Next Article in Journal
Immune Tolerance in the Oral Mucosa
Previous Article in Journal
RNAi Crop Protection Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transcriptional Regulation of Ripening in Chili Pepper Fruits (Capsicum spp.)

by
Maria Guadalupe Villa-Rivera
and
Neftalí Ochoa-Alejo
*
Departamento de Ingeniería Genética, Unidad Irapuato, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional, Irapuato 36824, Mexico
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(22), 12151; https://doi.org/10.3390/ijms222212151
Submission received: 12 October 2021 / Revised: 6 November 2021 / Accepted: 7 November 2021 / Published: 10 November 2021
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
Chili peppers represent a very important horticultural crop that is cultivated and commercialized worldwide. The ripening process makes the fruit palatable, desirable, and attractive, thus increasing its quality and nutritional value. This process includes visual changes, such as fruit coloration, flavor, aroma, and texture. Fruit ripening involves a sequence of physiological, biochemical, and molecular changes that must be finely regulated at the transcriptional level. In this review, we integrate current knowledge about the transcription factors involved in the regulation of different stages of the chili pepper ripening process.

Graphical Abstract

1. Introduction

The chili pepper belongs to the genus Capsicum and the Solanaceae family. The genus Capsicum encompasses 38 species, five of which have been domesticated: C. annuum, C. frutescens, C. chinense, C. baccatum and C. pubescens [1,2]. These species present different growth forms and a wide variety of shapes and colors of corollas and fruits [3]. Fresh or dry chili pepper fruits have economic and culinary importance, and they are cultivated and distributed worldwide; moreover, these fruits have generated growing interest because of their versatility and innumerable uses as condiments and medicinal properties [4,5]. Chili pepper fruits are commonly classified as berries, which are indehiscent fruits with many seeds bound to the pericarp. Capsicum fruits present variable shapes (spherical, conical, ovoid, fusiform, or elongate), colors (red, orange, yellow, white, purple or brown) and pungent and nonpungent varieties [3,4]. Traditionally, Capsicum fruits have been used for their flavor, spice, and color, and applied to cure toothache, muscle pain, neuralgia, rheumatism, and diabetes. Additionally, extracts of chili pepper have been investigated as antimicrobial agents against fungi and viruses and as insecticidal, anthelmintic and larvicidal agents; moreover, the anti-inflammatory properties, cardiovascular protection effects, anti-obesity effects, antioxidant activities, anticancer activities, antiangiogenic and anti-neoplastic effects, and antithrombotic and vasodilatory properties of chili pepper extracts have been widely described [6]. The nutraceutical effects of extracts of chili pepper fruits are due to the diverse bioactive compounds located in the placenta, pericarp and seeds [7], and these bioactive metabolites consist of capsaicinoids, carotenoids, phenolic compounds (such as flavonoids) and volatile compounds [8,9]. In addition, Capsicum fruits have high levels of vitamin C, pro vitamin A, and minerals, such as calcium and iron, among others, all of which have health-promoting potential. Furthermore, chili pepper phytochemicals constitute natural compounds for the cosmetic, pharmaceutical, and agro-food industries, where they are used as substitutes for synthetic ingredients [10,11].
The maturation of fruits in plants involves different steps: fruit set, fruit development, and fruit ripening. Fruit ripening requires a sequence of biochemical and physiological transformations (modifications of the plant cell wall, conversion of starch to sugars, alterations in pigment biosynthesis, accumulation of aromatic volatile compounds, among others) that are regulated genetically and take place at the final stage of fruit development [12,13]. This process includes visual changes, such as fruit coloration, and changes in flavor, aroma, texture and palatability, which make the fruit edible, attractive and desirable for consumers and increase its quality and nutritional value [14]. According to physiological variations in the pattern of respiration during the ripening process, fruits are classified as climacteric (increase in levels of CO2 and ethylene production detected) and non-climacteric (no peak in ethylene emission detected). For example, tomato, apple, and avocado are classified as climacteric fruits, while lemon, pineapple and chili pepper represent important models for the analysis of non-climacteric fruits [15]. During the ripening process of chili pepper fruits (Figure 1), gene expression, metabolite accumulation and antioxidant activity change through the different stages of growth and maturation [16]. These changes must be highly controlled, and transcription factors (TFs) are crucial components for regulating the expression of genes involved in the molecular mechanisms that occur in this process [12]. Although transcriptional regulation of the ripening process in tomato (the closest relative to Capsicum) has been widely studied [17], both species likely have different ripening mechanisms and therefore different regulatory effects on gene expression [18]. Transcriptional regulation is extremely important during the ripening process, and during the ripening process between tomato and chili pepper fruits, differences are observed in the transcription of genes encoding TFs, suggesting the existence of key factors that differentiate the ripening processes in climacteric and non-climacteric fruits [19]. In this review, we integrated current knowledge about the transcription factors involved in chili pepper fruit development and incorporated available information about the transcriptional regulation of capsaicinoid, carotenoid, anthocyanin and ascorbic acid biosynthetic pathways as ripening-related processes.

