Next Article in Journal
Trends and Limits for Quinoa Production and Promotion in Pakistan
Previous Article in Journal
Impact of Horse Grazing on Floristic Diversity in Mediterranean Small Standing-Water Ecosystems (SWEs)
Previous Article in Special Issue
Effect of Paper and Aluminum Bagging on Fruit Quality of Loquat (Eriobotrya japonica Lindl.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metabolism and Regulation of Ascorbic Acid in Fruits

1
Key Laboratory of Plant Hormones and Development Regulation of Chongqing, School of Life Sciences, Chongqing University, Chongqing 400044, China
2
Institute of Horticulture, Chengdu Academy of Agriculture and Forestry Sciences, Chengdu 611130, China
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(12), 1602; https://doi.org/10.3390/plants11121602
Submission received: 30 April 2022 / Revised: 26 May 2022 / Accepted: 14 June 2022 / Published: 18 June 2022
(This article belongs to the Special Issue Advance in Fruit Development)

Abstract

:
Ascorbic acid, also known as vitamin C, is a vital antioxidant widely found in plants. Plant fruits are rich in ascorbic acid and are the primary source of human intake of ascorbic acid. Ascorbic acid affects fruit ripening and stress resistance and plays an essential regulatory role in fruit development and postharvest storage. The ascorbic acid metabolic pathway in plants has been extensively studied. Ascorbic acid accumulation in fruits can be effectively regulated by genetic engineering technology. The accumulation of ascorbic acid in fruits is regulated by transcription factors, protein interactions, phytohormones, and environmental factors, but the research on the regulatory mechanism is still relatively weak. This paper systematically reviews the regulation mechanism of ascorbic acid metabolism in fruits in recent decades. It provides a rich theoretical basis for an in-depth study of the critical role of ascorbic acid in fruits and the cultivation of fruits rich in ascorbic acid.

1. Introduction

Ascorbic acid (AsA), also known as vitamin C, is an important antioxidant widely found in plants. AsA is closely related to human health. A lack of AsA can lead to scurvy [1]. In addition, AsA also plays an important role in the defense of cardiovascular diseases, cataracts, cancer, aging, and other diseases [2,3]. Since the human body lacks the gene encoding L-guluronic acid-1,4-lactone oxidase, the last step in the AsA synthesis pathway, it cannot synthesize AsA [4]. Therefore, fruit is the main source of human AsA intake.
As an important reactive oxygen species scavenger, AsA plays an important role in plant resistance to biotic stress. The increase in endogenous ascorbic acid content can improve the resistance of potato to Phytophthora infestans [5]. Seed priming with salicylic acid treatment enhanced tomato resistance to Fusarium wilt [6]. In addition to biotic stress conditions, more studies have shown that AsA played an important role in abiotic stress resistance [7,8]. The reactive oxygen species produced by drought, salinity, freezing, and other abiotic stresses can be eliminated by AsA, thereby protecting cells from damage [9,10,11]. Exogenously applied AsA can alleviate the adverse effects of various abiotic stresses and promote plant growth [12,13].
AsA plays an important role in the postharvest low-temperature storage of fruits. Exogenous AsA treatment can significantly improve the cold resistance of banana fruit and strawberry fruit under low-temperature storage [14,15]. Changes in AsA content can also affect fruit ripening [16]. In addition, AsA has a certain relationship with fruit development. After reducing the AsA content in tomato fruit, the size of tomato fruit is significantly reduced [17]. Understanding the metabolism and regulatory mechanism of AsA in fruit has important guiding significance for fruit quality improvement and postharvest preservation. This paper systematically reviews the AsA metabolism and regulatory mechanism in fruits in recent decades.

2. Metabolic Pathway of Ascorbic Acid

The main dietary sources of AsA for humans are fruits. The AsA contents of some fruits are organized, such as tomato, kiwifruit, orange, strawberry, carrot, sweet pepper, and so on (Table 1) [18,19,20,21,22,23,24,25,26,27,28,29,30]. Because of the important role of AsA in plants, in-depth studies have been carried out on the metabolic pathways of AsA in plants (Figure 1).

2.1. Biosynthesis of Ascorbic Acid

The AsA synthesis has been demonstrated to occur in the mitochondria via several proposed routes. The L-galactose pathway, also called the Smirnoff–Wheeler pathway, is the major biosynthesis pathway [31]. Arabidopsis is a model plant for gene function research, and most genes and enzymes involved in AsA biosynthesis have been well characterized and described. In the L-galactose pathway, D-glucose-6-phosphate is first converted to D-fructose-6-phosphate by phosphoglucose isomerase (PGI) [4]. Then phosphomannose isomerase (PMI) catalyzes the conversion of D-fructose-6-phosphate to D-Mannose-6-phosphate, which is then converted to D-mannose-1-phosphate by phosphomannomutase (PMM) [32,33]. GDP-mannose pyrophosphorylase (GMP), named VTC1 in Arabidopsis, catalyzes a rate-limiting step in AsA synthesis that converts D-mannose-1-phosphate to GDP-D-mannose [34]. GMP gene family has been identified in several horticultural crops, including tomato, apple, and pear [35,36,37]. In the next step, GDP-D-mannose3′, 5′-epimerase (GME) catalyzes GDP-D-mannose conversion to GDP-L-galactose [38]. The first dedicated step in the L-galactose pathway is accomplished by GDP-L-galactose-phosphorylase (GGP), which converts GDP-L-galactose to L-galactose-1-phosphate [39]. In Arabidopsis, GGP is encoded by the VTC2 and VTC5 gene and low content of AsA is presented in vtc2 and vtc5 mutants [40]. Then the removal of the phosphate group by L-galactose-1-phosphate phosphatase (GPP) to L-galactose-1-phosphate produces L-galactose [41]. Although GPP in Arabidopsis (encoded by the VTC4) has phosphatase activity, AsA is still accumulated in vtc4 mutants, suggesting that GPP is not a key rate-limiting enzyme in AsA biosynthesis [42,43]. Lastly, L-galactose dehydrogenase (GDH) converts L-galactose to L-galactono-1,4-lactone, which is finally restored to AsA by L-galactono-1,4-lactone dehydrogenase (GLDH) [41,44]. Almost all enzymes involved in AsA biosynthesis are located in the cytosol, but GLDH is located in the mitochondrial inner membrane, indicating that AsA is finally synthesized in the mitochondria [45,46].
In plants, the L-galactose pathway is the major but not the only pathway leading to AsA synthesis. Studies have found that, in addition to catalyzing GDP-D-mannose to produce GDP-L-galactose, GME can also catalyze GDP-D-mannose to GDP-L-gulose [47]. In the L-gulose pathway, GDP-L-gulose is successively converted to L-Gulose-1-phosphate, followed by L-Gulose, which can be converted into L-gulono-1,4-lactone [48]. The L-gulono-1,4-lactone is finally converted to AsA by L-gulono-1,4-lactone oxidase (GulLO), a paralogous gene of GLDH [49]. Two GulLO genes in Arabidopsis (AtGulLO3 and AtGulLO5) have been characterized [50]. However, there are almost no reports on the characterization of GulLO in fruits.
L-gulono-1,4-lactone oxidase is not only the last precursor of AsA biosynthesis in L-gulose pathway, but also the last precursor of AsA biosynthesis in the myoinositol pathway [51]. Like in animals, myoinositol can be converted to D-glucuronate by myoinositol oxygenase (MIOX) in plants [52]. It has been reported that overexpression of AtMIOX4 can increase the AsA content in Arabidopsis [51]. D-glucuronate is then sequentially converted to L-gulonate and L-gulono-1,4-lactone [4].
In addition, the D-galacturonate pathway is also a potential pathway of AsA biosynthesis [48]. Methyl-D-galacturonate is one of the cell wall degradation products and can be converted to D-galacturonate by methylesterase. Then, D-galacturonate reductase (GalUR) catalyzes the conversion of D-galacturonate to l-galactonate, which is then converted to L-galactono-1,4-lactone by aldono-lactonase (Alase) [53]. However, direct evidence for the characterization of Alase in plants is still missing. Although previous studies have shown that feeding L-galactose and D-galacturonate can enhance the content of AsA in red ripened fruits of tomatoes, there is a lack of sufficient genetic evidence to illustrate this pathway in AsA biosynthesis [54].

2.2. Regeneration and Degradation of Ascorbic Acid

Aside from the above biosynthesis pathways, regeneration of AsA through the Foyer-Halliwell–Asada cycle is also an approach for AsA production [55]. Ascorbate peroxidase (APX) catalyzes the conversion of AsA to monodehydroascorbic acid (MDHA), which is then converted to dehydroascorbic acid (DHA) by a nonenzymatic disproportionation [3]. Ascorbate oxidas (AO) is an ascorbate oxidase, and part of AsA can also form MDHA under the catalysis of AO. Meanwhile, both MDHA and DHA can be converted to AsA by monodehydroascorbate reductase (MDHAR) and dehydroascorbate reductase (DHAR), respectively [3]. Much evidence has verified that DHAR positively controls AsA production in some horticultural crops, such as tomatoes and potatoes [56,57]. Although MDHAR also governs AsA regeneration, the expression level of MDHAR does not always coincide with AsA content. Transgenic plans are overexpressing MDHAR to decrease AsA accumulation, while knockdown of MDHAR gene decreases the AsA content [56,58,59]. However, overexpression of acerola MDHAR in tobacco significantly increases AsA content [60].
AsA regeneration is often closely associated with environmental stresses. AsA will be degraded through multiple pathways in apoplast when supernumerary AsA is no longer needed in plants [48]. On the one hand, DHA is hydrolyzed to 2,3-diketo-gulonic acid, which is then hydrolyzed to multiple products, including L-threarate, oxalate, and tartaric acid [61,62,63]. On the other hand, DHA can be oxidated directly to L-threarate and oxalate [64]. Moreover, DHA can also be oxidated to 4-O-oxalyl-L-threonate, which is then oxidated to L-threarate and oxalate [64]. Tartaric acid is an essential determinant of the fruit quality of the grape. Still, it cannot be detected from degradation products, indicating that oxidation of DHA is the primary process involved in AsA degradation [65].

