxanthin-related (qZ), state-transition-related (qT), and photoinhibitory (qI) components (e.g., [131,132]).
The pH-regulated qE component is formed in response to light-intensity changes within a few minutes. In vascular plants, activation of qE is mediated by PsbS that acts as a lumen pH sensor and induces LHCII antenna rearrangement in a zeaxanthin-dependent manner [133,134].
The qZ component of NPQ is activated in the time range of 10 to 30 min, which correlates with the formation of zeaxanthin by violaxanthin de-epoxidase (VDE) in the lipid phase of the thylakoid membrane of vascular plants [135,136,137]. VDE belongs to the lipocalin protein family, and it utilizes Asc as a co-substrate (Figure 1), providing the reducing power for de-epoxidation [138]. At the pH-optimum of VDE of pH 5.0, the Km value for Asc is approximately 1.0 mM [138,139], which is probably in the range of luminal Asc concentration [102]. During the operation of the violaxanthin cycle, VDE attaches to the luminal side of the thylakoid membrane following its pH-dependent activation [140,141,142]. The active VDE is probably a dimer, capable of binding violaxanthin and Asc [141,143].
The maximum VDE activity is several times faster than zeaxanthin epoxidase activity [144]. Consequently, zeaxanthin accumulates in strong light when a low lumen pH is established. In contrast, zeaxanthin is epoxidized only when VDE activity is low, i.e., under low light or in darkness. It was recently demonstrated that up-regulation of VDE, PsbS, and zeaxanthin epoxidase in soybean significantly accelerated the violaxanthin cycle, leading to faster induction and relaxation of NPQ. This increased the efficiency of CO2 assimilation and PSII electron transport in fluctuating light conditions and significantly improved the biomass yield in field studies, demonstrating that the violaxanthin cycle plays a crucial role in the regulation of photosynthesis, thereby relating to plant productivity [145].
Besides playing a role in NPQ, zeaxanthin may also contribute to photoprotection by acting as an antioxidant and possibly by modulating thylakoid membrane properties [146,147]. In addition, zeaxanthin may be required for the PSII repair cycle [148].
The physiological importance of Asc for zeaxanthin formation was demonstrated in vivo using vtc mutants of Arabidopsis: it was shown that they accumulate less zeaxanthin in high light, and consequently, they have diminished and/or delayed NPQ induction [6,45,78]. In the miox4 Asc-overproducing mutant [107], a slightly higher NPQ level was obtained than in the wild type [78], demonstrating that Asc may be limiting NPQ formation [44].
In contrast to the vtc2 mutant of Arabidopsis, the C. reinhardtii vtc2 mutant (Crvtc2-1) with strongly decreased Asc content performs normal violaxanthin de-epoxidation [75]. C. reinhardtii lacks a plant-type VDE and instead uses an unrelated enzyme belonging to lycopene cyclases, called Chlorophycean VDE (CVDE). It is located on the stromal side of the thylakoid membrane [149], and it does not require Asc for violaxanthin de-epoxidation [75]. On the other hand, a slow, H2O2-dependent NPQ component (probably qI) is enhanced upon Asc-deficiency in C. reinhardtii [75].
Many green alga species, including Chlorella vulgaris, contain plant-type VDEs [150], which are crucial for photoprotective NPQ. Thus, there is an evolutionary divergence of photoprotective mechanisms among Chlorophyta [151]. Chromalveolate algae, such as Phaeodactylum tricornutum, use the diadinoxanthin cycle instead of the violaxanthin cycle in a fashion similar to vascular plants to support qE [152]. PtVDE converts the monoepoxide diadinoxanthin to diatoxanthin, possibly involving violaxanthin as an intermediate [153]. The reaction requires Asc, though less than vascular plants [154]. On the other hand, violaxanthin is also required for the biosynthesis of fucoxanthin and its derivatives, which are the main light-harvesting pigments in Chromalveolate algae. The conversion of violaxanthin to neoxanthin is catalyzed by the so-called violaxanthin de-epoxidase-like (VDL) protein, which does not require Asc as a reductant. It was found that PtVDL is modulated only by pH, whereas VDE activity is controlled on multiple levels, including the pH-dependent affinity for Asc [155].
Mosses have plant-type VDE enzymes [156], which probably require Asc as a reductant. The regulation of Asc biosynthesis in vascular plants and mosses is similar [72], but the Asc-dependence of NPQ in mosses has not been investigated.

7. Reactive Oxygen Species Management by Ascorbate in the Chloroplast

The best-known and most extensively discussed role of Asc is participation in ROS management. Excellent recent reviews are available on this topic [14,129,130]; therefore, only a few aspects related to photosynthesis are mentioned here.
Asc is essential for the enzymatic scavenging of ROS in the so-called Mehler reaction or the water–water cycle (reviewed by e.g., [14,157], Figure 1). This reaction becomes particularly relevant at high light intensities and/or when the Calvin-Benson cycle cannot work at high speed, for instance under drought, cold, or salt stress. Under such conditions, the outflow of electrons at the PSI acceptor side is inhibited, leading to a surplus of electrons; therefore, ferredoxin reduces O2, and superoxide is produced. It is then reduced to H2O2 by superoxide dismutase, which is reduced by APX to water. MDA can be directly reduced back to Asc by PSI, and/or in the Asc-glutathione cycle in which MDAR and DHAR use NADPH as reducing power [158]. Asc thereby participates in the mitigation of ROS. In this respect, it is to be considered that ROSs are key signaling molecules that enable cells to respond to changes in environmental conditions rapidly, and they integrate different environmental signals to activate stress-response networks and defense mechanisms [130].
In addition to its function as a powerful antioxidant, Asc is also considered as being a major player in redox homeostasis and part of the redox signaling network that regulates plant responses to biotic stress [14,159,160]. Interestingly, however, an approx. 80% decrease in Asc content led to only a minor increase in glutathione level when plants were grown under standard laboratory conditions [35,81], though cellular glutathione levels are redistributed and an approx. two-fold increase in chloroplastic glutathione concentration was observed [161]. In addition, Asc peroxidase activity and the redox state of Asc are also unaltered in vtc2 mutants, just as well as the photosynthetic activity and carotenoid content, and only moderate changes occur at constant high light in these parameters [33,35]. Thus, an approx. 80% reduction of cellular Asc content does not lead to photooxidative stress, and apparently, Asc is in large excess in vascular plants, or its deficiency, are compensated via various yet unexplored mechanisms.

8. Open Questions and Perspectives

Future work will be required to gain a complete overview of Asc functions within the cell. In recent years, several functions of Asc have been discovered in addition to its best-known role of mitigating ROS accumulation, and it is likely that the list will be further expanded. For instance, Asc serves as a chaperone for 2-ODD, a versatile oxidative enzyme group in plants, of which only a couple have been characterized [162], and it is conceivable that Asc regulates the activities of some of them. In addition, the role of Asc as a metabolic regulator and a key player in chloroplast-mitochondria signaling has been suggested [37], which warrants experimental confirmation. Moreover, Asc transporters may also play a key role in Asc metabolism and functions of which only two have been characterized on a molecular level in vascular plants (AtPHT4;4 and AtDTX25 in the chloroplast envelope membrane and the vacuole, respectively [163,164]). Therefore, future work needs to be directed at exploring other essential Asc transporters both in vascular plants and algae.

Funding

This work was supported by the National Research, Development, and Innovation Office (K132600), and the Lendület/Momentum Programme of the Hungarian Academy of Sciences (LP2014/19).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data presented in this study are available within this article. There are no special databases associated with this manuscript.