2. Chili Pepper Ripening Regulation Process

Metabolic changes that occur during the ripening process of chili peppers cause transformations in flavor, color, texture and aroma. Accordingly, different transcription profiles of genes that encode key enzymes of metabolite biosynthetic pathways and distinct antioxidant activity patterns during chili pepper ripening have been described at the early (16–25 DPA), breaker (36–28 DPA) and later stages (43–48 DPA). For instance, the production of glycosides (quercetin, luteolin and apigenin) characterized the early stages, while the production of other metabolites, such as shikimic acid, putrescin and γ-aminobutyric acid (GABA), which were highly accumulated, gradually decreased at the breaker stage. Finally, the content of amino acids (leucine, isoleucine, proline and phenylalanine) and other compounds, such as kaempferol, capsaicin and dihydrocapsaicin, increased significantly at the later stage [16]. Transcriptomic analyses conducted in two stages of ripening (immature (green fruit) and mature (red fruit)) of chili pepper fruits revealed 1008 differentially expressed genes in these fruits. Distinct expression patterns are associated with different metabolic pathways, such as carotenoid and phenylpropanoid biosynthesis, plant hormone signal transduction, secondary metabolites and sesquiterpenoid and triterpenoid biosynthesis pathways. Additionally, a comparison between the transcription patterns of C. annuum L. (‘Tampiqueño 74’, a domesticated variety) and C. annuum L. var. glabriusculum (chiltepin wild variety) resulted in 53 differentially expressed genes associated with size, shape and secondary metabolite biosynthesis, suggesting that these differences could be related to the domestication process [20].
Although chili pepper has been used as a model for the study of non-climacteric fruits, variations in the respiration rates and ethylene production among cultivars of Capsicum have been detected because of the diversity of this genus [21]. Accordingly, genes encoding key enzymes of ethylene biosynthesis, such as 1-aminoacyl cyclopropane 1-carboxilate (ACC) synthase (ACS) and ACC oxidase (ACO), were not expressed in chili pepper fruits [19]. Nevertheless, a proteomic analysis conducted with the aim of identifying the differentially expressed proteins in three ripening stages (green, breaker red and light red) of C. annuum L. demonstrated that at least one isoform of ACO (CaACO4) was expressed at the breaker stage of chili pepper ripening, and these results suggest that some aspects of regulation may be conserved in climacteric and non-climacteric fruits [22]. Other evaluations of the levels of transcription of genes encoding several isoforms of ACO, ACS and the ethylene receptor (ETR) demonstrated that except for CaACO4, the expression of these genes was limited to Capsicum fruits; additionally, the activity levels of ACS and ACC were very low, suggesting that ACS activity was the limiting step in the ethylene biosynthesis pathway in chili pepper, thus leading to the restriction of ACC accumulation [18]. Treatment with ethylene or 1-methylclopropene likely differentially regulated ethylene production in non-climacteric chili pepper fruit compared with climacteric tomato; moreover, the presence and upregulation of the CaACO4 gene in C. annuum L. fruits suggests that ethylene-independent pathways may exist in non-climacteric fruits [18,23].