2.3. Ascorbic Acid Transport

Although AsA is known to be synthesized in the mitochondrial inner membrane, AsA can also be detected in other subcellular structures, such as apoplast, chloroplast and vacuole, indicating the presence of AsA transmembrane transport [66,67,68] (Figure 2). At the outset, it is often assumed that AsA may be released from mitochondria into the cytosol via simple diffusion [69]. However, at physiological pH values, AsA has no membrane permeability [70]. Interestingly, it has been reported that mitochondrial ascorbic acid transporter from potato has unidirectional AsA transport activity, demonstrating the existence of transport proteins in mitochondria [71]. Currently, it is known that AsA is transferred from cytosol to chloroplast in a carrier-mediated manner [69]. In plants, AtPHT4;4 is the first chloroplast-localized ascorbate transporter identified from Arabidopsis [72]. In addition, some MDHAR and DHAR have also been shown to localize at chloroplast, suggesting that DHA can be taken up by chloroplast [73,74]. However, the mechanism of DHA uptake is still unknown. AsA utilizes concentration and pH gradients to enter the vacuole and thylakoid by passive diffusion [75]. Due to the missing AsA regenerated enzymes in the apoplast, AsA transport between protoplast and apoplast is necessary [66]. In mammals, AsA and DHA crossing the plasma membrane are mediated by Na-dependent transporters and glucose transporters [76,77]. However, the carriers that help AsA and DHA across the plasma membrane have not yet been identified [3]. Some nucleobase/ascorbate transporter (NAT) homologous genes have been identified and have functional redundancy according to the phenotype of multiple mutants. However, the regulation mechanism of NAT genes on AsA transmembrane transport is still unclear [78]. Furthermore, MDHA can reduce to AsA by accepting electrons from cytochrome b in the apoplast. MDHA can reduce to AsA by cytochrome b, which provides electrons from protoplast AsA oxidation [79].
AsA biosynthesis occurs in various tissues of plants, but some evidence shows that AsA has long-distance transport between tissues. AsA biosynthesis-related enzymes can be detected in the phloem of Cucurbita pepo fruits, indicating that AsA may be transported to other tissues through the phloem [80]. By labeling AsA with C-14, Franceschi and Tarlyn find that AsA can be transported from leaf to flower and root tip [81]. In addition, the AsA in the phloem and tubers of potato will be enriched with the increase in the AsA content in the leaves [82]. The inter-tissue transport of AsA still needs more evidence to support it.

3. Transcriptional Regulation of Ascorbic Acid Metabolic Genes in Fruit

Ascorbic acid is enriched in fruit and is closely related to fruit development and resistance. Changing the content of ascorbic acid in fruit is of great significance for improving fruit quality. Tomato, strawberry, kiwifruit, citrus, and other fruits are rich in ascorbic acid, with complete genome information and a mature genetic transformation system, which are the main objects of research related to the regulation of ascorbic acid content. For example, there were significant differences in the expression levels of AsA synthesis and metabolism-related genes in citrus fruits with different ascorbic acid contents [83,84]. It is of great significance to improve the ascorbic acid content in fruit by changing the expression of genes related to AsA synthesis and metabolism.
Substrate feeding experiments have shown that AsA synthesis in tomato leaves mainly depends on the L-galactose pathway, while the L-galactose pathway, the D-galacturonate pathway, and the myo-inositol pathway are all involved in AsA synthesis in tomato fruit [35]. Many genes encoding key enzymes in the AsA metabolism pathway have been identified. The content of AsA is significantly changed in fruits when the expression of these genes is altered through genetic engineering (Table 2). GMP is a key rate-limiting enzyme in the L-galactose pathway. There are four members of the GMP family in tomato (SlGMP1-SlGMP4). Upregulation or downregulation of SlGMP3 expression in tomato fruit can significantly increase or decrease the content of AsA [85]. Overexpression of tomato SlGME1 and SlGME2 increase AsA content to 1.60- and 1.24- fold in ripe red fruit, respectively [86]. However, silencing SlGME1 and SlGME2 alone does not affect AsA content in tomato fruit [87]. Conversely, simultaneous silencing of both SlGMEs genes significantly reduces AsA content in tomato fruit, suggesting a functional redundancy between SlGME1 and SlGME2 [38]. Overexpression of GGP significantly increases AsA content in tomato leaves but does not affect AsA content in fruit [35]. However, studies have also shown that the AsA content in leaves, flowers, and fruits is significantly reduced in tomato SlGGP1 mutants [88,89]. Heterologous expression of the GGP gene of kiwifruit in tomato and strawberry results in a twofold increase in AsA content in the fruits [90]. In addition, overexpression of AcGGP3 can significantly increase the AsA content in leaves and fruits of kiwifruit, indicating that the role of GGP in the AsA synthesis pathway is tissue-specific in different plants [91,92]. GPP catalyzes the dephosphorylation of L-galactose-1-phosphate to L-galactose, an essential part of the L-galactose pathway. However, no significant changes in AsA content are found in the fruits of the GGP overexpression line and the GGP × GPP pyramiding lines [35]. A recent study by Zheng et al. indicates that overexpression of SlIMP3, a gene encoding a bifunctional enzyme with GPP and IMP activities, significantly increased AsA content in tomato leaves, stems, and fruits [93]. Studies on the functions of GalDH and GLDH in AsA synthesis are limited. The only research shows that GLDH has no significant effect on AsA accumulation in tomato fruits [17]. L-gulono-γ-lactone oxidase (GLOase) is the homologous gene of GLDH in animals. A 1.5-fold increase in AsA levels is found in tomato fruit heterologously expressing rat GLOase [94]. In the myo-inositol pathway, MIOX is involved in the production of D-glucuronate. Heterologous expression of AtMIOX increases the AsA content of tomato fruit by 1.4-fold [95]. Munir et al. identifies five MIOX (MIOX1~MIOX5) in tomatoes, and overexpression of MIOX4 significantly increases the AsA content in leaves and red fruits [96]. GalUR is downstream of MIOX, which catalyzes the conversion of D-glucuronate to L-gulonate. Overexpression of GalUR from strawberries results in a 1.2- to 2.5-fold increase in AsA levels of tomato fruits [97]. Pepper fruits are rich in ascorbic acid, and the ascorbic acid content increases with fruit ripening [98]. Although the L-galactose pathway is the main AsA synthesis pathway in pepper fruits and leaves, the transcript levels of genes encoding key enzymes in the AsA biosynthesis pathway were not positively correlated with AsA concentrations in pepper pericarp [99].
In addition to changing the expression of genes related to the AsA synthesis pathway by targeting, the AsA content in fruits can also be regulated by changing the genes associated with AsA degradation and regeneration. For example, AsA content is increased 1.4- to 2.2-fold in the fruits of APX-downregulated tomato lines [100]. Downregulation of AO gene expression in tomatoes can also increase AsA content in fruit [101]. There are two SlDHARs (SlDHAR1 and SlDHAR2) in tomato. Overexpression of SlDHAR1 significantly increases the AsA content in tomato leaves and red ripe fruits. In contrast, overexpression of SlDHAR2 only increases the AsA content in leaves, indicating that the regeneration of AsA in fruits is mainly regulated by SlDHAR1 [102]. The upregulated [101] SlMDHAR expression leads to a 0.7-fold decrease in AsA content in tomato fruits but has no effect on AsA content in tomato leaves [56]. These results suggest that SlDHAR1 and SlMDHAR play regulatory functions in different processes of tomato fruit ripening but are not the main limiting factor in leaves. During citrus fruit ripening, the changes in ascorbic acid content were not completely consistent with the expression of ascorbic acid metabolism-related genes, but were regulated by complex and specific interactions of synthesis and recycling-related genes [103,104].

4. Regulatory Genes That Control Ascorbic Acid Accumulation

Aside from genetic engineering to control the expression of genes related to AsA metabolism, plants have a variety of transcriptional and translational regulatory mechanisms. The transcription factors can bind to gene promoters and control the expression of target genes at the transcriptional level. There have been some preliminary studies on the regulation of AsA accumulation in fruits by transcription factors (Table 3). SlICE1, a bHLH transcription factor, is closely related to chilling resistance in tomatoes. Overexpression of SlICE1 in tomatoes increases the AsA content in the ripe red fruit of tomato and improves the resistance of tomato fruit to low-temperature stress [105]. In tomato L1L4 transcription factor mutants, AsA content is significantly increased, suggesting that L1L4 may negatively influence AsA synthesis [106]. Heterologous expression of Arabidopsis brassinosteroid response transcription factor Brassinazole resistant 1 (BZR1-1D) in tomato can significantly increase AsA content in fruit [107]. It is found that MADS-box and CCAAT motifs are enriched in the promoters of tomato AsA synthesis-related genes, indicating that a single transcription factor may regulate multiple AsA synthesis genes [108]. Overexpression of the banana MaMADS7 transcription factor in tomatoes significantly increases the AsA content in the fruit [109]. The CCAAT-box transcription factor SlNFYA10 in tomatoes can bind to the promoter of SlGME1 and negatively regulate the expression of SlGME1 and the level of AsA in leaves and fruits [110]. An HD-Zip I family transcription factor SIHZ24 is identified in tomato, which binds to the promoter of SlGMP3, activates the expression of SlGMP3, and then positively regulates the accumulation of AsA [111]. In addition, SlHZ24 can bind to the promoters of GME2 and GGP and regulate the accumulation of AsA by multiple targets [111]. MdERF98 activates the expression of MdGMP1, thereby promoting the synthesis of AsA in apple [36]. Transient expression of AcERF91 in kiwifruit can increase AsA content. At the same time, AcERF91 can bind to the promoter of AcGGP3 and activate the transcription of AcGGP3 [112]. Pattern regulation of the AceMYBS1 gene indicated that AceMYBS1 positively regulates AsA accumulation in kiwifruit [92]. In addition, AceMYBS1 can also bind to the AceGGP3 promoter and activate its expression, indicating that multiple transcription factors may simultaneously regulate the expression of a single AsA synthesis-related gene [92]. In addition, some transcription factors that regulate AsA degradation and regeneration are also identified. MdMYB1 can bind to the promoter of MdDHAR, activate the expression of DHAR, and then increase the content of AsA in apples [113]. MsSCL26.1 can bind to the P1 region of the MsMDHAR promoter and activate the transcription of MsMDHAR, thereby inhibiting the synthesis of AsA in the apple [114]. After transient expression of MYB, NAC, and ZIF transcription factors in tomato, AsA metabolism-related genes, such as GMP2, GalUR, AO2, and APX6, are significantly upregulated [115]. Transcript levels of CaMYB16 gene and GLDH in pepper fruit are highly correlated, indicating that CaMYB16 may be involved in ascorbic acid synthesis in pepper fruit [116].
The research on the regulatory mechanism of AsA accumulation at the protein level is still minimal (Table 3). AcESE3 and AcMYBR can interact with AcGGP3 to regulate AsA synthesis in kiwifruit [91]. As a result of MdAMR1L1 overexpression or silencing in apples, a negative relationship between Asc levels and MdAMR1L1 is detected [36]. MdAMR1L1 protein stimulated MdGMP1 degradation through ubiquitination, inhibiting AsA biosynthesis at a post-translational level [36].
In addition to the genes that can encode proteins, there are many noncoding RNAs in the plant genome. In plant development, the function of small ncRNAs, such as microRNAs and small interfering RNAs, has been extensively studied in the past decade. In total, 118 differentially expressed lncRNAs (DE-lncRNAs) and 32 differentially expressed microRNAs are identified during seabuckthorn fruit development [117]. These DE-lncRNAs are particularly enriched in the biosynthesis of AsA, carotenoids, and flavonoids. A miRNA, mdm-miR171i, is identified in the apple that explicitly targets and degrades MsSCL26.1, thereby enhancing AsA accumulation (Table 3) [114].