Acknowledgments

I apologize to colleagues whose work I was unable to review. I thank Soujanya Kuntam (BRC Szeged) for the careful proofreading.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Macknight, R.C.; Laing, W.A.; Bulley, S.M.; Broad, R.C.; Johnson, A.A.; Hellens, R.P. Increasing ascorbate levels in crops to enhance human nutrition and plant abiotic stress tolerance. Curr. Opin. Biotechnol. 2017, 44, 153–160. [Google Scholar] [CrossRef]
  2. Fenech, M.; Amaya, I.; Valpuesta, V.; Botella, M.A. Vitamin C content in fruits: Biosynthesis and regulation. Front. Plant Sci. 2019, 9, 2006. [Google Scholar] [CrossRef] [Green Version]
  3. Wheeler, G.L.; Jones, M.A.; Smirnoff, N. The biosynthetic pathway of vitamin C in higher plants. Nature 1998, 393, 365–369. [Google Scholar] [CrossRef]
  4. Smirnoff, N. Ascorbate biosynthesis and function in photoprotection. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2000, 355, 1455–1464. [Google Scholar] [CrossRef]
  5. Millar, A.H.; Mittova, V.; Kiddle, G.; Heazlewood, J.L.; Bartoli, C.G.; Theodoulou, F.L.; Foyer, C.H. Control of ascorbate synthesis by respiration and its implications for stress responses. Plant Physiol. 2003, 133, 443–447. [Google Scholar] [CrossRef] [Green Version]
  6. Conklin, P.L.; Saracco, S.A.; Norris, S.R.; Last, R.L. Identification of ascorbic acid-deficient Arabidopsis thaliana mutants. Genetics 2000, 154, 847–856. [Google Scholar] [CrossRef]
  7. Fenech, M.; Amorim-Silva, V.; Esteban del Valle, A.; Arnaud, D.; Ruiz-Lopez, N.; Castillo, A.G.; Smirnoff, N.; Botella, M.A. The role of GDP-L-galactose phosphorylase in the control of ascorbate biosynthesis. Plant Physiol. 2021, 185, 1574–1594. [Google Scholar] [CrossRef]
  8. Laing, W.A.; Martínez-Sánchez, M.; Wright, M.A.; Bulley, S.M.; Brewster, D.; Dare, A.P.; Rassam, M.; Wang, D.; Storey, R.; Macknight, R.C.; et al. An upstream open reading frame is essential for feedback regulation of ascorbate biosynthesis in Arabidopsis. Plant Cell 2015, 27, 772–786. [Google Scholar] [CrossRef] [Green Version]
  9. Bulley, S.; Laing, W. The regulation of ascorbate biosynthesis. Curr. Opin. Plant Biol. 2016, 33, 15–22. [Google Scholar] [CrossRef]
  10. Deslous, P.; Bournonville, C.; Decros, G.; Okabe, Y.; Mauxion, J.P.; Jorly, J.; Gadin, S.; Brès, C.; Mori, K.; Ferrand, C.; et al. Overproduction of ascorbic acid impairs pollen fertility in tomato. J. Exp. Bot. 2021, 72, 3091–3107. [Google Scholar] [CrossRef]
  11. Terzaghi, M.; De Tullio, M.C. The perils of planning strategies to increase vitamin C content in plants: Beyond the hype. Front. Plant Sci. 2022, 13, 1096549. [Google Scholar] [CrossRef]
  12. Ntagkas, N.; Woltering, E.J.; Marcelis, L.F.M. Light regulates ascorbate in plants: An integrated view on physiology and biochemistry. Environ. Exp. Bot. 2018, 147, 271–280. [Google Scholar] [CrossRef]
  13. Kavkova, E.I.; Blöchl, C.; Tenhaken, R. The Myo-inositol pathway does not contribute to ascorbic acid synthesis. Plant Biol. 2019, 21, 95–102. [Google Scholar] [CrossRef] [Green Version]
  14. Foyer, C.H.; Kyndt, T.; Hancock, R.D. Vitamin C in plants: Novel concepts, new perspectives, and outstanding issues. Antiox. Redox Signal. 2020, 32, 463–485. [Google Scholar] [CrossRef]
  15. Smirnoff, N. Ascorbic acid metabolism and functions: A comparison of plants and mammals. Free Radic. Biol. Med. 2018, 122, 116–129. [Google Scholar] [CrossRef]
  16. Njus, D.; Kelleya, P.M.; Tub, Y.J.; Schlegel, H.B. Ascorbic acid: The chemistry underlying its antioxidant properties. Free Radic. Biol. Med. 2020, 159, 37–43. [Google Scholar] [CrossRef]
  17. Asada, K. The water-water cycle as alternative photon and electron sinks. Philos. Trans. R. Soc. Lond. B 2000, 355, 1419–1431. [Google Scholar] [CrossRef] [Green Version]
  18. Foyer, C.H.; Halliwell, B. Purification and properties of dehydroascorbate reductase from spinach leaves. Phytochemistry 1977, 16, 1347–1350. [Google Scholar] [CrossRef]
  19. Terai, Y.; Ueno, H.; Ogawa, T.; Sawa, Y.; Miyagi, A.; Kawai-Yamada, M.; Ishikawa, T.; Maruta, T. Dehydroascorbate reductases and glutathione set a threshold for high-light-induced ascorbate accumulation. Plant Physiol. 2020, 183, 112–122. [Google Scholar] [CrossRef] [Green Version]
  20. Green, M.; Fry, S. Vitamin C degradation in plant cells via enzymatic hydrolysis of 4-O-oxalyl-l-threonate. Nature 2005, 433, 83–87. [Google Scholar] [CrossRef]
  21. Truffault, V.; Fry, S.C.; Stevens, R.G.; Gautier, H. Ascorbate degradation in tomato leads to accumulation of oxalate, threonate and oxalyl threonate. Plant J. 2017, 89, 996–1008. [Google Scholar] [CrossRef] [Green Version]
  22. Fry, S.C. Oxidative scission of plant cell wall polysaccharides by ascorbate-induced hydroxyl radicals. Biochem. J. 1998, 332, 507–515. [Google Scholar] [CrossRef] [Green Version]
  23. Green, M.A.; Fry, S.C. Apoplastic degradation of ascorbate: Novel enzymes and metabolites permeating the plant cell wall. Plant Biosys. 2005, 139, 2–7. [Google Scholar] [CrossRef]
  24. Kliebenstein, D.J.; Lambrix, V.; Reichelt, M.; Gershenzon, J.; Mitchell-Olds, T. Gene duplication in the diversification of secondary metabolism: Tandem 2-oxoglutarate-dependent dioxygenases control glucosinolate biosynthesis in Arabidopsis. Plant Cell 2001, 13, 681–693. [Google Scholar]
  25. Brisson, L.; El Bakkali-Taheri, N.; Giorgi, M.; Fadel, A.; Kaizer, J.; Réglier, M.; Tron, T.; Ajandouz, E.H.; Simaan, A.J. 1-Aminocyclopropane-1-carboxylic acid oxidase: Insight into cofactor binding from experimental and theoretical studies. J. Biol. Inorg. Chem. 2012, 17, 939–949. [Google Scholar] [CrossRef]
  26. Mellor, N.; Band, L.R.; Pěnčík, A.; Novák, O.; Rashed, A.; Holman, T.; Wilson, M.H.; Voß, U.; Bishopp, A.; King, J.R.; et al. Dynamic regulation of auxin oxidase and conjugating enzymes AtDAO1 and GH3 modulates auxin homeostasis. Proc. Natl. Acad. Sci. USA 2016, 113, 11022–11027. [Google Scholar] [CrossRef] [Green Version]
  27. Monfort, A.; Wutz, A. Breathing in epigenetic change with vitamin C. EMBO Rep. 2013, 14, 337–346. [Google Scholar] [CrossRef] [Green Version]
  28. Young, J.I.; Züchner, S.; Wang, G. Regulation of the epigenome by vitamin C. Annu. Rev. Nutr. 2015, 35, 545–564. [Google Scholar] [CrossRef] [Green Version]
  29. Xue, J.H.; Chen, G.D.; Hao, F.; Chen, H.; Fang, Z.; Chen, F.-F.; Pang, B.; Yang, Q.-L.; Wei, X.; Fan, Q.-Q.; et al. A vitamin-C-derived DNA modification catalysed by an algal TET homologue. Nature 2019, 569, 581–585. [Google Scholar] [CrossRef]