The transcripts of three genes encoding TFs related to ethylene-mediated signaling (EIL1-like) were detected in C. annuum at the brown stage [19]. EIL TFs are homologous to the ethylene insensitive (EIN) TFs reported in Arabidopsis thaliana and Solanum lycopersicum, and they act downstream from ethylene receptors in the ethylene signal transduction pathway and are described as positive regulators of ethylene responses [24]. These results might suggest that activation of ethylene signaling and the subsequent alteration of ethylene sensitivity could be conserved in non-climacteric fruits [18,19]. In contrast, upregulation of ethylene receptor-like protein (ETR), a key regulator of the ethylene transcription pathway, was detected in a transcriptomic analysis of C. frutescens and C. annuum performed at different stages of ripening, suggesting that at least some species of chili pepper exhibit climacteric behavior [20,25].
A comparative transcriptome analysis performed between tomato (climacteric) and chili pepper (non-climacteric) showed that the transcript profiles of genes encoding TFs such as the ripening inhibitor (RIN; a regulator or ripening in tomato) [26], tomato AGAMOUS-LIKE1 (TAG1; which has an important role in the regulation of both fleshy fruit expansion and ripening process) [27], and non-ripening (NOR) (Figure 2A) [12] are conserved in both types of fruits [18]. In contrast, other genes encoding TFs involved in the regulation of fruit ripening, such as Colorless ripening (Cnr; a major component in the regulatory network controlling tomato ripening) [28], Uniform (encodes a Golden 2-like (SlGLK2) transcription factor involved in chlorophyll accumulation and distribution during fruit development) [29], and LeHB-1 (participates in the control of ripening) [30], were expressed at low levels during chili pepper ripening and highly expressed during the tomato ripening process [18].
Some transcription factors have been identified and proposed as regulators of the transcription of structural genes that encode proteins involved in the ripening process of chili pepper fruits. Among the differentially expressed genes (DEGs) identified in the transcriptomic analysis conducted in immature and mature chili pepper fruits, only four genes encoding TFs were upregulated (FUL2, two bZIPs and one homeobox TF), while 25 TF genes were downregulated [20].
With the aim of evaluating the role of some TFs during the chili pepper ripening process, functional and interaction assays were performed. The gene expression profile and yeast two-hybrid assays suggested that the interaction of two transcription factors (CaMADS, subfamily AGL2; and CAMADS6, subfamily SQUA) could have an important role in the fruit development of a hot chili pepper [31]. The MADS (MCM1 AG DEFA SRF) box is a conserved family of TFs involved in the control of development and signal transduction [32]. Structurally, MADS-box TFs bear two conserved domains: the MADS domain of 56 amino acids, which includes the motif of union to the consensus DNA sequence (Table 1) and the heterodimerization zone, and the K domain (required for MADS protein interaction), which plays an important role in MADS-box proteins. Moreover, the structure of these TFs includes two divergent regions: the interdomain region (I-region) located between the MADS and K domains and the C-region, which is located at the C-terminus [33,34,35,36]. With the aim of investigating the function of other MADS-box TFs from chili pepper, the role of CaMADS-ripening inhibitor (CaMADS-RIN), a TF belonging to the SEPALLATA (SEP) group, was evaluated. The CaMADS-RIN gene showed high transcription levels during the ripening process of C. annuum, whereas its heterologous expression in tomato plants revealed that this TF had a positive role in ethylene biosynthesis and fruit ripening in tomato, suggesting its involvement in the fruit ripening of either climacteric or non-climacteric fruits [37]. A comparative genomic analysis based on the tomato genome was carried out to identify genes involved in the development/ripening process of chili pepper fruits; in this case, 32 orthologous genes encoding proteins involved in cell wall formation, metabolite accumulation, coloring, softening and aroma production were expressed, and only 12 of them were differentially expressed in Capsicum. Interestingly, it was possible to identify a gene with high levels of expression at the breaker stage and mature fruits by quantitative expression analysis; this gene was identified as an ortholog gene of MADS-RIN by Dong in 2014 [37], which supported its possible participation in the fruit ripening process of Capsicum [25].