5. Effects of Hormones on the Ascorbic Acid Accumulation in Fruits

Classical plant hormones mainly include auxin, cytokinin, gibberellin (GA), ethylene, abscisic acid (ABA), salicylic acid (SA), jasmonic acid (JA), and brassinolide, which play an important regulatory role in fruit ripening and nutritional quality. Many studies have used phytohormones to regulate the accumulation of AsA in fruits. Application of GA3 to plum fruit delays the decrease in AsA concentration and reduces flesh browning development during storage at low temperature [118]. The combined treatments of GA3 and phenylurea significantly delay the losses in AsA contents and suppress fruit softening of harvested banana fruit [119]. GA3 (50 ppm) + 1-methyl-cyclopropane-treated strawberry fruits increase the retention of vitamin C over their shelf life compared to the control group [120]. Exogenous gibberellin and ethylene treatments can also significantly increase ascorbic acid content in citrus fruits [121]. According to a recent study, 6-benzylaminopurine and kinetin increase the AsA content of strawberry fruits by 33.96% and 27.22%, respectively, compared to the control [122].
The plant growth regulator ABA plays a role in fruit ripening. Abscisic acid and ethylene regulate AsA synthesis through antagonism [123]. It has been found that ABA alters the AsA redox state at the early stages of fruit development and more than doubles AsA levels at the end of fruit ripening in red raspberry (Rubus idaeus L.) [124]. Strawberry had a 1.6-fold increase in ascorbic acid content after treatment with 1 μM abscisic acid [125]. La(NO3)3 treatment can induce ABA synthesis in strawberry fruit [126]. At the same time, La(NO3)3 increases the activities of DHAR, MDHAR, and GalLDH and decreases the activities of APX and AOO, resulting in increased AsA content in strawberry fruit [126]. Tungstate is an ABA synthesis inhibitor. Tungstate treatment can activate the expression of GR, MDHAR, and GalLDH, while inhibiting the expression of DHAR and APX, thereby inhibiting the accumulation of AsA in strawberry fruit [126].
SA is a vital phytohormone involved in regulating plant resistance to various biotic and abiotic stresses. Treatment with 2 mM SA can maintain the level of AsA and reduce the chilling injury of pomegranate fruit during low-temperature storage [127]. SA treatment also significantly delays the reduction in AsA content in plum, cornelian cherry, pineapple, pear, and strawberry fruits during low-temperature storage and improves the resistance of fruits to low-temperature stress [128,129,130,131,132]. Under the combined treatment of 1.0 mM SA and 2% chitosan, the content of AsA in lychee fruit is maintained at a high level [133]. In addition, using chitosan combined with SA treatment can effectively increase the AsA content of grapefruit fruit and activate the disease resistance to green mold [134].
In addition to SA, JA is also believed to play a variety of essential roles in regulating stress responses and plant growth [135]. MeJA treatment can significantly increase the AsA content in star fruit, blueberry, and pineapple fruits and delay the quality decline of fruits under low-temperature storage [136,137,138]. Postharvest cherry tomato fruits treated with MeJA contain significantly higher AsA and carotenoids, especially lycopene [139]. MeJA treatment could inhibit the activity of AO and increase the activity of DHAR, thereby increasing the content of AsA in loquat fruit and delaying the occurrence of internal browning caused by chilling injury [140]. MYC2 acts as a regulatory center of JA signaling and is involved in cold resistance in many horticultural crops [141]. MYC2 participates in the AsA-GSH cycle and regulates the cold tolerance of tomato fruit by regulating the accumulation of AsA [142]. Treatment with 0.25 mM MeJA can significantly increase ‘Kumato’ tomato fruit yield and AsA content [143]. The treatment of pomegranate and blueberry fruits with MeJA can also increase the AsA content and total antioxidant activity in the fruits and improve the fruits’ antioxidant capacity and storage stability [144,145]. In addition, MeJA pretreatment effectively prevents wound-induced loss of AsA and organic acids and the deterioration of the flesh color of freshly cut pitaya fruits [146].
Melatonin (MT), a hormone found in the pineal gland, has also been found in plants. As a new research-hot phytohormone, melatonin plays a vital role in scavenging reactive oxygen species and improving the resistance of plants to environmental stress [147]. Little research has been conducted on the relationship between melatonin and AsA. A 100 μM melatonin treatment increases the levels of AsA and delays senescence in sweet cherries [148]. After treatment of pomegranate fruit with 100 μM melatonin, the activities of APX and GR are increased and the activity of AAO is decreased, resulting in a higher accumulation of AA and GSH and improved resistance of the fruit to cold [149]. As well as being essential antioxidants, there may also be a deeper relationship between melatonin and AsA.

6. Regulation of Ascorbic Acid Accumulation by Environmental Factors

Environmental factors, such as temperature, light, and water, profoundly impact fruit development and nutritional quality. The accumulation of AsA in the fruit helps to improve the nutritional value of the fruit and improve the resistance of the fruit to various environmental stresses. Studying the effect of environmental factors on the accumulation of AsA in fruits has essential theoretical and practical significance for cultivating high-quality fruits rich in AsA (Figure 3). Light and AsA accumulation in fruit are closely related. After shading treatment of tomatoes, fruit ripening is delayed and AsA content is also significantly decreased [150,151]. Similarly, light affects the AsA content of citrus fruits [121,152]. The research on the mechanism of light-induced AsA synthesis is still very preliminary. The D-mannose/L-galactose pathway produces a majority of AsA in plants. However, although light induces tomato fruit ripening and AsA accumulation, the carbohydrate levels in the fruit did not change significantly. Changes in carbohydrate content in fruit also did not affect the light-induced AsA synthesis, indicating that light-induced AsA synthesis in tomato fruit is independent of carbohydrates in vivo [153]. In addition, light induces the expression of critical genes in the D-mannose/L-galactose pathway, while inhibiting the expression of genes related to AsA degradation, thereby enhancing the synthesis of AsA in tomato fruit [154].
Long-term low-temperature storage can damage the quality of the fruit. With the increase in low-temperature storage time, low-temperature damage occurs in mango fruit, and the content of antioxidants, such as AsA, increased [155]. However, a recent study showed that a 1-min heat treatment at 55 °C can significantly reduce the damage of red bell pepper fruits during low-temperature storage. Preheating can substantially increase the content of AsA and glutathione during refrigeration [156]. Mild drought and salt stress can also increase the content of anthocyanin and AsA in strawberry fruit without affecting the yield [157]. These studies show that moderate environmental stress stimulation has a good application prospect in improving fruit quality.

7. Conclusions

AsA plays an important role in regulating fruit quality and stress resistance. Although most genes related to ascorbic acid synthesis and metabolism have been found in model plants, such as Arabidopsis, studies on fruits are very limited. The discovery of the functions of ascorbic acid anabolism-related genes in fruits has important guiding significance for the use of genetic engineering technology to change the AsA content in fruits. At the same time, the regulatory mechanism of AsA accumulation in fruits is still lacking, and more regulators at the transcriptional and post-translational levels need to be discovered. Plant hormones, such as gibberellin, ethylene, salicylic acid, jasmonic acid, and melatonin, can all affect the accumulation of AsA in fruits, but the regulatory mechanism is still unclear and needs further study. In addition, light and moderate environmental stress are beneficial to the accumulation of ascorbic acid in fruit. Using different environmental factors to increase the content of AsA in fruit also has important application value for fruit quality improvement.