  30. Brabson, J.P.; Leesang, T.; Mohammad, S.; Cimmino, L. Epigenetic regulation of genomic stability by Vitamin C. Front. Genet. 2021, 12, 675780. [Google Scholar] [CrossRef]
  31. Grillet, L.; Ouerdane, L.; Flis, P.; Hoang, M.T.T.; Isaure, M.-P.; Lobinski, R.; Curie, C.; Mari, S. Ascorbate efflux as a new strategy for iron reduction and transport in plants. J. Biol. Chem. 2014, 289, 2515–2525. [Google Scholar] [CrossRef]
  32. Makavitskaya, M.; Svistunenko, D.; Navaselsky, I.; Hryvusevich, P.; Mackievic, V.; Rabadanova, C.; Tyutereva, E.; Samokhina, V.; Straltsova, D.; Sokolik, A.; et al. Novel roles of ascorbate in plants: Induction of cytosolic Ca2+ signals and efflux from cells via anion channels. J. Exp. Bot. 2018, 69, 3477–3489. [Google Scholar] [CrossRef]
  33. Lim, B.; Smirnoff, N.; Cobbett, C.S.; Golz, J.F. Ascorbate-deficient vtc2 mutants in Arabidopsis do not exhibit decreased growth. Front. Plant Sci. 2016, 7, 1025. [Google Scholar] [CrossRef] [Green Version]
  34. Plumb, W.; Townsend, A.J.; Rasool, B.; Alomrani, S.; Razak, N.; Karpinska, B.; Ruban, A.V.; Foyer, C.H. Ascorbate-mediated regulation of growth, photoprotection, and photoinhibition in Arabidopsis thaliana. J. Exp. Bot. 2018, 69, 2823–2835. [Google Scholar] [CrossRef] [Green Version]
  35. Müller-Moulé, P.; Golan, T.; Niyogi, K.K. Ascorbate-deficient mutants of Arabidopsis grow in high light despite chronic photooxidative stress. Plant Physiol. 2004, 134, 1163–1172. [Google Scholar] [CrossRef] [Green Version]
  36. Höller, S.; Ueda, Y.; Wu, L.; Wang, Y.; Hajirezaei, M.-R.; Ghaffari, M.-R.; von Wirén, N.; Frei, M. Ascorbate biosynthesis and its involvement in stress tolerance and plant development in rice (Oryza sativa L.). Plant Mol. Biol. 2015, 88, 545–560. [Google Scholar] [CrossRef]
  37. Rosado-Souza, L.; Fernie, A.R.; Aarabi, F. Ascorbate and thiamin: Metabolic modulators in plant acclimation responses. Plants 2020, 9, 101. [Google Scholar] [CrossRef] [Green Version]
  38. Broad, R.C.; Bonneau, J.P.; Hellens, R.P.; Johnson, A.A.T. Manipulation of ascorbate biosynthetic, recycling, and regulatory pathways for improved abiotic stress tolerance in plants. Int. J. Mol. Sci. 2020, 21, 1790. [Google Scholar] [CrossRef] [Green Version]
  39. Xiao, M.; Li, Z.; Zhu, L.; Wang, J.; Zhang, B.; Zheng, F.; Zhao, B.; Zhang, H.; Wang, Y.; Zhang, Z. The multiple roles of ascorbate in the abiotic stress response of plants: Antioxidant, cofactor, and regulator. Front. Plant Sci. 2021, 12, 598173. [Google Scholar] [CrossRef]
  40. Maruta, T. How does light facilitate vitamin C biosynthesis in leaves? Biosci. Biotechnol. Biochem. 2022, 86, 1173–1182. [Google Scholar] [CrossRef]
  41. Kiyota, M.; Numayama, N.; Goto, K. Circadian rhythms of the l-ascorbic acid level in Euglena and spinach. J. Photochem. Photobiol. B Biol. 2006, 84, 197–203. [Google Scholar] [CrossRef]
  42. Dowdle, J.; Ishikawa, T.; Gatzek, S.; Rolinski, S.; Smirnoff, N. Two genes in Arabidopsis thaliana encoding GDP-l-galactose phosphorylase are required for ascorbate biosynthesis and seedling viability. Plant J. 2007, 52, 673–689. [Google Scholar] [CrossRef]
  43. Maruta, T.; Yonemitsu, M.; Yabuta, Y.; Tamoi, M.; Ishikawa, T.; Shigeoka, S. Arabidopsis Phosphomannose Isomerase 1, but not Phosphomannose Isomerase 2, is essential for ascorbic acid biosynthesis. J. Biol. Chem. 2008, 283, 28842–28851. [Google Scholar] [CrossRef] [Green Version]
  44. Müller-Moulé, P.; Conklin, P.L.; Niyogi, K.K. Ascorbate deficiency can limit violaxanthin de-epoxidase activity in vivo. Plant Physiol. 2002, 128, 970–977. [Google Scholar] [CrossRef] [Green Version]
  45. Müller-Moulé, P.; Havaux, M.; Niyogi, K.K. Zeaxanthin deficiency enhances the high light sensitivity of an ascorbate-deficient mutant of Arabidopsis. Plant Physiol. 2003, 133, 748–760. [Google Scholar] [CrossRef] [Green Version]
  46. Conklin, P.L.; Williams, E.H.; Last, R.L. Ozone-induced expression of stress sensitivity of an ascorbic acid-deficient Arabidopsis mutant. Proc. Natl. Acad. Sci. USA 1996, 93, 9970–9974. [Google Scholar] [CrossRef] [Green Version]
  47. Bellini, E.; De Tullio, M.C. Ascorbic acid and ozone: Novel perspectives to explain an elusive relationship. Plants 2019, 8, 122. [Google Scholar] [CrossRef] [Green Version]
  48. Kakan, X.; Yu, Y.; Li, S.; Li, X.; Huang, R.; Wang, J. Ascorbic acid modulation by ABI4 transcriptional repression of VTC2 in the salt tolerance of Arabidopsis. BMC Plant Biol. 2021, 21, 112. [Google Scholar] [CrossRef]
  49. Zhang, Z.; Wang, J.; Zhang, R.; Huang, R. The ethylene response factor AtERF98 enhances tolerance to salt through the transcriptional activation of ascorbic acid synthesis in Arabidopsis. Plant J. 2012, 71, 273–287. [Google Scholar] [CrossRef]
  50. Zhang, H.; Xiang, Y.; He, N.; Liu, X.; Liu, H.; Fang, L.; Zhang, F.; Sun, X.; Zhang, D.; Li, X.; et al. Enhanced vitamin C production mediated by an ABA-induced PTP-like nucleotidase improves plant drought tolerance in Arabidopsis and maize. Mol. Plant 2020, 13, 760–776. [Google Scholar] [CrossRef]
  51. Iwagami, T.; Ogawa, T.; Ishikawa, T.; Maruta, T. Activation of ascorbate metabolism by nitrogen starvation and its physiological impacts in Arabidopsis thaliana. Biosci. Biotechnol. Biochem. 2022, 86, 476–489. [Google Scholar] [CrossRef]
  52. Yabuta, Y.; Maruta, T.; Nakamura, A.; Mieda, T.; Yoshimura, K.; Ishikawa, T.; Shigeoka, S. Conversion of L-galactono-1,4-lactone to L-ascorbate is regulated by the photosynthetic electron transport chain in Arabidopsis. Biosci. Biotechnol. Biochem. 2008, 72, 2598–2607. [Google Scholar] [CrossRef] [Green Version]
  53. Bulley, S.M.; Cooney, J.M.; Laing, W. Elevating ascorbate in Arabidopsis stimulates the production of abscisic acid, phaseic acid, and to a lesser extent auxin (IAA) and jasmonates, resulting in increased expression of DHAR1 and multiple transcription factors associated with abiotic stress tolerance. Int. J. Mol. Sci. 2021, 22, 6743. [Google Scholar]
  54. Tamaoki, M.; Mukai, F.; Asai, N.; Nakajima, N.; Kubo, A.; Aono, M.; Saji, H. Light-controlled expression of a gene encoding L-galactono-γ-lactone dehydrogenase which affects ascorbate pool size in Arabidopsis thaliana. Plant Sci. 2003, 164, 1111–1117. [Google Scholar] [CrossRef]