Moreover, a wide genomic and subcellular localization analysis of the auxin response factor (ARF) family from C. annuum L. identified CaARF genes that exhibited higher levels of expression in flowers (CaARF18) and fruits (CaARF4 and CaARF5) than in other chili pepper organs [38]. ARFs are a family of TFs involved in the transcriptional regulation of genes of the auxin response, and they interact with the auxin response elements located in the promoters of genes via the conserved B3 DNA binding domain (DBD) of ARFs. In addition, a variable middle region (MR) that works as a repression or activation domain and a conserved C-terminal domain (CTD), which enables protein–protein interactions, compose the ARF structure [50,51]. In this sense, it is possible to hypothesize that CaARF18, CaARF4 and CaARF5 could participate in the regulation of chili pepper fruit development and ripening processes.
Other important candidates for chili pepper ripening regulation are WRKY TFs. CaWRKYs are expressed at higher levels in vegetative tissues, and a previous study indicated that almost 60% of these TFs were expressed during chili pepper fruit maturation and regulated by plant hormones and abiotic stress, suggesting their key regulatory role in the ripening process [52]. A highly conserved WRKY domain at the N-terminal region characterizes WRKY TFs. The WRKY domain contains a conserved peptide WRKYGQK and a zinc-finger structure (Cx4–5Cx22–23HxH or Cx7Cx23HxC) [53].
Finally, the role of noncoding RNAs (ncRNAs) was evaluated. The transcriptional profile of ncRNAs showed that they are differentially expressed in mature green and red ripe chili pepper fruits. Moreover, the targets of these ncRNAs were suggested to include TFs, such as ARF, bHLH, bZIP, ERF, MYB, NAC and WRKY, and enzymes involved in cell wall remodeling, pigment accumulation, fruit flavor and aroma development, and ethylene production; thus, these ncRNAs constitute a new level of control during the fruit ripening process [54].

3. Transcriptional Regulation of Capsaicinoids Biosynthesis

A characteristic quality of chili pepper fruits is pungency, and they produce and accumulate hot compounds called capsaicinoids, which are alkaloids found exclusively in Capsicum species [55]. The most abundant capsaicinoids in chili pepper fruits are capsaicin (trans-8-methyl-N-vanillyl-6-nonenamide), a derivative of homovallinic acid, and dihydrocapsaicin. Interestingly, capsaicin has become a very important bioactive compound as a cancer preventive agent and potentially as a treatment against various types of cancer [56]. Capsaicinoid accumulation in Capsicum fruits is genetically determined and depends on the developmental stage, and it occurs in the fruits from the early stages (ten days after pollination), starting at the instant that growth begins, reaches a maximum concentration at 33–40 dpa depending on the cultivar, and then decreases due to the activity of different peroxidases [57,58]. Biotic (wounding, attack by pathogens or insects) and abiotic factors (temperature, light, water, levels of CO2, altitude, and nutrient availability) influence the content and accumulation of capsaicinoids. Abiotic stresses stimulate a drastic change in the expression levels of genes involved in the production of capsaicinoids [59,60]. For instance, in hot chili pepper, the level of pungency is highly variable in response to temperature and the accumulation of capsaicinoids was reported to increase by exposure to water stress. Finally, mineral nutrition (mainly nitrogen and potassium) has been observed to have a key role in the metabolism of capsaicinoids [57]. Moreover, the differences in the levels of pungency in distinct species of Capsicum are partially dependent on the genotype. The absence of capsaicinoids in species or varieties of chili pepper is due to the presence of mutations in genes encoding key enzymes or TFs that regulate the capsaicinoid biosynthetic pathway or related pathways [59,60]. Finally, the content of capsaicinoids in chili pepper fruits depends on the type of tissue, with synthesis and accumulation of these compounds occurring in the epidermal cells of placental tissues and an ambiguous boundary observed in the placental septum and pericarp [61].
Capsaicinoids are produced by the convergence of two metabolic pathways: the phenylpropanoid and branched-chain fatty acid pathways [62]. Transcriptional regulation of both metabolic pathways, as well as amino acid precursor synthesis, results in determinants of capsaicinoid production and accumulation (Figure 2B) [60]. The complexity of the regulatory network of capsaicinoid biosynthesis was evidenced through an analysis of the expression profile of fruits of the Indian chili pepper C. frutescens cv. Guijianwang during different developmental stages, which found that at least 20 candidate enzyme-encoding genes could be related to capsaicinoid biosynthesis [63].