Author Contributions

X.Z., M.P. and W.D. conceived the review’s outline. M.G. and Q.Z. reviewed the metabolic pathway of ascorbic acid. H.T. and L.L. reviewed regulation of gene transcription and protein levels of ascorbic acid metabolism-related genes in fruit. X.Z., Y.T. and Z.L. reviewed regulation of hormones and environmental factors on the ascorbic acid accumulation in fruits. X.Z. wrote and edited the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Tianfu Scholar Program of Sichuan Province (Wei Deng, Department of Human Resources and Social Security of Sichuan Province 2021-58) and the Technology Innovation and Application Development Project in Chongqing (Wei Deng, cstc2021jscx-cylhX0115).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the Tianfu Scholar Program of Sichuan Province (Department of Human Resources and Social Security of Sichuan Province 2021-58) and the Technology Innovation and Application Development Project in Chongqing (cstc2021jscx-cylhX0115). The authors thank Mingchao Peng and Wei Deng for their contributions to the preparation and publication of the paper. Compliance with ethical standards.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Macknight, R.C.; Laing, W.A.; Bulley, S.M.; Broad, R.C.; Johnson, A.A.T.; 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] [PubMed]
  2. Lykkesfeldt, J.; Tveden-Nyborg, P. The pharmacokinetics of vitamin C. Nutrients 2019, 11, 2412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Smirnoff, N. Ascorbic acid metabolism and functions: A comparison of plants and mammals. Free Radic. Biol. Med. 2018, 122, 116–129. [Google Scholar] [CrossRef] [PubMed]
  4. Valpuesta, V.; Botella, M.A. Biosynthesis of l-ascorbic acid in plants: New pathways for an old antioxidant. Trends Plant Sci. 2004, 9, 573–577. [Google Scholar] [CrossRef]
  5. Chung, I.M.; Venkidasamy, B.; Upadhyaya, C.P.; Packiaraj, G.; Rajakumar, G.; Thiruvengadam, M. Alleviation of phytophthora infestans mediated necrotic stress in the transgenic potato (Solanum tuberosum L.) with enhanced ascorbic acid accumulation. Plants 2019, 8, 365. [Google Scholar] [CrossRef] [Green Version]
  6. Singh, P.; Singh, J.; Ray, S.; Rajput, R.S.; Vaishnav, A.; Singh, R.K.; Singh, H.B. Seed biopriming with antagonistic microbes and ascorbic acid induce resistance in tomato against Fusarium wilt. Microbiol. Res. 2020, 237, 126482. [Google Scholar] [CrossRef]
  7. Akram, N.A.; Shafiq, F.; Ashraf, M. Ascorbic acid-a potential oxidant scavenger and its role in plant development and abiotic stress tolerance. Front. Plant Sci. 2017, 8, 613. [Google Scholar] [CrossRef]
  8. Mellidou, I.; Koukounaras, A.; Kostas, S.; Patelou, E.; Kanellis, A.K. Regulation of vitamin c accumulation for improved tomato fruit quality and alleviation of abiotic stress. Genes 2021, 12, 694. [Google Scholar] [CrossRef]
  9. Li, Y.; Chu, Z.N.; Luo, J.Y.; Zhou, Y.H.; Cai, Y.J.; Lu, Y.G.; Xia, J.H.; Kuang, H.H.; Ye, Z.B.; Ouyang, B. The C2H2 zinc-finger protein SlZF3 regulates asa synthesis and salt tolerance by interacting with CSN5B. Plant Biotechnol. J. 2018, 16, 1201–1213. [Google Scholar] [CrossRef] [Green Version]
  10. Batth, R.; Singh, K.; Kumari, S.; Mustafiz, A. Transcript profiling reveals the presence of abiotic stress and developmental stage specific ascorbate oxidase genes in plants. Front. Plant Sci. 2017, 8, 198. [Google Scholar] [CrossRef]
  11. Min, K.; Chen, K.; Arora, R. A metabolomics study of ascorbic acid-induced in situ freezing tolerance in spinach (Spinacia oleracea L.). Plant Direct 2020, 4, e00202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Ahmad, I.; Basra, S.M.A.; Wahid, A. Exogenous application of ascorbic acid, salicylic acid and hydrogen peroxide improves the productivity of hybrid maize at low temperature stress. Int. J. Agric. Biol. 2014, 16, 825–830. [Google Scholar]
  13. Kamal, M.A.; Saleem, M.F.; Shahid, M.; Awais, M.; Khan, H.Z.; Ahmed, K. Ascorbic acid triggered physiochemical transformations at different phenological stages of heat-stressed bt cotton. J. Agron. Crop. Sci. 2017, 203, 323–331. [Google Scholar] [CrossRef]
  14. Lo’ay, A.A.; EL-Khateeb, A.Y. Antioxidant enzyme activities and exogenous ascorbic acid treatment of ‘williams’ banana during long-term cold storage stress. Sci. Hortic. 2018, 234, 210–219. [Google Scholar] [CrossRef]
  15. Saleem, M.S.; Anjum, M.A.; Naz, S.; Ali, S.; Hussain, S.; Azam, M.; Sardar, H.; Khaliq, G.; Canan, I.; Ejaz, S. Incorporation of ascorbic acid in chitosan-based edible coating improves postharvest quality and storability of strawberry fruits. Int. J. Biol. Macromol. 2021, 189, 160–169. [Google Scholar] [CrossRef] [PubMed]
  16. Steelheart, C.; Alegre, M.L.; Baldet, P.; Rothan, C.; Bres, C.; Just, D.; Okabe, Y.; Ezura, H.; Ganganelli, I.; Grozeff, G.E.G.; et al. The effect of low ascorbic acid content on tomato fruit ripening. Planta 2020, 252, 36. [Google Scholar] [CrossRef]
  17. Alhagdow, M.; Mounet, F.; Gilbert, L.; Nunes-Nesi, A.; Garcia, V.; Just, D.; Petit, J.; Beauvoit, B.; Fernie, A.R.; Rothan, C.; et al. Silencing of the mitochondrial ascorbate synthesizing enzyme L-galactono-1,4-lactone dehydrogenase affects plant and fruit development in tomato. Plant Physiol. 2007, 145, 1408–1422. [Google Scholar] [CrossRef] [Green Version]
  18. 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] [Green Version]
  19. Opara, U.L.; Al-Ani, M.R.; Al-Rahbi, N.M. Effect of fruit ripening stage on physico-chemical properties, nutritional composition and antioxidant components of tomato (lycopersicum esculentum) cultivars. Food Bioprocess. Technol. 2012, 5, 3236–3243. [Google Scholar] [CrossRef]
  20. Ilahy, R.; Hdider, C.; Lenucci, M.S.; Tlili, I.; Dalessandro, G. Antioxidant activity and bioactive compound changes during fruit ripening of high-lycopene tomato cultivars. J. Food Compos. Anal. 2011, 24, 588–595. [Google Scholar] [CrossRef]
  21. Nishiyama, I.; Yamashita, Y.; Yamanaka, M.; Shimohashi, A.; Fukuda, T.; Oota, T. Varietal difference in vitamin c content in the fruit of kiwifruit and other actinidia species. J. Agr. Food Chem. 2004, 52, 5472–5475. [Google Scholar] [CrossRef] [PubMed]
  22. Boonyakiat, D.; Chuamuangphan, C.; Maniwara, P.; Seehanam, P. Comparison of physico-chemical quality of different strawberry cultivars at three maturity stages. Int. Food Res. J. 2016, 23, 2405–2412. [Google Scholar]
  23. Kim, S.K.; Kim, D.S.; Kim, D.Y.; Chun, C. Variation of bioactive compounds content of 14 oriental strawberry cultivars. Food Chem. 2015, 184, 196–202. [Google Scholar] [CrossRef] [PubMed]
  24. Abaci, Z.T.; Sevindik, E.; Ayvaz, M. Comparative study of bioactive components in pear genotypes from ardahan/turkey. Biotechnol. Biotec. Eq. 2016, 30, 36–43. [Google Scholar] [CrossRef] [Green Version]
  25. Dhuique-Mayer, C.; Caris-Veyrat, C.; Ollitrault, P.; Curk, F.; Amiot, M.J. Varietal and interspecific influence on micronutrient contents in citrus from the mediterranean area. J. Agric. Food Chem. 2005, 53, 2140–2145. [Google Scholar] [CrossRef]
  26. Tlili, I.; Hdider, C.; Lenucci, M.S.; Ilahy, R.; Jebari, H.; Dalessandro, G. Bioactive compounds and antioxidant activities during fruit ripening of watermelon cultivars. J. Food Compos. Anal. 2011, 24, 923–928. [Google Scholar] [CrossRef]
  27. Di Matteo, A.; Simeone, G.D.; Cirillo, A.; Rao, M.A.; Di Vaio, C. Morphological characteristics, ascorbic acid and antioxidant activity during fruit ripening of four lemon (Citrus limon (L.) Burm. F.) cultivars. Sci. Hortic. 2021, 276, 109741. [Google Scholar] [CrossRef]
  28. Li, Y.J.; Sun, H.X.; Li, J.D.; Qin, S.; Yang, W.; Ma, X.Y.; Qiao, X.W.; Yang, B.R. Effects of genetic background and altitude on sugars, malic acid and ascorbic acid in fruits of wild and cultivated apples (Malus sp.). Foods 2021, 10, 2950. [Google Scholar] [CrossRef]
  29. Langova, R.; Juzl, M.; Cwikova, O.; Kos, I. Effect of different method of drying of five varieties grapes (vitis vinifera L.) on the bunch stem on physicochemical, microbiological, and sensory quality. Foods 2020, 9, 1183. [Google Scholar] [CrossRef]
  30. Setiasih, I.S.; Sukarminah, E. Effect of type and maturity on water content of three varieties of hot chilli (Capsicum frustencent L, catas, segana end domba Variety). IOP Conf. Ser. Mater. Sci. Eng. 2019, 506, 012041. [Google Scholar]
  31. 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] [PubMed]
  32. 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] [PubMed] [Green Version]
  33. Qian, W.Q.; Yu, C.M.; Qin, H.J.; Liu, X.; Zhang, A.M.; Johansen, I.E.; Wang, D.W. Molecular and functional analysis of phosphomannomutase (PMM) from higher plants and genetic evidence for the involvement of pmm in ascorbic acid biosynthesis in Arabidopsis and Nicotiana benthamiana. Plant J. 2007, 49, 399–413. [Google Scholar] [CrossRef] [PubMed]
  34. 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] [PubMed]
  35. Li, X.J.; Ye, J.; Munir, S.; Yang, T.; Chen, W.F.; Liu, G.Z.; Zheng, W.; Zhang, Y.Y. Biosynthetic gene pyramiding leads to ascorbate accumulation with enhanced oxidative stress tolerance in tomato. Int. J. Mol. Sci. 2019, 20, 1558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ma, S.Y.; Li, H.X.; Wang, L.; Li, B.Y.; Wang, Z.Y.; Ma, B.Q.; Ma, F.W.; Li, M.J. F-box protein mdamr1l1 regulates ascorbate biosynthesis in apple by modulating gdp-mannose pyrophosphorylase. Plant Physiol. 2022, 188, 653–669. [Google Scholar] [CrossRef] [PubMed]
  37. Cascia, G.; Bulley, S.M.; Punter, M.; Bowen, J.; Rassam, M.; Schotsmans, W.C.; Larrigaudiere, C.; Johnston, J.W. Investigation of ascorbate metabolism during inducement of storage disorders in pear. Physiol. Plant. 2013, 147, 121–134. [Google Scholar] [CrossRef]