  55. Yabuta, Y.; Mieda, T.; Rapolu, M.; Nakamura, A.; Motoki, T.; Maruta, T.; Yoshimura, K.; Ishikawa, T.; Shigeoka, S. Light regulation of ascorbate biosynthesis is dependent on the photosynthetic electron transport chain but independent of sugars in Arabidopsis. J. Exp. Bot. 2007, 58, 2661–2671. [Google Scholar] [CrossRef] [Green Version]
  56. Massot, C.; Stevens, R.; Genard, M.; Longuenesse, J.-J.; Gautier, H. Light affects ascorbate content and ascorbate-related gene expression in tomato leaves more than in fruits. Planta 2012, 235, 153–163. [Google Scholar] [CrossRef]
  57. Pallanca, J.E.; Smirnoff, N. Ascorbic acid metabolism in pea seedlings. A comparison of D-glucosone, L-sorbosone, and L-galactono-1,4-lactone as ascorbate precursors. Plant Physiol. 1999, 120, 453–461. [Google Scholar] [CrossRef] [Green Version]
  58. Nishikawa, F.; Kato, M.; Hyodo, H.; Ikoma, Y.; Sugiura, M.; Yano, M. Effect of sucrose on ascorbate level and expression of genes involved in the ascorbate biosynthesis and recycling pathway in harvested broccoli florets. J. Exp. Bot. 2005, 56, 65–72. [Google Scholar] [CrossRef] [Green Version]
  59. Badejo, A.A.; Wada, K.; Gao, Y.; Maruta, T.; Sawa, Y.; Shigeoka, S.; Ishikawa, T. Translocation and the alternative D-galacturonate pathway contribute to increasing the ascorbate level in ripening tomato fruits together with the D-mannose/L-galactose pathway. J. Exp. Bot. 2012, 63, 229–239. [Google Scholar] [CrossRef] [Green Version]
  60. Yoshimura, K.; Nakane, T.; Kume, S.; Shiomi, Y.; Maruta, T.; Ishikawa, T.; Shigeoka, S. Transient expression analysis revealed the importance of VTC2 expression level in light/dark regulation of ascorbate biosynthesis in Arabidopsis. Biosci. Biotechnol. Biochem. 2014, 78, 60–66. [Google Scholar] [CrossRef]
  61. Tanaka, H.; Maruta, T.; Tamoi, M.; Yabuta, Y.; Yoshimura, K.; Ishikawa, T.; Shigeoka, S. Transcriptional control of vitamin C defective 2 and tocopherol cyclase genes by light and plastid-derived signals: The partial involvement of GENOMES UNCOUPLED 1. Plant Sci. 2015, 231, 20–29. [Google Scholar] [CrossRef]
  62. Chen, H.; Huang, X.; Gusmaroli, G.; Terzaghi, W.; Lau, O.S.; Yanagawa, Y.; Zhang, Y.; Li, J.; Lee, J.H.; Zhu, D.; et al. Arabidopsis CULLIN4-damaged DNA binding protein 1 interacts with CONSTITUTIVELY PHOTOMORPHOGENIC1-SUPPRESSOR OF PHYA complexes to regulate photomorphogenesis and flowering time. Plant Cell 2010, 22, 108–123. [Google Scholar] [CrossRef]
  63. Nezames, C.D.; Deng, X.W. The COP9 signalosome: Its regulation of cullin-based E3 ubiquitin ligases and role in photomorphogenesis. Plant Physiol. 2012, 160, 38–46. [Google Scholar] [CrossRef] [Green Version]
  64. Wang, J.; Yu, Y.; Zhang, Z.; Quan, R.; Zhang, H.; Ma, L.; Deng, X.W.; Huang, R. Arabidopsis CSN5B interacts with VTC1 and modulates ascorbic acid synthesis. Plant Cell 2013, 25, 625–636. [Google Scholar] [CrossRef] [Green Version]
  65. Zhang, W.; Lorence, A.; Gruszewski, H.A.; Chevone, B.I.; Nessler, C.L. AMR1, an Arabidopsis gene that coordinately and negatively regulates the mannose/L-galactose ascorbic acid biosynthetic pathway. Plant Physiol. 2009, 150, 942–950. [Google Scholar] [CrossRef] [Green Version]
  66. Hu, T.; Ye, J.; Tao, P.; Li, H.; Zhang, J.; Zhang, Y.; Ye, Z. The tomato HD-Zip I transcription factor SlHZ24 modulates ascorbate accumulation through positive regulation of the D-mannose/L-galactose pathway. Plant J. 2016, 85, 16–29. [Google Scholar] [CrossRef] [Green Version]
  67. Li, J.; Liang, D.; Li, M.; Ma, F. Light and abiotic stresses regulate the expression of GDP-L-galactose phosphorylase and levels of ascorbic acid in two kiwifruit genotypes via light-responsive and stress-inducible cis-elements in their promoters. Planta 2013, 238, 535–547. [Google Scholar] [CrossRef]
  68. Fukunaga, K.; Fujikawa, Y.; Esaka, M. Light regulation of ascorbic acid biosynthesis in rice via light responsive cis-elements in genes encoding ascorbic acid biosynthetic enzymes. Biosci. Biotechnol. Biochem. 2010, 74, 888–891. [Google Scholar] [CrossRef]
  69. Conklin, P.L.; DePaolo, D.; Wintle, B.; Schatz, C.; Buckenmeyer, G. Identification of Arabidopsis VTC3 as a putative and unique dual function protein kinase::protein phosphatase involved in the regulation of the ascorbic acid pool in plants. J. Exp. Bot. 2013, 64, 2793–2804. [Google Scholar] [CrossRef] [Green Version]
  70. Leferink, N.G.; van Duijn, E.; Barendregt, A.; Heck, A.J.R.; van Berkel, W.J.H. Galactono-lactone dehydrogenase requires a redox-sensitive thiol for optimal production of vitamin C. Plant Physiol. 2009, 150, 596–605. [Google Scholar] [CrossRef] [Green Version]
  71. Paciolla, C.; Tommasi, F. The ascorbate system in two bryophytes: Brachythecium velutinum and Marchantia polymorpha. Biol. Plant. 2003, 47, 387–393. [Google Scholar] [CrossRef]
  72. Sodeyama, T.; Nishikawa, H.; Harai, K.; Takeshima, D.; Sawa, Y.; Maruta, T.; Ishikawa, T. The D-mannose/L-galactose pathway is the dominant ascorbate biosynthetic route in the moss Physcomitrium patens. Plant J. 2021, 107, 1724–1738. [Google Scholar] [CrossRef]
  73. Urzica, E.I.; Adler, L.N.; Page, M.D.; Linster, C.L.; Arbing, M.A.; Casero, D.; Pellegrini, M.; Merchant, S.S.; Clarke, S.G. Impact of oxidative stress on ascorbate biosynthesis in Chlamydomonas via regulation of the VTC2 gene encoding a GDP-L-galactose phosphorylase. J. Biol. Chem. 2012, 287, 14234–14245. [Google Scholar] [CrossRef] [Green Version]
  74. Vidal-Meireles, A.; Neupert, J.; Zsigmond, L.; Rosado-Souza, L.; Kovács, L.; Nagy, V.; Galambos, A.; Fernie, A.R.; Bock, R.; Tóth, S.Z. Regulation of ascorbate biosynthesis in green algae has evolved to enable rapid stress-induced response via the VTC2 gene encoding GDP-l-galactose phosphorylase. New Phytol. 2017, 214, 668–681. [Google Scholar] [CrossRef] [Green Version]
  75. Vidal-Meireles, A.; Tóth, D.; Kovács, L.; Neupert, J.; Tóth, S.Z. Ascorbate deficiency does not limit non-photochemical quenching in Chlamydomonas reinhardtii. Plant Physiol. 2020, 182, 597–611. [Google Scholar] [CrossRef] [Green Version]
  76. Nagy, V.; Vidal-Meireles, A.; Tengölics, R.; Rákhely, G.; Garab, G.; Kovács, L.; Tóth, S.Z. Ascorbate accumulation during sulphur deprivation and its effects on photosystem II activity and H2 production of the green alga Chlamydomonas reinhardtii. Plant. Cell Environ. 2016, 39, 1460–1472. [Google Scholar] [CrossRef] [Green Version]