With the aim of elucidating the mechanisms of transcriptional regulation of capsaicinoid biosynthesis, research has evaluated the potential participation of the MYB TF family in the regulation of several secondary metabolism pathways (lignin, vitamin C, phenylpropanoid, capsaicinoid, and carotenoid biosynthesis) [47]. Structurally, MYB TFs are organized in two regions: an N-terminal conserved DNA-binding domain of 52 amino acids (R2R3-type domain in plants), and a variable modulator region located at the C-terminal region responsible for the regulatory activity [64,65]. MYB TFs are involved in several cellular processes, such as cellular and organ morphogenesis, primary and secondary metabolism, chloroplast development, cell cycle control, and biotic and abiotic stress responses [43,65].
The isolation and characterization of CaMYB31, an R2R3-MYB gene from C. annuum ‘Tampiqueño 74’, and a positive correlation was observed between the transcription profile of CaMYB31 and structural genes encoding enzymes of capsaicinoid biosynthesis [44]. Additionally, silencing of CaMYB31 resulted in a decrease in structural gene expression and capsaicinoid content. Remarkably, the expression levels of CaMYB31 were affected by plant hormones, wounding, temperature, and light. These results together suggest that the CaMYB31 TF has an important role in the regulation of the capsaicinoid biosynthesis pathway [44]. Moreover, a homologous gene of CaMYB31 in fruits of C. annuum with different degrees of pungency (pungent cultivar ‘Tean’, and nonpungent accession ‘YMC334’) was identified [66], and this homolog was described as a single recessive gene named Pun3. Interestingly, Pun3 was only transcribed in the pungent cultivar ‘Tean’, and a sequence analysis detected a premature stop codon in Pun3 from the nonpungent accession; consequently, genes encoding enzymes of the capsaicinoid biosynthetic pathway were significantly downregulated in ‘YCM334’ compared to the pungent chili pepper cultivar—these results confirm the role of MYB31 as a master regulator of capsaicinoid biosynthesis [66]. Accordingly, a genetic and functional analysis was performed in five Capsicum species, including C. chinense (with a high content of capsaicinoids), and it identified the locus Cap1 (identical to Pun3), which encodes the MYB31 TF and contributes to the level of pungency. MYB31 is specifically expressed in placental tissue and activates the transcription of capsaicinoid biosynthesis-related genes, resulting in genus-specialized metabolite production. Additionally, an increase in the expression of MYB31 (from C. chinense) caused by natural variations in its promoter allowed for the binding of the transcriptional activator WRKY9 and triggered an increase in the expression of capsaicinoid biosynthesis-related genes (Table 1) [67,68].
In addition to CaMYB31, the jasmonate-inducible transcription factor CaMYB108 was described [47]. Virus-induced silencing (VIGS) of CaMYB108 led to a reduction in the transcript levels of capsaicinoid biosynthetic genes and capsaicinoid content in C. annuum [45]. Moreover, a weighted gene coexpression network analysis (WGCNA) identified a module related to capsaicinoid biosynthesis within this module, and a MYB TF (CaMYB48) was described as a key regulator of capsaicinoid biosynthesis in chili pepper [46]. Finally, expression analyses of CaR2R3-MYB genes and capsaicinoid biosynthesis-related genes reported that at least six CaR2R3-MYBs may regulate capsaicinoid biosynthesis in chili pepper fruits [69]. Interestingly, expression analysis of the candidates mentioned above (CaMIB31, CaMYB108 and CaMYB48) and the six CaR2R3-MYBs CaMYB47, CaMYB64, CaMYB73, CaMYB74, CaMYB87 and CaMYB92 showed that only CaMYB31 had a positive correlation with the transcription patterns of genes related to the capsaicinoid biosynthesis pathway [47]. Additionally, coexpression analysis allowed us to identify three strong candidate genes for capsaicinoid biosynthesis regulation: CaMYB103, CaMYB115 and CaDIV14, whose expression profiles correlated positively with those of AT3 (encoding an acyl transferase; putatively capsaicinoid synthase) and CaMYB31 [47]. In particular, CaDIV14 was grouped with DIVARICATA-like genes, which encode a subgroup of MYB-related proteins involved in the determination of the dorsoventral asymmetry pattern in Antirrhinum flowers [70].