  38. Gilbert, L.; Alhagdow, M.; Nunes-Nesi, A.; Quemener, B.; Guillon, F.; Bouchet, B.; Faurobert, M.; Gouble, B.; Page, D.; Garcia, V.; et al. GDP-d-mannose 3,5-epimerase (GME) plays a key role at the intersection of ascorbate and non-cellulosic cell-wall biosynthesis in tomato. Plant J. 2009, 60, 499–508. [Google Scholar] [CrossRef]
  39. Laing, W.A.; Wright, M.A.; Cooney, J.; Bulley, S.M. The missing step of the L-galactose pathway of ascorbate biosynthesis in plants, an l-galactose guanyltransferase, increases leaf ascorbate content. Proc. Natl. Acad. Sci. USA 2007, 104, 9534–9539. [Google Scholar] [CrossRef] [Green Version]
  40. 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]
  41. Laing, W.A.; Bulley, S.; Wright, M.; Cooney, J.; Jensen, D.; Barraclough, D.; MacRae, E. A highly specific L-galactose-1-phosphate phosphatase on the path to ascorbate biosynthesis. Proc. Natl. Acad. Sci. USA 2004, 101, 16976–16981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Conklin, P.L.; Gatzek, S.; Wheeler, G.L.; Dowdle, J.; Raymond, M.J.; Rolinski, S.; Isupov, M.; Littlechild, J.A.; Smirnoff, N. Arabidopsis thaliana vtc4 encodes l-galactose-1-p phosphatase, a plant ascorbic acid biosynthetic enzyme. J. Biol. Chem. 2006, 281, 15662–15670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Torabinejad, J.; Donahue, J.L.; Gunesekera, B.N.; Allen-Daniels, M.J.; Gillaspy, G.E. VTC4 is a bifunctional enzyme that affects myoinositol and ascorbate biosynthesis in plants. Plant Physiol. 2009, 150, 951–961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Pineau, B.; Layoune, O.; Danon, A.; De Paepe, R. L-galactono-1,4-lactone dehydrogenase is required for the accumulation of plant respiratory complex i. J. Biol. Chem. 2008, 283, 32500–32505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Siendones, E.; Gonzalez-Reyes, J.A.; Santos-Ocana, C.; Navas, P.; Cordoba, F. Biosynthesis of ascorbic acid in kidney bean. L-galactono-gamma-lactone dehydrogenase is an intrinsic protein located at the mitochondrial inner membrane. Plant Physiol. 1999, 120, 907–912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. 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] [PubMed] [Green Version]
  47. Wolucka, B.A.; Van Montagu, M. GDP-mannose 3′,5′-epimerase forms GDP-L-gulose, a putative intermediate for the de novo biosynthesis of vitamin c in plants. J. Biol. Chem. 2003, 278, 47483–47490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. 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]
  49. Maruta, T.; Ichikawa, Y.; Mieda, T.; Takeda, T.; Shigeoka, S. The contribution of arabidopsis homologs of L-gulono-1,4-lactone oxidase to the biosynthesis of ascorbic acid. Biosci. Biotechnol. Biochem. 2010, 74, 1494–1497. [Google Scholar] [CrossRef] [Green Version]
  50. Aboobucker, S.I.; Suza, W.P.; Lorence, A. Characterization of two arabidopsis L-gulono-1,4-lactone oxidases, AtGulLO3 and AtGulLO5, involved in ascorbate biosynthesis. React. Oxyg. Species 2017, 4, 389–417. [Google Scholar] [CrossRef]
  51. 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] [PubMed] [Green Version]
  52. Siddique, S.; Endres, S.; Atkins, J.M.; Szakasits, D.; Wieczorek, K.; Hofmann, J.; Blaukopf, C.; Urwin, P.E.; Tenhaken, R.; Grundler, F.M.W.; et al. Myo-inositol oxygenase genes are involved in the development of syncytia induced by heterodera schachtii in arabidopsis roots. New Phytol. 2009, 184, 457–472. [Google Scholar] [CrossRef] [PubMed]
  53. 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] [PubMed]
  54. Badejo, A.A.; Wada, K.; Gao, Y.S.; 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] [PubMed] [Green Version]
  55. Foyer, C.H.; Halliwell, B. Presence of glutathione and glutathione reductase in chloroplasts—Proposed role in ascorbic-acid metabolism. Planta 1976, 133, 21–25. [Google Scholar] [CrossRef] [PubMed]
  56. Haroldsen, V.M.; Chi-Ham, C.L.; Kulkarni, S.; Lorence, A.; Bennett, A.B. Constitutively expressed dhar and mdhar influence fruit, but not foliar ascorbate levels in tomato. Plant Physiol. Bioch. 2011, 49, 1244–1249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Qin, A.G.; Shi, Q.H.; Yu, X.C. Ascorbic acid contents in transgenic potato plants overexpressing two dehydroascorbate reductase genes. Mol. Biol. Rep. 2011, 38, 1557–1566. [Google Scholar] [CrossRef]
  58. Gest, N.; Garchery, C.; Gautier, H.; Jimenez, A.; Stevens, R. Light-dependent regulation of ascorbate in tomato by a monodehydroascorbate reductase localized in peroxisomes and the cytosol. Plant Biotechnol. J. 2013, 11, 344–354. [Google Scholar] [CrossRef]
  59. Ren, J.; Duan, W.K.; Chen, Z.W.; Zhang, S.; Song, X.M.; Liu, T.K.; Hou, X.L.; Li, Y. Overexpression of the monodehydroascorbate reductase gene from non-heading chinese cabbage reduces ascorbate level and growth in transgenic tobacco. Plant Mol. Biol. Rep. 2015, 33, 881–892. [Google Scholar] [CrossRef]
  60. Eltelib, H.A.; Fujikawa, Y.; Esaka, M. Overexpression of the acerola (malpighia glabra) monodehydroascorbate reductase gene in transgenic tobacco plants results in increased ascorbate levels and enhanced tolerance to salt stress. S. Afr. J. Bot. 2012, 78, 295–301. [Google Scholar] [CrossRef] [Green Version]
  61. DeBolt, S.; Hardie, J.; Tyerman, S.; Ford, C.M. Composition and synthesis of raphide crystals and druse crystals in berries of vitis vinifera l. Cv. Cabernet sauvignon: Ascorbic acid as precursor for both oxalic and tartaric acids as revealed by radiolabelling studies. Aust. J. Grape Wine R. 2004, 10, 134–142. [Google Scholar] [CrossRef]
  62. Yang, J.C.; Loewus, F.A. Metabolic conversion of L-ascorbic acid to oxalic acid in oxalate-accumulating plants. Plant Physiol. 1975, 56, 283–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Burbidge, C.A.; Ford, C.M.; Melino, V.J.; Wong, D.C.J.; Jia, Y.; Jenkins, C.L.D.; Soole, K.L.; Castellarin, S.D.; Darriet, P.; Rienth, M.; et al. Biosynthesis and cellular functions of tartaric acid in grapevines. Front. Plant Sci. 2021, 12, 309. [Google Scholar] [CrossRef] [PubMed]
  64. Green, M.A.; Fry, S.C. Vitamin c degradation in plant cells via enzymatic hydrolysis of 4-o-oxalyl-l-threonate. Nature 2005, 433, 83–87. [Google Scholar] [CrossRef] [PubMed]
  65. 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] [PubMed] [Green Version]
  66. Sharova, E.I.; Medvedev, S.S.; Demidchik, V.V. Ascorbate in the apoplast: Metabolism and functions. Russ. J. Plant Physiol. 2020, 67, 207–220. [Google Scholar] [CrossRef]
  67. Smirnoff, N. Ascorbic acid: Metabolism and functions of a multi-facetted molecule. Curr. Opin. Plant Biol. 2000, 3, 229–235. [Google Scholar] [CrossRef]
  68. Pradedova, E.V.; Nimaeva, O.D.; Semenova, N.V.; Salyaev, R.K. Comparative study on redox state of ascorbic acid and ascorbate oxidase activity in vacuoles and leucoplasts of red beet taproots during physiological dormancy. Russ. J. Plant Physiol. 2021, 68, 74–84. [Google Scholar] [CrossRef]
  69. 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]
  70. Horemans, N.; Foyer, C.H.; Asard, H. Transport and action of ascorbate at the plant plasma membrane. Trends Plant Sci. 2000, 5, 263–267. [Google Scholar] [CrossRef]
  71. Scalera, V.; Giangregorio, N.; De Leonardis, S.; Console, L.; Carulli, E.S.; Tonazzi, A. Characterization of a novel mitochondrial ascorbate transporter from rat liver and potato mitochondria. Front. Mol. Biosci. 2018, 5, 58. [Google Scholar] [CrossRef] [PubMed]
  72. 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] [Green Version]
  73. Sano, S. Molecular and functional characterization of monodehydro-ascorbate and dehydroascorbate reductases. In Ascorbic Acid in Plant Growth, Development and Stress Tolerance; Hossain, M.A., Munné-Bosch, S., Burritt, D.J., Diaz-Vivancos, P., Fujita, M., Lorence, A., Eds.; Springer International Publishing: Cham, Switzerland, 2017; pp. 129–156. [Google Scholar]
  74. Noshi, M.; Yamada, H.; Hatanaka, R.; Tanabe, N.; Tamoi, M.; Shigeoka, S. Arabidopsis dehydroascorbate reductase 1 and 2 modulate redox states of ascorbate-glutathione cycle in the cytosol in response to photooxidative stress. Biosci. Biotechnol. Biochem. 2017, 81, 523–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Rautenkranz, A.A.F.; Li, L.J.; Machler, F.; Martinoia, E.; Oertli, J.J. Transport of ascorbic and dehydroascorbic acids across protoplast and vacuole membranes isolated from barley (hordeum-vulgare l cv gerbel) leaves. Plant Physiol. 1994, 106, 187–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Harrison, F.E.; Dawes, S.M.; Meredith, M.E.; Babaev, V.R.; Li, L.; May, J.M. Low vitamin c and increased oxidative stress and cell death in mice that lack the sodium-dependent vitamin c transporter svct2. Free Radic. Biol. Med. 2010, 49, 821–829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Michels, A.J.; Hagen, T.M.; Frei, B. Human genetic variation influences vitamin c homeostasis by altering vitamin c transport and antioxidant enzyme function. Annu. Rev. Nutr. 2013, 33, 45–70. [Google Scholar] [CrossRef] [Green Version]
  78. Maurino, V.G.; Grube, E.; Zielinski, J.; Schild, A.; Fischer, K.; Flugge, U.I. Identification and expression analysis of twelve members of the nucleobase-ascorbate transporter (nat) gene family in arabidopsis thaliana. Plant Cell Physiol. 2006, 47, 1381–1393. [Google Scholar] [CrossRef] [Green Version]
  79. Preger, V.; Scagliarini, S.; Pupillo, P.; Trost, P. Identification of an ascorbate-dependent cytochrome b of the tonoplast membrane sharing biochemical features with members of the cytochrome b561 family. Planta 2005, 220, 365–375. [Google Scholar] [CrossRef]
  80. Hancock, R.D.; McRae, D.; Haupt, S.; Viola, R. Synthesis of L-ascorbic acid in the phloem. BMC Plant Biol. 2003, 3, 7. [Google Scholar] [CrossRef] [Green Version]