  77. Nagy, V.; Vidal-Meireles, A.; Podmaniczki, A.; Szentmihályi, K.; Rákhely, G.; Zsigmond, L.; Kovács, L.; Tóth, S.Z. The mechanism of photosystem-II inactivation during sulphur deprivation-induced H2 production in Chlamydomonas reinhardtii. Plant J. 2018, 94, 548–561. [Google Scholar] [CrossRef] [Green Version]
  78. Tóth, S.Z.; Nagy, V.; Puthur, J.T.; Kovács, L.; Garab, G. The physiological role of ascorbate as photosystem II electron donor: Protection against photoinactivation in heat-stressed leaves. Plant Physiol. 2011, 156, 382–392. [Google Scholar] [CrossRef] [Green Version]
  79. Senn, M.E.; Gergoff Grozeff, G.E.; Alegre, M.L.; Barrile, F.; De Tullio, M.C.; Bartoli, C.G. Effect of mitochondrial ascorbic acid synthesis on photosynthesis. Plant Physiol. Biochem. 2016, 104, 29–35. [Google Scholar] [CrossRef]
  80. Chen, Z.; Gallie, D.R. The ascorbic acid redox state controls guard cell signaling and stomatal movement. Plant Cell 2004, 16, 1143–1162. [Google Scholar] [CrossRef] [Green Version]
  81. Kerchev, P.I.; Pellny, T.K.; Vivancos, P.D.; Kiddle, G.; Hedden, P.; Driscoll, S.; Vanacker, H.; Verrier, P.; Hancock, R.D.; Foyer, C.H. The transcription factor ABI4 Is required for the ascorbic acid-dependent regulation of growth and regulation of jasmonate-dependent defense signaling pathways in Arabidopsis. Plant Cell 2011, 23, 3319–3334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Mapson, L.W.; Isherwood, F.A.; Chen, Y.T. Biological synthesis of L-ascorbic acid: The conversion of L-galactono-γ-lactone into L-ascorbic acid by plant mitochondria. Biochem. J. 1954, 56, 21–28. [Google Scholar] [CrossRef] [PubMed]
  83. Leferink, N.G.; van den Berg, W.A.; van Berkel, W.J. L-Galactono-γ-lactone dehydrogenase from Arabidopsis thaliana, a flavoprotein involved in vitamin C biosynthesis. FEBS J. 2008, 275, 713–726. [Google Scholar] [CrossRef] [PubMed]
  84. Meyer, E.H.; Letts, J.A.; Maldonado, M. Structural insights into the assembly and the function of the plant oxidative phosphorylation system. New Phytol. 2022, 235, 1315–1329. [Google Scholar] [CrossRef]
  85. Sazanov, L.A. A giant molecular proton pump: Structure and mechanism of respiratory complex I. Nat. Rev. Mol. Cell Biol. 2015, 16, 375–388. [Google Scholar] [CrossRef]
  86. Schimmeyer, J.; Bock, R.; Meyer, E.H. L-Galactono-1,4-lactone dehydrogenase is an assembly factor of the membrane arm of mitochondrial complex I in Arabidopsis. Plant Mol. Biol. 2016, 90, 117–126. [Google Scholar] [CrossRef] [Green Version]
  87. Bartoli, C.G.; Pastori, G.M.; Foyer, C.H. Ascorbate biosynthesis in mitochondria is linked to the electron transport chain between complexes III and IV. Plant Physiol. 2000, 123, 335–344. [Google Scholar] [CrossRef] [Green Version]
  88. Welchen, E.; Gonzalez, D.H. Cytochrome c, a hub linking energy, redox, stress and signaling pathways in mitochondria and other cell compartments. Physiol. Plant. 2016, 157, 310–321. [Google Scholar] [CrossRef]
  89. Welchen, E.; Hildebrand, T.; Lewejohann, D.; Gonzalez, D.H.; Braun, H.P. Lack of cytochrome c in Arabidopsis decreases stability of Complex IV and modifies redox metabolism without affecting Complexes I and III. Biochim. Biophys. Acta 2012, 1817, 990–1001. [Google Scholar] [CrossRef] [Green Version]
  90. Bartoli, C.G.; Yu, J.; Gómez, F.; Fernández, L.; McIntosh, L.; Foyer, C.H. Inter-relationships between light and respiration in the control of ascorbic acid synthesis and accumulation in Arabidopsis thaliana leaves. J. Exp. Bot. 2006, 57, 1621–1631. [Google Scholar] [CrossRef]
  91. Nunes-Nesi, A.; Sulpice, R.; Gibon, Y.; Fernie, A.R. The enigmatic contribution of mitochondrial function in photosynthesis. J. Exp. Bot. 2008, 59, 1675–1684. [Google Scholar] [CrossRef] [Green Version]
  92. Nunes-Nesi, A.; Carrari, F.; Lytovchenko, A.; Smith, A.M.; Loureiro, M.E.; Ratcliffe, R.G.; Sweetlove, L.J.; Fernie, A.R. Enhanced photosynthetic performance and growth as a consequence of decreasing mitochondrial malate dehydrogenase activity in transgenic tomato plants. Plant Physiol. 2005, 137, 611–622. [Google Scholar] [CrossRef]
  93. Tomaz, T.; Bagard, M.; Pracharoenwattana, I.; Linden, P.; Lee, C.P.; Carroll, A.J.; Stroher, E.; Smith, S.M.; Gardestrom, P.; Millar, A.H. Mitochondrial malate dehydrogenase lowers leaf respiration and alters photorespiration and plant growth in Arabidopsis. Plant Physiol. 2010, 154, 1143–1157. [Google Scholar] [CrossRef] [Green Version]
  94. Gebicki, J.M.; Nauser, T.; Domazou, A.; Steinmann, D.; Bounds, P.L.; Koppenol, W.H. Reduction of protein radicals by GSH and ascorbate: Potential biological significance. Amino Acids 2010, 39, 1131–1137. [Google Scholar] [CrossRef]
  95. Katoh, S.; San Pietro, A. Ascorbate-supported NADP photoreduction by heated Euglena chloroplasts. Arch. Biochem. Biophys. 1967, 122, 144–152. [Google Scholar] [CrossRef]
  96. Yamashita, T.; Butler, W.L. Photoreduction and photophosphorylation with Tris-washed chloroplasts. Plant Physiol. 1968, 43, 1978–1986. [Google Scholar] [CrossRef] [Green Version]
  97. Mano, J.; Hideg, É.; Asada, K. Ascorbate in thylakoid lumen functions as an alternative electron donor to photosystem II and photosystem I. Arch. Biochem. Biophys. 2004, 429, 71–80. [Google Scholar] [CrossRef]
  98. Tóth, S.Z.; Schansker, G.; Garab, G.; Strasser, R.J. Photosynthetic electron transport activity in heat-treated barley leaves: The role of internal alternative electron donors to photosystem II. Biochim. Biophys. Acta 2007, 1767, 295–305. [Google Scholar] [CrossRef] [Green Version]
  99. Tóth, S.Z.; Puthur, J.T.; Nagy, V.; Garab, G. Experimental evidence for ascorbate-dependent electron transport in leaves with inactive oxygen-evolving complexes. Plant Physiol. 2009, 149, 1568–1578. [Google Scholar] [CrossRef]
  100. Yamane, Y.; Kashino, Y.; Koike, H.; Satoh, K. Effects of high temperatures on the photosynthetic systems in spinach: Oxygen-evolving activities, fluorescence characteristics and the denaturation process. Photosynth. Res. 1998, 57, 51–59. [Google Scholar] [CrossRef]
  101. Barra, M.; Haumann, M.; Dau, H. Specific loss of the extrinsic 18 kDa protein from photosystem II upon heating to 47 °C causes inactivation of oxygen evolution likely due to Ca release from the Mn-complex. Photosynth. Res. 2005, 84, 231–237. [Google Scholar] [CrossRef] [PubMed]
  102. Foyer, C.H.; Lelandais, M.A. A comparison of the relative rates of transport of ascorbate and glucose across the thylakoid, chloroplast and plasmalemma membranes of pea leaf mesophyll cells. J. Plant Physiol. 1996, 148, 391–398. [Google Scholar] [CrossRef]