On the other hand, multiple members of the ethylene response factor (ERF) transcription factor family have been proposed as regulator candidates of genes encoding structural capsaicinoid biosynthesis-related enzymes. ERFs belong to the apetala2/ERF (AP2/ERF) superfamily of TFs. Members of the ERF family possess a single DNA binding domain AP2/ERF, which is composed of 60–70 amino acids; moreover, the induction of ERF genes is not always dependent on ethylene action, and these genes are involved in the regulation of several biological processes related to growth and development as well as in the responses to biotic and abiotic stresses [41,71,72]. An evaluation of the transcription profiles of two ERF TFs (Erf and Jerf) in nine cultivars of chili pepper with different capsaicinoid contents demonstrated that these genes were positively correlated with pungency. Maximum levels of expression of both genes were detected at 16–20 dpa prior to the accumulation of capsaicinoids in the placental tissues. Thus, these two ERF family members could participate in the regulation of the capsaicinoid biosynthetic pathway [73]. Subsequent analysis confirmed that the transcription patterns of three genes encoding TFs of the ERF family (CaERF102, CaERF53, CaERF111 and CaERF92) were similar to the capsaicinoid accumulation pattern, and CaERF92, CaERF102 and CaERF111 may be involved in the regulation of capsaicinoid biosynthesis mediated by temperature [42].
Ultimately, TFs belonging to the basic/helix-loop-helix (bHLH) family have emerged as new regulator candidates for capsaicinoid accumulation in Capsicum placenta. This superfamily of proteins is characterized by the presence of the bHLH conserved domain of approximately 60 amino acids and two functionally distinctive regions: an N-terminal basic region of approximately 15 amino acids (mainly basic residues) involved in DNA binding, and a C-terminal HLH region that functions as a dimerization domain, which is composed of hydrophobic residues [74]. Accordingly, the expression levels of the CabHLH007, CabHLH009, CabHLH026, CabHLH063 and CabHLH086 genes were positively correlated with the profile of capsaicinoid accumulation [18]. Moreover, CabHLH007, CabHLH009, CabHLH026 and CabHLH086 could be involved in temperature-mediated capsaicinoid biosynthesis. Interestingly, yeast two-hybrid (Y2H) assays demonstrated that CabHLH007, CabHLH009, CabHLH026, CabHLH063 and CabHLH086 might interact with MYB31, the master regulator of capsaicinoid production [39].

4. Transcriptional Regulation of Carotenoid Biosynthesis

The different colors of chili pepper fruits are due to variations in carotenoid composition and content in the pericarp [10,75]. Carotenoids are lipophilic molecules chemically composed of 40 carbons with conjugated double bonds. According to their composition, they are classified as carotenes (containing C-H atoms) and xathophylls (containing C-H-O atoms) [76]. In plants, carotenoids have important functions because they constitute the major attractants for pollinators and enhance the flavor of food crops. Additionally, carotenoids offer photoprotection against photooxidative damage in plant cells, increase heat and light stress tolerance, are involved in photosystem assembly, represent precursors to phytohormones, such as abscisic acid (ABA) and strigolactons, and are rhizosphere signal molecules [77]. Similar to capsaicinoids, carotenoids have outstanding nutraceutical properties in the prevention and treatment of human diseases and present antioxidant, cardiovascular disorder preventive, anti-obesity, anti-inflammatory and cancer preventive activities [78].
The variable accumulation of carotenoids in the fruits of Capsicum spp. is determined genetically and depends primarily on the cultivar and the development and ripening stages. For instance, red chili pepper fruits present high contents of capsanthin, capsorubin, β-cryptoxanthin, β-carotene, zeaxanthin and antheraxanthin, whereas brown fruits accumulate the same pigments as red fruits as well as high amounts of chlorophyll B and lutein [79]. Yellow fruits contain high levels of violaxanthin and lutein and antheraxanthin in a minor proportion. Finally, the orange color of chili pepper fruits is conferred mainly by the accumulation of zeaxanthin and lutein [79,80].