  81. Franceschi, V.R.; Tarlyn, N.M. L-ascorbic acid is accumulated in source leaf phloem and transported to sink tissues in plants. Plant Physiol. 2002, 130, 649–656. [Google Scholar] [CrossRef] [Green Version]
  82. Tedone, L.; Hancock, R.D.; Alberino, S.; Haupt, S.; Viola, R. Long-distance transport of L-ascorbic acid in potato. BMC Plant Biol. 2004, 4, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Yang, X.Y.; Xie, J.X.; Wang, F.F.; Zhong, J.; Liu, Y.Z.; Li, G.H.; Peng, S.A. Comparison of ascorbate metabolism in fruits of two citrus species with obvious difference in ascorbate content in pulp. J. Plant Physiol. 2011, 168, 2196–2205. [Google Scholar] [CrossRef] [PubMed]
  84. Caruso, P.; Russo, M.P.; Caruso, M.; Di Guardo, M.; Russo, G.; Fabroni, S.; Timpanaro, N.; Licciardello, C. A transcriptional analysis of the genes involved in the ascorbic acid pathways based on a comparison of the juice and leaves of navel and anthocyanin-rich sweet orange varieties. Plants 2021, 10, 1291. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, C.J.; Ouyang, B.; Yang, C.X.; Zhang, X.H.; Liu, H.; Zhang, Y.Y.; Zhang, J.H.; Li, H.X.; Ye, Z.B. Reducing asa leads to leaf lesion and defence response in knock-down of the asa biosynthetic enzyme GDP-D-mannose pyrophosphorylase gene in tomato plant. PLoS ONE 2013, 8, e61987. [Google Scholar] [CrossRef] [Green Version]
  86. Zhang, C.J.; Liu, J.X.; Zhang, Y.Y.; Cai, X.F.; Gong, P.J.; Zhang, J.H.; Wang, T.T.; Li, H.X.; Ye, Z.B. Overexpression of slgmes leads to ascorbate accumulation with enhanced oxidative stress, cold, and salt tolerance in tomato. Plant Cell Rep. 2011, 30, 389–398. [Google Scholar] [CrossRef]
  87. Mounet-Gilbert, L.; Dumont, M.; Ferrand, C.; Bournonville, C.; Monier, A.; Jorly, J.; Lemaire-Chamley, M.; Mori, K.; Atienza, I.; Hernould, M.; et al. Two tomato GDP-D-mannose epimerase isoforms involved in ascorbate biosynthesis play specific roles in cell wall biosynthesis and development. J. Exp. Bot. 2016, 67, 4767–4777. [Google Scholar] [CrossRef] [Green Version]
  88. Alegre, M.L.; Steelheart, C.; Baldet, P.; Rothan, C.; Just, D.; Okabe, Y.; Ezura, H.; Smirnoff, N.; Grozeff, G.E.G.; Bartoli, C.G. Deficiency of gdp-l-galactose phosphorylase, an enzyme required for ascorbic acid synthesis, reduces tomato fruit yield. Planta 2020, 251, 54. [Google Scholar] [CrossRef]
  89. Li, T.D.; Yang, X.P.; Yu, Y.; Si, X.M.; Zhai, X.W.; Zhang, H.W.; Dong, W.X.; Gao, C.X.; Xu, C. Domestication of wild tomato is accelerated by genome editing. Nat. Biotechnol. 2018, 36, 1160–1163. [Google Scholar] [CrossRef]
  90. Bulley, S.; Wright, M.; Rommens, C.; Yan, H.; Rassam, M.; Lin-Wang, K.; Andre, C.; Brewster, D.; Karunairetnam, S.; Allan, A.C.; et al. Enhancing ascorbate in fruits and tubers through over-expression of the L-galactose pathway gene gdp-l-galactose phosphorylase. Plant Biotechnol. J. 2012, 10, 390–397. [Google Scholar] [CrossRef]
  91. Liu, X.Y.; Xie, X.D.; Zhong, C.H.; Li, D.W. Comparative transcriptome analysis revealed the key genes regulating ascorbic acid synthesis in actinidia. Int. J. Mol. Sci. 2021, 22, 12894. [Google Scholar] [CrossRef]
  92. Liu, X.Y.; Wu, R.M.; Bulley, S.M.; Zhong, C.H.; Li, D.W. Kiwifruit MYBS1-like and GBF3 transcription factors influence L-ascorbic acid biosynthesis by activating transcription of GDP-L-galactose phosphorylase 3. New Phytol. 2022, 234, 1782–1800. [Google Scholar] [CrossRef] [PubMed]
  93. Zheng, X.Z.; Yuan, Y.J.; Huang, B.W.; Hu, X.W.; Tang, Y.W.; Xu, X.; Wu, M.B.; Gong, Z.H.; Luo, Y.Q.; Gong, M.; et al. Control of fruit softening and ascorbic acid accumulation by manipulation of SlIMP3 in tomato. Plant Biotechnol. J. 2022, 20, 1213–1225. [Google Scholar] [CrossRef] [PubMed]
  94. Lim, M.Y.; Pulla, R.K.; Park, J.M.; Harn, C.H.; Jeong, B.R. Over-expression of L-gulono-gamma-lactone oxidase (GLoase) gene leads to ascorbate accumulation with enhanced abiotic stress tolerance in tomato. In Vitro Cell Dev. 2012, 48, 453–461. [Google Scholar] [CrossRef]
  95. Cronje, C.; George, G.M.; Fernie, A.R.; Bekker, J.; Kossmann, J.; Bauer, R. Manipulation of L-ascorbic acid biosynthesis pathways in solanum lycopersicum: Elevated GDP-mannose pyrophosphorylase activity enhances L-ascorbate levels in red fruit. Planta 2012, 235, 553–564. [Google Scholar] [CrossRef] [PubMed]
  96. Munir, S.; Mumtaz, M.A.; Ahiakpa, J.K.; Liu, G.Z.; Chen, W.F.; Zhou, G.L.; Zheng, W.; Ye, Z.B.; Zhang, Y.Y. Genome-wide analysis of myo-inositol oxygenase gene family in tomato reveals their involvement in ascorbic acid accumulation. BMC Genom. 2020, 21, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Cai, X.F.; Zhang, C.J.; Ye, J.; Hu, T.X.; Ye, Z.B.; Li, H.X.; Zhang, Y.Y. Ectopic expression of fagalur leads to ascorbate accumulation with enhanced oxidative stress, cold, and salt tolerance in tomato. Plant Growth Regul. 2015, 76, 187–197. [Google Scholar] [CrossRef]
  98. Villa-Rivera, M.G.; Ochoa-Alejo, N. Transcriptional regulation of ripening in chili pepper fruits (Capsicum spp.). Int. J. Mol. Sci. 2021, 22, 12151. [Google Scholar] [CrossRef]
  99. Gomez-Garcia, M.D.; Ochoa-Alejo, N. Predominant role of the l-galactose pathway in L-ascorbic acid biosynthesis in fruits and leaves of the Capsicum annuum L. Chili pepper. Braz. J. Bot. 2016, 39, 157–168. [Google Scholar] [CrossRef]
  100. Zhang, Y.Y.; Li, H.X.; Shu, W.B.; Zhang, C.J.; Ye, Z.B. Rna interference of a mitochondrial apx gene improves vitamin c accumulation in tomato fruit. Sci. Hortic. 2011, 129, 220–226. [Google Scholar] [CrossRef]
  101. Abdelgawad, K.F.; El-Mogy, M.M.; Mohamed, M.I.A.; Garchery, C.; Stevens, R.G. Increasing ascorbic acid content and salinity tolerance of cherry tomato plants by suppressed expression of the ascorbate oxidase genes. Agronomy 2019, 9, 51. [Google Scholar] [CrossRef] [Green Version]
  102. Li, Q.Z.; Li, Y.S.; Li, C.H.; Yu, X.C. Enhanced ascorbic acid accumulation through overexpression of dehydroascorbate reductase confers tolerance to methyl viologen and salt stresses in tomato. Czech J. Genet. Plant 2012, 48, 74–86. [Google Scholar] [CrossRef] [Green Version]
  103. Alos, E.; Rey, F.; Gil, J.V.; Rodrigo, M.J.; Zacarias, L. Ascorbic acid content and transcriptional profiling of genes involved in its metabolism during development of petals, leaves, and fruits of orange (Citrus sinensis cv. Valencia late). Plants 2021, 10, 2590. [Google Scholar] [CrossRef]
  104. Alos, E.; Rodrigo, M.J.; Zacarias, L. Differential transcriptional regulation of ascorbic acid content in peel and pulp of citrus fruits during development and maturation. Planta 2014, 239, 1113–1128. [Google Scholar] [CrossRef] [PubMed]
  105. Miura, K.; Sato, A.; Shiba, H.; Kang, S.W.; Kamada, H.; Ezura, H. Accumulation of antioxidants and antioxidant activity in tomato, solanum lycopersicum, are enhanced by the transcription factor SlICE1. Plant Biotechnol. 2012, 29, 261–269. [Google Scholar] [CrossRef] [Green Version]
  106. Gago, C.; Drosou, V.; Paschalidis, K.; Guerreiro, A.; Miguel, G.; Antunes, D.; Hilioti, Z. Targeted gene disruption coupled with metabolic screen approach to uncover the leafy cotyledon1-like4 (L1L4) function in tomato fruit metabolism. Plant Cell Rep. 2017, 36, 1065–1082. [Google Scholar] [CrossRef] [PubMed]
  107. Liu, L.H.; Jia, C.G.; Zhang, M.; Chen, D.L.; Chen, S.X.; Guo, R.F.; Guo, D.P.; Wang, Q.M. Ectopic expression of a BZR1-1D transcription factor in brassinosteroid signalling enhances carotenoid accumulation and fruit quality attributes in tomato. Plant Biotechnol. J. 2014, 12, 105–115. [Google Scholar] [CrossRef]
  108. Qiu, Z.K.; Li, R.; Zhang, S.B.; Wang, K.T.; Xu, M.; Li, J.Y.; Du, Y.C.; Yu, H.; Cui, X. Identification of regulatory DNA elements using genome-wide mapping of dnase i hypersensitive sites during tomato fruit development. Mol. Plant 2016, 9, 1168–1182. [Google Scholar] [CrossRef] [Green Version]
  109. Liu, J.H.; Liu, L.; Li, Y.J.; Jia, C.H.; Zhang, J.B.; Miao, H.X.; Hu, W.; Wang, Z.; Xu, B.Y.; Jin, Z.Q. Role for the banana agamous-like gene MaMADS7 in regulation of fruit ripening and quality. Physiol. Plantarum 2015, 155, 217–231. [Google Scholar] [CrossRef]
  110. Chen, W.F.; Hu, T.X.; Ye, J.; Wang, B.; Liu, G.Z.; Wang, Y.; Yuan, L.; Li, J.M.; Li, F.M.; Ye, Z.B.; et al. A ccaat-binding factor, SlNFYA10, negatively regulates ascorbate accumulation by modulating the D-mannose/L-galactose pathway in tomato. Hortic. Res. 2020, 7, 1–12. [Google Scholar] [CrossRef]
  111. Hu, T.X.; Ye, J.; Tao, P.W.; Li, H.X.; Zhang, J.H.; Zhang, Y.Y.; Ye, Z.B. 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]
  112. Chen, Y.; Shu, P.; Wang, R.C.; Du, X.F.; Xie, Y.; Du, K.; Deng, H.; Li, M.Z.; Zhang, Y.; Grierson, D.; et al. Ethylene response factor AcERF91 affects ascorbate metabolism via regulation of GDP-galactose phosphorylase encoding gene (AcGGP3) in kiwifruit. Plant Sci. 2021, 313, 111063. [Google Scholar] [CrossRef] [PubMed]
  113. An, J.P.; Wang, X.F.; You, C.X.; Hao, Y.J. The anthocyanin biosynthetic regulator MdMYB1 positively regulates ascorbic acid biosynthesis in apple. Front. Agric. Sci. Eng. 2021, 8, 231–235. [Google Scholar] [CrossRef]
  114. Wang, Y.T.; Feng, C.; Zhai, Z.F.; Peng, X.; Wang, Y.Y.; Sun, Y.T.; Li, J.; Shen, X.S.; Xiao, Y.Q.; Zhu, S.J.; et al. The apple microR171i-SCARECROW-LIKE PROTEINS26.1 module enhances drought stress tolerance by integrating ascorbic acid metabolism. Plant Physiol. 2020, 184, 194–211. [Google Scholar] [CrossRef]
  115. Ye, J.; Hu, T.X.; Yang, C.M.; Li, H.X.; Yang, M.Z.; Ijaz, R.N.; Ye, Z.B.; Zhang, Y.Y. Transcriptome profiling of tomato fruit development reveals transcription factors associated with ascorbic acid, carotenoid and flavonoid biosynthesis. PLoS ONE 2015, 10, e0130885. [Google Scholar] [CrossRef] [PubMed]