  103. Chen, G.X.; Blubaugh, D.J.; Homann, P.H.; Golbeck, J.H.; Cheniae, G.M. Superoxide contributes to the rapid inactivation of specific secondary donors of the photosystem II reaction center during photodamage of manganese-depleted photosystem II membranes. Biochemistry 1995, 34, 2317–2332. [Google Scholar] [CrossRef] [PubMed]
  104. Spetea, C.; Hideg, E.; Vass, I. Low pH accelerates light-induced damage of photosystem II by enhancing the probability of the donor-side mechanism of photoinhibition. Biochim. Biophys. Acta 1997, 1318, 275–283. [Google Scholar] [CrossRef] [Green Version]
  105. Blubaugh, D.J.; Cheniae, G.M. Kinetics of photoinhibition in hydroxylamine-extracted photosystem II membranes: Relevance to photoactivation and sites of electron donation. Biochemistry 1990, 29, 5109–5118. [Google Scholar] [CrossRef]
  106. Jegerschöld, C.; Styring, S. Spectroscopic characterization of intermediate steps involved in donor-side-induced photoinhibition of photosystem II. Biochemistry 1996, 35, 7794–7801. [Google Scholar] [CrossRef]
  107. Lorence, A.; Chevone, B.I.; Mendes, P.; Nessler, C.L. myo-Inositol oxygenase offers a possible entry point into plant ascorbate biosynthesis. Plant Physiol. 2004, 134, 1200–1205. [Google Scholar] [CrossRef] [Green Version]
  108. Tóth, S.Z.; Schansker, G.; Garab, G. The physiological roles and metabolism of ascorbate in chloroplasts. Physiol. Plant. 2013, 148, 161–175. [Google Scholar] [CrossRef]
  109. Ivanov, B.N.; Asada, K.; Kramer, D.; Edwards, G. Characterization of photosynthetic electron transport in bundle sheath cells of maize. I. Ascorbate effectively stimulates cyclic electron flow around PSI. Planta 2005, 20, 572–581. [Google Scholar] [CrossRef]
  110. Ivanov, B.; Asada, K.; Edwards, G.E. Analysis of donors of electrons to photosystem I and cyclic electron flow by redox kinetics of P700 in chloroplasts of isolated bundle sheath strands of maize. Photosynth. Res. 2007, 92, 65–74. [Google Scholar] [CrossRef]
  111. Ivanov, B.N. Role of ascorbic acid in photosynthesis. Biochemistry 2014, 79, 282–289. [Google Scholar] [CrossRef]
  112. Trubitsin, B.V.; Mamedov, M.D.; Semenov, A.Y.; Tikhonov, A.N. Interaction of ascorbate with photosystem I. Photosynth. Res. 2014, 122, 215–231. [Google Scholar] [CrossRef]
  113. Ivanov, B.N.; Sacksteder, C.A.; Kramer, D.M.; Edwards, G.E. Light-Induced ascorbate-dependent electron transport and membrane energization in chloroplasts of bundle sheath cells of the C4 plant maize. Arch. Biochem. Biophys. 2001, 385, 145–153. [Google Scholar] [CrossRef]
  114. Tóth, S.Z.; Lőrincz, T.; Szarka, A. Concentration does matter: The beneficial and potentially harmful effects of ascorbate in humans and plants. Antioxid. Redox Signal. 2018, 29, 1516–1533. [Google Scholar] [CrossRef] [Green Version]
  115. Zechmann, B.; Stumpe, M.; Mauch, F. Immunocytochemical determination of the subcellular distribution of ascorbate in plants. Planta 2011, 233, 1–12. [Google Scholar] [CrossRef] [Green Version]
  116. Razeghifard, M.R.; Klughammer, C.; Pace, R.J. Electron paramagnetic resonance kinetic studies of the S states in spinach thylakoids. Biochemistry 1997, 36, 86–92. [Google Scholar] [CrossRef]
  117. Godaux, D.; Bailleul, B.; Berne, N.; Cardol, P. Induction of photosynthetic carbon fixation in anoxia relies on hydrogenase activity and Proton-Gradient Regulation-Like1-mediated cyclic electron flow in Chlamydomonas reinhardtii. Plant Physiol. 2015, 168, 648–658. [Google Scholar] [CrossRef] [Green Version]
  118. Kim, J.P.; Kim, K.-R.; Choi, S.P.; Han, S.J.; Kim, M.S.; Sim, S.J. Repeated production of hydrogen by sulfate re-addition in sulfur deprived culture of Chlamydomonas reinhardtii. Int. J. Hydrogen Energy 2010, 35, 13387–13391. [Google Scholar] [CrossRef]
  119. Tamura, N.; Inoue, H.; Inoue, Y. Inactivation of the water-oxidizing complex by exogenous reductants in PSII membranes depleted of extrinsic proteins. Plant Cell Physiol. 1990, 31, 469–477. [Google Scholar]
  120. Anderson, J.M.; Chow, W.S.; De Las Rivas, J. Dynamic flexibility in the structure and function of photosystem II in higher plant thylakoid membranes: The grana enigma. Photosynth. Res. 2008, 98, 575–587. [Google Scholar] [CrossRef] [Green Version]
  121. Ifuku, K. Localization and functional characterization of the extrinsic subunits of photosystem II: An update. Biosci. Biotechnol. Biochem. 2015, 79, 1223–1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Nagao, R.; Suzuki, T.; Okumura, A.; Niikura, A.; Iwai, M.; Dohmae, N.; Tomo, T.; Shen, J.-R.; Ikeuchi, M.; Enami, I. Topological analysis of the extrinsic PsbO, PsbP and PsbQ proteins in a green algal PSII complex by cross-linking with a water-soluble carbodiimide. Plant Cell Physiol. 2010, 51, 718–727. [Google Scholar] [CrossRef] [PubMed]
  123. Allahverdiyeva, Y.; Suorsa, M.; Rossi, F.; Pavesi, A.; Kater, M.M.; Antonacci, A.; Tadini, L.; Pribil, M.; Schneider, A.; Wanner, G.; et al. Arabidopsis plants lacking PsbQ and PsbR subunits of the oxygen-evolving complex show altered PSII super-complex organization and short-term adaptive mechanisms. Plant J. 2013, 75, 671–684. [Google Scholar] [CrossRef]
  124. Gest, N.; Gautier, H.; Stevens, R. Ascorbate as seen through plant evolution: The rise of a successful molecule? J. Exp. Bot. 2013, 64, 33–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Podmaniczki, A.; Nagy, V.; Vidal-Meireles, A.; Tóth, D.; Patai, R.; Kovács, L.; Tóth, S.Z. Ascorbate inactivates the oxygen-evolving complex in prolonged darkness. Physiol. Plant. 2021, 171, 232–245. [Google Scholar] [CrossRef]
  126. Kirchhoff, H.; Hall, C.; Wood, M.; Herbstová, M.; Tsabari, O.; Nevo, R.; Charuvi, D.; Shimoni, E.; Reich, Z. Dynamic control of protein diffusion within the granal thylakoid lumen. Proc. Natl. Acad. Sci. USA 2011, 108, 20248–20253. [Google Scholar] [CrossRef] [Green Version]
  127. Castro, J.L.S.; Lima-Melo, Y.; Carvalho, F.E.L.; Feitosa, A.G.S.; Lima Neto, M.C.; Caverzan, A.; Margis-Pinheiro, M.; Silveira, J.A.G. Ascorbic acid toxicity is related to oxidative stress and enhanced by high light and knockdown of chloroplast ascorbate peroxidases in rice plants. Theor. Exp. Plant Physiol. 2018, 30, 41–55. [Google Scholar] [CrossRef] [Green Version]