In the initial stages of chili pepper fruit development, the characteristic immature green color is obviously due to the presence of chlorophyll. Chloroplast development, compartment size and chlorophyll content variations in immature chili pepper fruits have been associated with the allelic diversity of three genes: C. annuum Golden2-like (CaGLK2) [81], C. chinense zinc finger LSD-One Like 1 (CcLOL1) and CcAPRR2 [82]. Additionally, the color of mature chili pepper fruits is regulated by three independent loci: C1, C2 (encoding phytoene synthase (PSY)) and Y (encoding capsanthin/capsorubin synthase (CCS)) [83]. A recent report demonstrated that locus C1, which controls the white fruit color, encodes a homolog of a pseudoresponse regulator 2-like (PRR2) of A. thaliana. Virus-induced gene silencing (VIGS) assays performed in C. frutescens plants with a PRR2 construct produced a lighter fruit color [84]. Moreover, the overexpression of PRR2 in tomato showed enhanced levels of chlorophyll in fruits at the early stages of development and a slight increase in the amount of carotenoids in red fruits due to the expansion of the plastid number [85]. Moreover, PRR2 transcription appears to be regulated by the circadian rhythm and could interact with the abscisic acid insensitive 3 (ABI3) protein [86].
In plant cells, carotenoids are stored in chromoplasts (plastids that have lost photosynthetic activity) [87]. During the ripening process, the characteristic grana-intergranal structure of chloroplasts is degraded and substituted by nonchlorophyll single thylakoids derived from the inner envelope membrane, and these morphological changes are correlated with metabolic transformations, such as the loss of galactolipids, a slight increase in phospholipid content, and the esterification of xanthophylls with fatty acids [88,89]. Chromoplasts are unique plastids capable of accumulating massive amounts of carotenoids, and these pigments are stored in membranes and fibrilar, crystal and tubule structures [87]. Particularly, plastoglobules (oval or tubular structures rich in lipids that are located in all plastids) constitute active metabolic sites of carotenoid biosynthesis and recycling. Proteomic analysis of chromoplast plastoglobules revealed the presence of seven fibrillins and 25 proteins involved in the metabolism of lipids, isoprenoid-derived molecules, and carotenoid conversion [90]. Interestingly, differences in the abundance of chromoplast proteins were observed in other plastid types and among different crops, and such differences have been observed for Capsicum spp., which accumulate high levels of CCS enzyme and fibrillins inside the chromoplasts [91,92].
Carotenoids in plants are biosynthesized through the methylerythritol phosphate-derived pathway [93]. With the exception of ζ-carotene isomerase (Z-ISO), most carotenogenic enzymes have been localized in the chloroplast proteomes of chili pepper fruits [92,94]; additionally, the transcription profiles of genes encoding carotenogenic enzymes during the chili pepper fruit development process have been reported [95]. Phytoene, the first colorless carotenoid, is synthetized in the stroma of chromoplasts, and the final desaturation and cyclization stages that produce colored carotenoids occur in the membrane [96]. The biosynthesis of phytoene, carotene and capsanthin depends on the phospholipid environment inside the chromoplast and the development of new membranes during the conversion of chloroplasts into chromoplasts [88].
In plants, the biosynthesis of carotenoids is regulated at different levels, such as metabolite feedback and transcriptional and epigenetic control [97]. Mechanisms of the transcriptional regulation of carotenoid biosynthesis are highly relevant and have been widely studied in climacteric fruits, with tomato (Solanum lycopersicum) used as a model [17]; nevertheless, our understanding of the transcriptional regulation of carotenoid biosynthesis remains fragmented and incomplete, especially in non-climacteric fruits (Figure 2C) [98].
Transcriptional regulation of the first step of the carotenoid biosynthetic pathway seems to be ubiquitous in plants, and phytochrome interacting factor 1 (PIF1) downregulates the content of carotenoids by repressing the expression of the gene encoding the phytoene synthase (PSY) enzyme, which is the main rate-limiting enzyme of the pathway [99]. PIF proteins belong to the bHLH family of TFs, and they are categorized as negative regulators (repressors) of chlorophyll production [100]. PIFs are capable of transducing light signals to control transcription and, consequently, carotenoid accumulation [99]. Additionally, it has been suggested that carotenoid biosynthesis is subjected to fine regulation integrated by specific modules of antagonistic TFs and cofactors. In this sense, the transcriptional cofactor phytochrome-rapidly regulated 1 has been proposed to prevent the union of TFs to their target promoters, and this cofactor interacts with the PIF1 repressor and promotes the transcription of the PSY enzyme-encoding gene [101].