  116. Arce-Rodriguez, M.L.; Martinez, O.; Ochoa-Alejo, N. Genome-wide identification and analysis of the myb transcription factor gene family in chili pepper (Capsicum spp.). Int. J. Mol. Sci. 2021, 22, 2229. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, G.Y.; Chen, D.G.; Zhang, T.; Duan, A.G.; Zhang, J.G.; He, C.Y. Transcriptomic and functional analyses unveil the role of long non-coding RNAs in anthocyanin biosynthesis during sea buckthorn fruit ripening. DNA Res. 2018, 25, 465–476. [Google Scholar] [CrossRef] [PubMed]
  118. Li, H.; Jiang, Y.; Li, J. Use of GA(3) to inhibit flesh browning development of plum fruit during storage at low temperature. Eur. J. Hortic. Sci. 2006, 71, 231–235. [Google Scholar]
  119. Huang, H.; Jing, G.X.; Wang, H.; Duan, X.W.; Qu, H.X.; Jiang, Y.M. The combined effects of phenylurea and gibberellins on quality maintenance and shelf life extension of banana fruit during storage. Sci. Hortic. 2014, 167, 36–42. [Google Scholar] [CrossRef]
  120. Tas, A.; Berk, S.K.; Orman, E.; Gundogdu, M.; Ercisli, S.; Karatas, N.; Jurikova, T.; Adamkova, A.; Nedomova, S.; Mlcek, J. Influence of pre-harvest gibberellic acid and post-harvest 1-methyl cyclopropane treatments on phenolic compounds, vitamin c and organic acid contents during the shelf life of strawberry fruits. Plants 2021, 10, 121. [Google Scholar] [CrossRef]
  121. Magwaza, L.S.; Mditshwa, A.; Tesfay, S.Z.; Opara, U.L. An overview of preharvest factors affecting vitamin c content of citrus fruit. Sci. Hortic. 2017, 216, 12–21. [Google Scholar] [CrossRef]
  122. Dong, Y.H.; Song, M.Y.; Liu, X.X.; Tian, R.P.; Zhang, L.Y.; Gan, L.J. Effects of exogenous kt and ba on fruit quality in strawberry (Fragaria vesca). J. Hortic. Sci. Biotechnol. 2022, 97, 236–243. [Google Scholar] [CrossRef]
  123. Yu, Y.W.; Wang, J.; Li, S.H.; Kakan, X.; Zhou, Y.; Miao, Y.; Wang, F.F.; Qin, H.; Huang, R.F. Ascorbic acid integrates the antagonistic modulation of ethylene and abscisic acid in the accumulation of reactive oxygen species. Plant Physiol. 2019, 179, 1861–1875. [Google Scholar] [CrossRef] [Green Version]
  124. Miret, J.A.; Munne-Bosch, S. Abscisic acid and pyrabactin improve vitamin c contents in raspberries. Food Chem. 2016, 203, 216–223. [Google Scholar] [CrossRef] [PubMed]
  125. Li, D.D.; Li, L.; Luo, Z.S.; Mou, W.S.; Mao, L.C.; Ying, T.J. Comparative transcriptome analysis reveals the influence of abscisic acid on the metabolism of pigments, ascorbic acid and folic acid during strawberry fruit ripening. PLoS ONE 2015, 10, e0130037. [Google Scholar] [CrossRef]
  126. Shan, C.J.; Zhang, Y.Y.; Zhang, H.X. Aba participates in the regulation of vitamin c content in the fruit of strawberry using lanthanum nitrate. Sci. Hortic. 2018, 233, 455–459. [Google Scholar] [CrossRef]
  127. Sayyari, M.; Babalar, M.; Kalantari, S.; Serrano, M.; Valero, D. Effect of salicylic acid treatment on reducing chilling injury in stored pomegranates. Postharvest Biol. Technol. 2009, 53, 152–154. [Google Scholar] [CrossRef]
  128. Davarynejad, G.H.; Zarei, M.; Nasrabadi, M.E.; Ardakani, E. Effects of salicylic acid and putrescine on storability, quality attributes and antioxidant activity of plum cv. ‘Santa rosa’. J. Food Sci. Technol. 2015, 52, 2053–2062. [Google Scholar] [CrossRef] [Green Version]
  129. Dokhanieh, A.Y.; Aghdam, M.S.; Fard, J.R.; Hassanpour, H. Postharvest salicylic acid treatment enhances antioxidant potential of cornelian cherry fruit. Sci. Hortic. 2013, 154, 31–36. [Google Scholar] [CrossRef]
  130. Lu, X.H.; Sun, D.Q.; Li, Y.H.; Shi, W.Q.; Sun, G.M. Pre- and post-harvest salicylic acid treatments alleviate internal browning and maintain quality of winter pineapple fruit. Sci. Hortic. 2011, 130, 97–101. [Google Scholar] [CrossRef]
  131. Adhikary, T.; Gill, P.S.; Jawandha, S.K.; Bhardwaj, R.D.; Anurag, R.K. Browning and quality management of pear fruit by salicylic acid treatment during low temperature storage. J. Sci. Food Agr. 2021, 101, 853–862. [Google Scholar] [CrossRef]
  132. El-Mogy, M.M.; Ali, M.R.; Darwish, O.S.; Rogers, H.J. Impact of salicylic acid, abscisic acid, and methyl jasmonate on postharvest quality and bioactive compounds of cultivated strawberry fruit. J. Berry Res. 2019, 9, 333–348. [Google Scholar] [CrossRef]
  133. Kumari, P.; Barman, K.; Patel, V.B.; Siddiqui, M.W.; Kole, B. Reducing postharvest pericarp browning and preserving health promoting compounds of litchi fruit by combination treatment of salicylic acid and chitosan. Sci. Hortic. 2015, 197, 555–563. [Google Scholar] [CrossRef]
  134. Shi, Z.J.; Wang, F.; Lu, Y.Y.; Deng, J. Combination of chitosan and salicylic acid to control postharvest green mold caused by penicillium digitatum in grapefruit fruit. Sci. Hortic. 2018, 233, 54–60. [Google Scholar] [CrossRef]
  135. Creelman, R.A.; Mullet, J.E. Biosynthesis and action of jasmonates in plants. Annu. Rev. Plant Biol. 1997, 48, 355–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Mustafa, M.A.; Ali, A.; Seymour, G.; Tucker, G. Enhancing the antioxidant content of carambola (averrhoa carambola) during cold storage and methyl jasmonate treatments. Postharvest Biol. Technol. 2016, 118, 79–86. [Google Scholar] [CrossRef]
  137. Huang, X.J.; Li, J.; Shang, H.L.; Meng, X. Effect of methyl jasmonate on the anthocyanin content and antioxidant activity of blueberries during cold storage. J. Sci. Food Agr. 2015, 95, 337–343. [Google Scholar] [CrossRef]
  138. Boonyaritthongchai, P.; Supapvanich, S. Effects of methyl jasmonate on physicochemical qualities and internal browning of ‘queen’ pineapple fruit during cold storage. Hortic. Environ. Biotechnol. 2017, 58, 479–487. [Google Scholar] [CrossRef]
  139. Liu, H.R.; Meng, F.L.; Miao, H.Y.; Chen, S.S.; Yin, T.T.; Hu, S.S.; Shao, Z.Y.; Liu, Y.Y.; Gao, L.X.; Zhu, C.Q.; et al. Effects of postharvest methyl jasmonate treatment on main health-promoting components and volatile organic compounds in cherry tomato fruits. Food Chem. 2018, 263, 194–200. [Google Scholar] [CrossRef]
  140. Cai, Y.T.; Cao, S.F.; Yang, Z.F.; Zheng, Y.H. Meja regulates enzymes involved in ascorbic acid and glutathione metabolism and improves chilling tolerance in loquat fruit. Postharvest Biol. Technol. 2011, 59, 324–326. [Google Scholar] [CrossRef]
  141. Kazan, K.; Manners, J.M. MYC2: The master in action. Mol. Plant 2013, 6, 686–703. [Google Scholar] [CrossRef] [Green Version]
  142. Li, Z.L.; Min, D.D.; Fu, X.D.; Zhao, X.M.; Wang, J.H.; Zhang, X.H.; Li, F.J.; Li, X.A. The roles of slmyc2 in regulating ascorbate-glutathione cycle mediated by methyl jasmonate in postharvest tomato fruits under cold stress. Sci. Hortic. 2021, 288, 110406. [Google Scholar] [CrossRef]
  143. Baek, M.W.; Choi, H.R.; Jae, L.Y.; Kang, H.M.; Lee, O.H.; Jeong, C.S.; Tilahun, S. Preharvest treatment of methyl jasmonate and salicylic acid increase the yield, antioxidant activity and gaba content of tomato. Agronomy 2021, 11, 2293. [Google Scholar] [CrossRef]
  144. Garcia-Pastor, M.E.; Serrano, M.; Guillen, F.; Gimenez, M.J.; Martinez-Romero, D.; Valero, D.; Zapata, P.J. Preharvest application of methyl jasmonate increases crop yield, fruit quality and bioactive compounds in pomegranate ‘mollar de elche’ at harvest and during postharvest storage. J. Sci. Food Agr. 2020, 100, 145–153. [Google Scholar] [CrossRef] [PubMed]
  145. Wang, H.B.; Wu, Y.; Yu, R.P.; Wu, C.E.; Fan, G.J.; Li, T.T. Effects of postharvest application of methyl jasmonate on physicochemical characteristics and antioxidant system of the blueberry fruit. Sci. Hortic. 2019, 258, 108785. [Google Scholar] [CrossRef]
  146. Li, X.A.; Li, M.L.; Wang, J.; Wang, L.; Han, C.; Jin, P.; Zheng, Y.H. Methyl jasmonate enhances wound-induced phenolic accumulation in pitaya fruit by regulating sugar content and energy status. Postharvest Biol. Technol. 2018, 137, 106–112. [Google Scholar] [CrossRef]
  147. Wu, X.; Ren, J.; Huang, X.Q.; Zheng, X.Z.; Tian, Y.C.; Shi, L.; Dong, P.; Li, Z.G. Melatonin: Biosynthesis, content, and function in horticultural plants and potential application. Sci. Hortic. 2021, 288, 110392. [Google Scholar] [CrossRef]
  148. Wang, F.; Zhang, X.P.; Yang, Q.Z.; Zhao, Q.F. Exogenous melatonin delays postharvest fruit senescence and maintains the quality of sweet cherries. Food Chem. 2019, 301, 125311. [Google Scholar] [CrossRef]
  149. Aghdam, M.S.; Luo, Z.S.; Li, L.; Jannatizadeh, A.; Fard, J.R.; Pirzad, F. Melatonin treatment maintains nutraceutical properties of pomegranate fruits during cold storage. Food Chem. 2020, 303, 125385. [Google Scholar] [CrossRef]
  150. Gautier, H.; Massot, C.; Stevens, R.; Serino, S.; Genard, M. Regulation of tomato fruit ascorbate content is more highly dependent on fruit irradiance than leaf irradiance. Ann. Bot. 2009, 103, 495–504. [Google Scholar] [CrossRef] [Green Version]
  151. 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]
  152. Lado, J.; Alos, E.; Rodrigo, M.J.; Zacarias, L. Light avoidance reduces ascorbic acid accumulation in the peel of citrus fruit. Plant Sci. 2015, 231, 138–147. [Google Scholar] [CrossRef] [PubMed]
  153. Ntagkas, N.; Woltering, E.; Bouras, S.; de Vos, R.C.H.; Dieleman, J.A.; Nicole, C.C.S.; Labrie, C.; Marcelis, L.F.M. Light-induced vitamin c accumulation in tomato fruits is independent of carbohydrate availability. Plants 2019, 8, 86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Zhang, Y.T.; Ntagkas, N.; Fanourakis, D.; Tsaniklidis, G.; Zhao, J.T.; Cheng, R.F.; Yang, Q.C.; Li, T. The role of light intensity in mediating ascorbic acid content during postharvest tomato ripening: A transcriptomic analysis. Postharvest Biol. Technol. 2021, 180, 111622. [Google Scholar] [CrossRef]