  128. Bassi, R.; Dall’Osto, L. Dissipation of light energy absorbed in excess: The molecular mechanisms. Ann. Rev. Plant Biol. 2021, 72, 47–76. [Google Scholar] [CrossRef]
  129. Krieger-Liszkay, A.; Shimakawa, G. Regulation of the generation of reactive oxygen species during photosynthetic electron transport. Biochem. Soc. Trans. 2022, 50, 1025–1034. [Google Scholar] [CrossRef]
  130. Mittler, R.; Zandalinas, S.I.; Fichman, Y.; Van Breusegem, F. Reactive oxygen species signalling in plant stress responses. Nat. Rev. Mol. Cell Biol. 2022, 23, 663–679. [Google Scholar] [CrossRef]
  131. Pinnola, A.; Bassi, R. Molecular mechanisms involved in plant photoprotection. Biochem. Soc. Trans. 2018, 46, 467–482. [Google Scholar] [CrossRef]
  132. Kaiser, E.; Correa-Galvis, V.; Armbruster, U. Efficient photosynthesis in dynamic light environments: A chloroplast’s perspective. Biochem. J. 2019, 476, 2725–2741. [Google Scholar] [CrossRef] [Green Version]
  133. Correa-Galvis, V.; Poschmann, G.; Melzer, M.; Stühler, K.; Jahns, P. PsbS interactions involved in the activation of energy dissipation in Arabidopsis. Nat. Plants 2016, 2, 15225. [Google Scholar] [CrossRef]
  134. Sacharz, J.; Giovagnetti, V.; Ungerer, P.; Mastroianni, G.; Ruban, A.V. The xanthophyll cycle affects reversible interactions between PsbS and light-harvesting complex II to control non-photochemical quenching. Nat. Plants 2017, 3, 16225. [Google Scholar] [CrossRef]
  135. Dall’Osto, L.; Caffarri, S.; Bassi, R. A mechanism of nonphotochemical energy dissipation, independent from PsbS, revealed by a conformational change in the antenna protein CP26. Plant Cell 2005, 17, 1217–1232. [Google Scholar] [CrossRef] [Green Version]
  136. Nilkens, M.; Kress, E.; Lambrev, P.; Miloslavina, Y.; Müller, M.; Holzwarth, A.R.; Jahns, P. Identification of a slowly inducible zeaxanthin-dependent component of non-photochemical quenching of chlorophyll fluorescence generated under steady-state conditions in Arabidopsis. Biochim. Biophys. Acta 2010, 1797, 466–475. [Google Scholar] [CrossRef] [Green Version]
  137. Jahns, P.; Latowski, D.; Strzalka, K. Mechanism and regulation of the violaxanthin cycle: The role of antenna proteins and membrane lipids. Biochim. Biophys. Acta 2009, 1787, 3–14. [Google Scholar] [CrossRef] [Green Version]
  138. Bratt, C.; Arvidsson, P.; Carlsson, M.; Akerlund, H. Regulation of violaxanthin de-epoxidase activity by pH and ascorbate. Photosynth. Res. 1995, 45, 169–175. [Google Scholar] [CrossRef]
  139. Pfündel, E.E.; Renganathan, M.; Gilmore, A.M.; Yamamoto, H.Y.; Dilley, R.A. Intrathylakoid pH in isolated pea chloroplasts as probed by violaxanthin deepoxidation. Plant Physiol. 1994, 106, 1647–1658. [Google Scholar] [CrossRef] [Green Version]
  140. Hieber, A.D.; Bugos, R.C.; Verhoeven, A.S.; Yamamoto, H.Y. Overexpression of violaxanthin de-epoxidase: Properties of C-terminal deletions on activity and pH-dependent lipid binding. Planta 2002, 214, 476–483. [Google Scholar] [CrossRef]
  141. Arnoux, P.; Morosinotto, T.; Saga, G.; Bassi, R.; Pignol, D. A structural basis for the pH-dependent xanthophyll cycle in Arabidopsis thaliana. Plant Cell 2009, 21, 2036–2044. [Google Scholar] [CrossRef] [Green Version]
  142. Hallin, E.I.; Hasan, M.; Guo, K.; Åkerlund, H.-E. Molecular studies on structural changes and oligomerisation of violaxanthin de-epoxidase associated with the pH-dependent activation. Photosynth. Res. 2016, 129, 29–41. [Google Scholar] [CrossRef]
  143. Saga, G.; Giorgetti, A.; Fufezan, C.; Giacometti, G.M.; Bassi, R.; Morosinotto, T. Mutation analysis of violaxanthin de-epoxidase identifies substrate-binding sites and residues involved in catalysis. J. Biol. Chem. 2010, 285, 23763–23770. [Google Scholar] [CrossRef]
  144. Hartel, H.; Lokstein, H.; Grimm, B.; Rank, B. Kinetic studies on the xanthophyll cycle in barley leaves (Influence of antenna size and relations to nonphotochemical chlorophyll fluorescence quenching). Plant Physiol. 1996, 110, 471–482. [Google Scholar] [CrossRef] [Green Version]
  145. De Souza, A.P.; Burgess, S.J.; Doran, L.; Hansen, J.; Manukyan, L.; Maryn, N.; Gotarkar, D.; Leonelli, L.; Niyogi, K.K.; Long, S.P. Soybean photosynthesis and crop yield are improved by accelerating recovery from photoprotection. Science 2022, 377, 851–854. [Google Scholar] [CrossRef] [PubMed]
  146. Havaux, M.; Dall’Osto, L.; Bassi, R. Zeaxanthin has enhanced antioxidant capacity with respect to all other xanthophylls in Arabidopsis leaves and functions independent of binding to PSII antennae. Plant Physiol. 2007, 145, 1506–1520. [Google Scholar] [CrossRef] [Green Version]
  147. Havaux, M. Carotenoids as membrane stabilizers in chloroplasts. Trends Plant Sci. 1998, 3, 147–151. [Google Scholar] [CrossRef]
  148. Bethmann, S.; Melzer, M.; Schwarz, N.; Jahns, P. The zeaxanthin epoxidase is degraded along with the D1 protein during photoinhibition of photosystem II. Plant Direct 2019, 3, e00185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Li, Z.; Peers, G.; Dent, R.M.; Bai, Y.; Yang, S.Y.; Apel, W.; Leonelli, L.; Niyogi, K.K. Evolution of an atypical de-epoxidase for photoprotection in the green lineage. Nat. Plants 2016, 2, 16140. [Google Scholar] [CrossRef] [Green Version]
  150. Coesel, S.; Oborník, M.; Varela, J.; Falciatore, A.; Bowler, C. Evolutionary origins and functions of the carotenoid biosynthetic pathway in marine diatoms. PLoS ONE 2008, 3, e2896. [Google Scholar] [CrossRef] [Green Version]
  151. Girolomoni, L.; Bellamoli, F.; de la Cruz Valbuena, G.; Perozeni, F.; D’Andrea, C.; Cerullo, G.; Cazzaniga, S.; Ballottari, M. Evolutionary divergence of photoprotection in the green algal lineage: A plant-like violaxanthin de-epoxidase enzyme activates the xanthophyll cycle in the green alga Chlorella vulgaris modulating photoprotection. New Phytol. 2020, 228, 136–150. [Google Scholar] [CrossRef]
  152. Goss, R.; Jakob, T. Regulation and function of xanthophyll cycle-dependent photoprotection in algae. Photosynth. Res. 2010, 106, 103–122. [Google Scholar] [CrossRef]
  153. Dambek, M.; Eilers, U.; Breitenbach, J.; Steiger, S.; Büchel, C.; Sandmann, G. Biosynthesis of fucoxanthin and diadinoxanthin and function of initial pathway genes in Phaeodactylum tricornutum. J. Exp. Bot. 2012, 63, 5607–5612. [Google Scholar] [CrossRef]
  154. Grouneva, I.; Jakob, T.; Wilhelm, C.; Goss, R. Influence of ascorbate and pH on the activity of the diatom xanthophyll cycle-enzyme diadinoxanthin de-epoxidase. Physiol. Plant. 2006, 126, 205–211. [Google Scholar] [CrossRef]
  155. Dautermann, O.; Lyska, D.; Andersen-Ranberg, J.; Becker, M.; Fröhlich-Nowoisky, J.; Gartmann, H.; Kramer, L.C.; Mayr, K.; Pieper, D.; Rij, L.M.; et al. An algal enzyme required for biosynthesis of the most abundant marine carotenoids. Sci. Adv. 2020, 6, eaaw9183. [Google Scholar] [CrossRef] [Green Version]
  156. Pinnola, A.; Dall’Osto, L.; Gerotto, C.; Morosinotto, T.; Bassi, R.; Alboresi, A. Zeaxanthin binds to Light-Harvesting Complex Stress-Related Protein to enhance nonphotochemical quenching in Physcomitrella patens. Plant Cell 2013, 25, 3519–3534. [Google Scholar] [CrossRef] [Green Version]
  157. Ivanov, B.; Borisova-Mubarakshina, M.; Vilyanen, D.; Vetoshkina, D.; Kozuleva, M. Cooperative pathway of O2 reduction to H2O2 in chloroplast thylakoid membrane: New insight into the Mehler reaction. Biophys. Rev. 2022, 14, 857–869. [Google Scholar] [CrossRef]
  158. 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]
  159. Pastori, G.M.; Kiddle, G.; Antoniw, J.; Bernard, S.; Veljovic-Jovanovic, S.; Verrier, P.J.; Noctor, G.; Foyer, C.H. Leaf vitamin C contents modulate plant defense transcripts and regulate genes that control development through hormone signaling. Plant Cell 2003, 15, 939–951. [Google Scholar] [CrossRef] [Green Version]
  160. 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] [Green Version]
  161. Luschin-Ebengreuth, N.; Zechmann, B. Compartment-specific investigations of antioxidants and hydrogen peroxide in leaves of Arabidopsis thaliana during dark-induced senescence. Acta Physiol. Plant. 2016, 38, 133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Farrow, S.C.; Facchini, P.J. Functional diversity of 2-oxoglutarate/Fe(II)-dependent dioxygenases in plant metabolism. Front. Plant Sci. 2004, 5, 524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. 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] [PubMed]
  164. Hoang, M.T.T.; Almeida, D.; Chay, S.; Alcon, C.; Corratge-Faillie, C.; Curie, C.; Mari, S. AtDTX25, a member of the multidrug and toxic compound extrusion family, is a vacuolar ascorbate transporter that controls intracellular iron cycling in Arabidopsis. New Phytol. 2021, 231, 1956–1967. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Ascorbate (Asc) biosynthesis and the roles of chloroplastic Asc in vascular plants. Asc is synthesized via the Smirnoff-Wheeler pathway with the majority of the steps taking place in the cytosol and the last step in the mitochondria at Complex I. Asc biosynthesis is regulated by light, circadian clock, photosynthetic electron transport, and a feedback inhibition by Asc. Asc is found in all cell compartments; PHT4;4 transports it into the chloroplast. Chloroplastic Asc plays multiple roles: (i) it is an alternative electron donor of photosystem II (PSII) when the oxygen-evolving complex (OEC) is inactive, (ii) it may inactivate the OEC under specific circumstances, (iii) it is a co-substrate of violaxanthin de-epoxidase (VDE) thereby plays a role in non-photochemical quenching, (iv) Asc participates in reactive oxygen species management. Abbreviations: APX, Asc peroxidase; cp, chloroplast; cs, cytosol; Cytb6f, cytochrome b6f; Cytc, cytochrome c; DHA, dehydroascorbate; DHAR, dehydroascorbate reductase; DTX25, vacuolar ascorbate transporter; Fd, ferredoxin; FNR, ferredoxin-NADP oxidoreductase; GDH, L-galactose dehydrogenase; GGP, GDP-L-galactose phosphorylase (encoded by VTC2); GLDH, L-galactono-1,4 lactone dehydrogenase; GME, GDP-D-mannose 3′,5′-epimerase; GMP, GDP-D-mannose phosphorylase; GPP, L-galactose-1-P phosphatase; GR, glutathione reductase; L-Gal-L, L-galactono-lactone; LHC, Light-harvesting complex; m, mitochondrion; MDA, monodehydroascorbate; MDAR Monodehydroascorbate reductase; n, nucleus; PHT4;4, chloroplastic Asc transporter; pm, plasma membrane; PSI, photosystem I; SOD, superoxide dismutase; t, thylakoid; TCA, tricarboxylic acid; UQ, ubiquitine; v, vacuole.
Figure 1. Ascorbate (Asc) biosynthesis and the roles of chloroplastic Asc in vascular plants. Asc is synthesized via the Smirnoff-Wheeler pathway with the majority of the steps taking place in the cytosol and the last step in the mitochondria at Complex I. Asc biosynthesis is regulated by light, circadian clock, photosynthetic electron transport, and a feedback inhibition by Asc. Asc is found in all cell compartments; PHT4;4 transports it into the chloroplast. Chloroplastic Asc plays multiple roles: (i) it is an alternative electron donor of photosystem II (PSII) when the oxygen-evolving complex (OEC) is inactive, (ii) it may inactivate the OEC under specific circumstances, (iii) it is a co-substrate of violaxanthin de-epoxidase (VDE) thereby plays a role in non-photochemical quenching, (iv) Asc participates in reactive oxygen species management. Abbreviations: APX, Asc peroxidase; cp, chloroplast; cs, cytosol; Cytb6f, cytochrome b6f; Cytc, cytochrome c; DHA, dehydroascorbate; DHAR, dehydroascorbate reductase; DTX25, vacuolar ascorbate transporter; Fd, ferredoxin; FNR, ferredoxin-NADP oxidoreductase; GDH, L-galactose dehydrogenase; GGP, GDP-L-galactose phosphorylase (encoded by VTC2); GLDH, L-galactono-1,4 lactone dehydrogenase; GME, GDP-D-mannose 3′,5′-epimerase; GMP, GDP-D-mannose phosphorylase; GPP, L-galactose-1-P phosphatase; GR, glutathione reductase; L-Gal-L, L-galactono-lactone; LHC, Light-harvesting complex; m, mitochondrion; MDA, monodehydroascorbate; MDAR Monodehydroascorbate reductase; n, nucleus; PHT4;4, chloroplastic Asc transporter; pm, plasma membrane; PSI, photosystem I; SOD, superoxide dismutase; t, thylakoid; TCA, tricarboxylic acid; UQ, ubiquitine; v, vacuole.
Ijms 24 02537 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tóth, S.Z. The Functions of Chloroplastic Ascorbate in Vascular Plants and Algae. Int. J. Mol. Sci. 2023, 24, 2537. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032537

AMA Style

Tóth SZ. The Functions of Chloroplastic Ascorbate in Vascular Plants and Algae. International Journal of Molecular Sciences. 2023; 24(3):2537. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032537

Chicago/Turabian Style

Tóth, Szilvia Z. 2023. "The Functions of Chloroplastic Ascorbate in Vascular Plants and Algae" International Journal of Molecular Sciences 24, no. 3: 2537. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032537

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