  155. Vithana, M.D.K.; Singh, Z.; Johnson, S.K. Cold storage temperatures and durations affect the concentrations of lupeol, mangiferin, phenolic acids and other health-promoting compounds in the pulp and peel of ripe mango fruit. Postharvest Biol. Technol. 2018, 139, 91–98. [Google Scholar] [CrossRef]
  156. Kantakhoo, J.; Imahori, Y. Antioxidative responses to pre-storage hot water treatment of red sweet pepper (Capsicum annuum L.) fruit during cold storage. Foods 2021, 10, 3031. [Google Scholar] [CrossRef]
  157. Perin, E.C.; Messias, R.D.; Borowski, J.M.; Crizel, R.L.; Schott, I.B.; Carvalho, I.R.; Rombaldi, C.V.; Galli, V. Aba-dependent salt and drought stress improve strawberry fruit quality. Food Chem. 2019, 271, 516–526. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The AsA metabolic pathway in plants. The main pathways for ascorbic acid synthesis in plants are L-galactose pathway, Myoinositol pathway, L-gulose pathway, and D-galacturonate pathway. Alase, aldono-lactonase; AO, ascorbate oxidas; APX, ascorbate peroxidase; DHA, dehydroascorbic acid; DHAR, dehydroascorbate reductase; GalUR, D-galacturonate reductase; GDH, L-galactose dehydrogenase; GGP, GDP-L-galactose-phosphorylase; GLDH, L-galactono-1,4-lactone dehydrogenase; GME, GDP-D-mannose3′, 5′-epimerase; GMP, GDP-mannose pyrophosphorylase; GPP, L-galactose-1-phosphate phosphatase; GSH, glutathione; GSSH, oxidized glutathione; GulLO, L-gulono-1,4-lactone oxidase; IMP, myoinositol monophosphatase; MDHA, monodehydroascorbic acid; MDHAR, monodehydroascorbate reductase; MIOX, myoinositol oxygenase; MIPS, myoinositol phosphate synthase; PGI, phosphoglucose isomerase; PMI, phosphomannose isomerase; PMM, phosphomannomutase.
Figure 1. The AsA metabolic pathway in plants. The main pathways for ascorbic acid synthesis in plants are L-galactose pathway, Myoinositol pathway, L-gulose pathway, and D-galacturonate pathway. Alase, aldono-lactonase; AO, ascorbate oxidas; APX, ascorbate peroxidase; DHA, dehydroascorbic acid; DHAR, dehydroascorbate reductase; GalUR, D-galacturonate reductase; GDH, L-galactose dehydrogenase; GGP, GDP-L-galactose-phosphorylase; GLDH, L-galactono-1,4-lactone dehydrogenase; GME, GDP-D-mannose3′, 5′-epimerase; GMP, GDP-mannose pyrophosphorylase; GPP, L-galactose-1-phosphate phosphatase; GSH, glutathione; GSSH, oxidized glutathione; GulLO, L-gulono-1,4-lactone oxidase; IMP, myoinositol monophosphatase; MDHA, monodehydroascorbic acid; MDHAR, monodehydroascorbate reductase; MIOX, myoinositol oxygenase; MIPS, myoinositol phosphate synthase; PGI, phosphoglucose isomerase; PMI, phosphomannose isomerase; PMM, phosphomannomutase.
Plants 11 01602 g001
Figure 2. The AsA transport in plant cell. AsA is synthesized on the inner mitochondrial membrane and transported into the cytoplasm. AsA in the cytoplasm can enter organelles, such as vacuoles, chloroplasts, and nuclei, through diffusion or carriers. In addition, AsA can also be transported outside the cell membrane by simple diffusion or transport proteins. DHA in the apoplast can also enter the cell membrane and participate in the regeneration of AsA. Arrows indicate the direction of material transport. NAT, nucleobase/ascorbate transporter; Cytb, cytochrome b.
Figure 2. The AsA transport in plant cell. AsA is synthesized on the inner mitochondrial membrane and transported into the cytoplasm. AsA in the cytoplasm can enter organelles, such as vacuoles, chloroplasts, and nuclei, through diffusion or carriers. In addition, AsA can also be transported outside the cell membrane by simple diffusion or transport proteins. DHA in the apoplast can also enter the cell membrane and participate in the regeneration of AsA. Arrows indicate the direction of material transport. NAT, nucleobase/ascorbate transporter; Cytb, cytochrome b.
Plants 11 01602 g002
Figure 3. Effects of hormones and environmental factors on the accumulation of AsA in fruits. Light, heat, mild drought, and mild salt can stimulate the accumulation of ascorbic acid in the fruit. In addition, phytohormones, such as gibberellin, abscisic acid, salicylic acid, jasmonic acid, and melatonin, can also regulate the accumulation of ascorbic acid in fruits. These environmental factors and phytohormones affect fruit ripening and stress resistance by regulating the accumulation of ascorbic acid in fruits.
Figure 3. Effects of hormones and environmental factors on the accumulation of AsA in fruits. Light, heat, mild drought, and mild salt can stimulate the accumulation of ascorbic acid in the fruit. In addition, phytohormones, such as gibberellin, abscisic acid, salicylic acid, jasmonic acid, and melatonin, can also regulate the accumulation of ascorbic acid in fruits. These environmental factors and phytohormones affect fruit ripening and stress resistance by regulating the accumulation of ascorbic acid in fruits.
Plants 11 01602 g003
Table 1. AsA content in fruits of different cultivars.
Table 1. AsA content in fruits of different cultivars.
Common NameCultivarContent of AsA
(mg/100 g FW)
Reference
TomatoSolanum pennellii4.40–17.61[18]
Cherry31.98–42.68[19]
Monika41.41–53.10[19]
Isabella41.19–48.65[19]
HLY5.58–18[20]
Rio Grande6.78–10.8[20]
KiwifruitHayward51.3–79.7[21]
Jiangxi 79-153.8–93.6[21]
Awaji22.2–28.8[21]
Kosui31.5–50.3[21]
StrawberryPraratchatan63.73–72.57[22]
Sagahonoka56.8[23]
Sugyeong108.1[23]
PearBanda10.2[24]
Limon10.1[24]
İncir4.4[24]
OrangeMandarin41.3 [25]
Hamlin62.7[25]
Salustiana56.8[25]
WatermelonCrimson sweet11.86–15.27[26]
Giza15.58–30.45[26]
Dumara10.84–23.34[26]
LemonOvale di Sorrento29.91[27]
Sfusato Amalfitano27.71[27]
Femminello Cerza26.90[27]
Femminello Adamo26.69[27]
AppleM. pumila ‘Saiwaihong’0.96 ± 0.06[28]
Yantai Fuji No. 30.37 ± 0.09[28]
Xinshiji0.25 ± 0.05[28]
Liuyuehong0.20 ± 0.04[28]
Gala0.10 ± 0.06[28]
Starkrimson0.05 ± 0.02[28]
GrapeVrboska0.46[29]
Jakubské0.41[29]
Perlette0.36[29]
PepperSegana71.99[30]
Catas68.56[30]
Domba61.22[30]
Table 2. Relationship between expression patterns of AsA metabolism-related genes and AsA accumulation in fruits.
Table 2. Relationship between expression patterns of AsA metabolism-related genes and AsA accumulation in fruits.
Common NameGene NameGene SourceStrategyChange of AsA ContentFold ChangeReference
tomatoGMP3tomatooverexpressionup1.1–1.6[85]
GMP3tomatoRNAidown1.3–2.4[85]
GME1tomatooverexpressionup1.6[86]
GME1tomatoRNAino change0[87]
GME2tomatooverexpressionup1.2[86]
GME2tomatoRNAino change0[87]
GME1 × GME2tomatoRNAidown0.2–0.6[38]
GGPtomatooverexpressionno change0[35]
GGPtomatomutantdown0.5[88,89]
GGPkiwifruitoverexpressionup2.0[90]
GGP × GPPtomatooverexpressionno change0[35]
SlIMP3tomatooverexpressionup0.3–0.6[93]
SlIMP3tomatoantisensedown0.3–0.7[93]
GLDHtomatoRNAino change0[17]
GLOaseratoverexpressionup1.5[94]
MIOXArabidopsisoverexpressionup1.4[95]
MIOX4tomatooverexpressionup1.5–2.3[96]
GalURstrawberryoverexpressionup1.2–2.5[97]
APXtomatoRNAiup1.4–2.2[100]
AOtomatoRNAiup0.3[101]
DHAR1tomatooverexpressionup0.4[102]
DHAR2tomatooverexpressionno change0[102]
MDHARtomatooverexpressionup0.7[56]
kiwifruitGGP3kiwifruitoverexpressionup2.0–6.4[91,92]
strawberryGGPkiwifruitoverexpressionup2.0[90]
Table 3. Regulation of AsA metabolism-related genes at the transcriptional and protein levels.
Table 3. Regulation of AsA metabolism-related genes at the transcriptional and protein levels.
Common NameGene NameGene SourceTarget GeneCombinationEffectReference
tomatoICE1tomatonot describednot describedPositively regulate the accumulation of AsA[105]
L1L4tomatonot describednot describedNegatively regulate the accumulation of AsA[106]
BZR1-1DArabidopsisnot describednot describedPositively regulate the accumulation of AsA[107]
MADS7banananot describednot describedPositively regulate the accumulation of AsA[109]
NFYA10tomatoGME1protein-DNANegatively regulate the expression of GME1 and the accumulation of AsA.[110]
HZ24tomatoGMP3protein-DNAPositively regulate the expression of GMP3 and the accumulation of AsA.[111]
HZ24tomatoGME2protein-DNAPositively regulate the expression of GME2 and the accumulation of AsA.[111]
HZ24tomatoGGPprotein-DNAPositively regulate the expression of GGP and the accumulation of AsA.[111]
appleERF98appleGMP1protein-DNAPositively regulate the expression of GMP1 and the accumulation of AsA.[36]
MYB1appleDHARprotein-DNAPositively regulate the expression of DHAR and the accumulation of AsA.[113]
SCL26.1appleMDHARprotein-DNAPositively regulate the expression of MDHAR and negatively regulate the accumulation of AsA.[114]
AMR1L1appleGMP1protein-proteinStimulate GMP1 degradation and negatively regulate the accumulation of AsA.[36]
mdm-miR171iappleSCL26.1RNA-DNAStimulate SCL26.1 degradation and positively regulate the accumulation of AsA.[114]
kiwifruitERF91kiwifruitGGP3protein-DNAPositively regulate the expression of GGP3 and the accumulation of AsA.[112]
MYBS1kiwifruitGGP3protein-DNAPositively regulate the expression of GGP3 and the accumulation of AsA.[92]
ESE3kiwifruitGGP3protein-proteinPositively regulate the accumulation of AsA[91]
MYBRkiwifruitGGP3protein-proteinPositively regulate the accumulation of AsA[91]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, X.; Gong, M.; Zhang, Q.; Tan, H.; Li, L.; Tang, Y.; Li, Z.; Peng, M.; Deng, W. Metabolism and Regulation of Ascorbic Acid in Fruits. Plants 2022, 11, 1602. https://doi.org/10.3390/plants11121602

AMA Style

Zheng X, Gong M, Zhang Q, Tan H, Li L, Tang Y, Li Z, Peng M, Deng W. Metabolism and Regulation of Ascorbic Acid in Fruits. Plants. 2022; 11(12):1602. https://doi.org/10.3390/plants11121602

Chicago/Turabian Style

Zheng, Xianzhe, Min Gong, Qiongdan Zhang, Huaqiang Tan, Liping Li, Youwan Tang, Zhengguo Li, Mingchao Peng, and Wei Deng. 2022. "Metabolism and Regulation of Ascorbic Acid in Fruits" Plants 11, no. 12: 1602. https://doi.org/10.3390/plants11121602

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