Although the transcriptional regulation of genes encoding carotenogenic enzymes is poorly understood in Capsicum spp., transcriptomic and functional analyses have identified the role of some TFs that control this important process in non-climacteric fruits. Transcriptomic comparisons during the maturation stages of climacteric and non-climacteric fruits have revealed that although there is a common set of metabolic genes that controls the biosynthesis and accumulation of carotenoids, some specific regulators could differ between climacteric and non-climacteric fruits, which is probably because of their distinct use and susceptibility to ethylene during the ripening process [102].
An analysis of the ripening of chili pepper fruits (C. frutescens) showed that ethylene emissions increased at the initial coloring stage and reached a peak at the brown stage [103]. In contrast, the levels of ABA showed a slight increase before the brown stage and peaked at the full red stage. RNA sequencing (RNA-seq) and VIGS analysis showed the role of ACO3 and NCED1/3 during the ripening of C. frutescens fruits. Higher ACO3 transcription is usually related to an ethylene emission peak, which positively regulates the change in fruit color through the carotenoid biosynthesis pathway, whereas higher NCED1/3 expression is associated with ABA accumulation and subsequent degradation of chlorophyll during the ripening process [103].
Some transcriptional regulators of the carotenoid biosynthesis pathway have been described for non-climacteric fruits. The citrus gene CrMB68, which encodes an R2R3-MYB TF, showed direct and negative regulation of β-carotene hydroxylase (CrBCH) and 9-cis-epoxycarotenoid dioxygenase (CrNCED) gene expression [104]. On the other hand, overexpression of the CsMADS6 gene in sweet orange (Citrus sinensis) triggered enhanced expression of the CsPSY, phytoene desaturase (CsPDS), carotene isomerase (CsCRTISO) and CsBCH genes, and inhibited the expression of the ε-lycopene cyclase (CsLCYE) gene. Additionally, a yeast one-hybrid assay established that the CsMADS6 TF interacts with the promoters of the β-lycopene cyclase 1 (CsLCYB1) and CsLCYB2 genes [105].
Regarding the transcription factors that may regulate the expression of genes encoding carotenogenic enzymes in chili pepper fruits, the expression profile of the ERF TFs CaERF82, CaERF97, CaERF66, CaERF107 and CaERF101 showed a positive correlation with the levels of β-carotene, zeaxanthin and capsorubin, suggesting their possible regulatory role in the carotenoid biosynthetic pathway [42]. Related analyses conducted with the bHLH family of TFs from chili pepper demonstrated that the transcription profiles of the CabHLH009, CabHLH032, CabHLH048, CabHLH095 and CabHLH100 genes were similar to the carotenoid biosynthesis profile in the pericarp, including lutein, zeaxanthin and capsorubin (Table 1) [39].
Transcriptomic analysis and carotenoid content profiling were carried out in orange and red chili peppers at 25, 40 and 55 days DPA. Twenty-three DEGs involved in the carotenoid biosynthetic pathway were reported [49]; interestingly, significant differences were observed in the expression profiles of the β-carotene hydroxylase 1 (CHYB1) and CCS genes of red and orange fruits during growth and maturation. Differences in the regulation of the CCS gene likely produce distinct accumulation of carotenoids and fruit colors.
On the other hand, a positive correlation in the expression profile of the CCS gene and two genes encoding TFs belonging to the MYB family, CADIV1 and CaMYB3R-5, has been reported [47], and these TFs constitute interesting candidates for future functional analysis.
Finally, a gene coexpression network analysis identified the following six TF candidates for CCS gene regulation: U-box domain-containing protein 52, GATA transcription factor 26, RING/FYVE/PHD-type, F-box protein SKIP23, CONSTANS-LIKE 9, and zinc finger family FYVE/PHD-type (Figure 2C, Table 1) [49]. However, additional experiments are required to confirm its regulatory role in the carotenoid biosynthetic pathway.

5. Transcriptional Regulation of the Anthocyanin Biosynthetic Pathway

Some species of chili pepper biosynthesize and accumulate anthocyanins in their tissues and organs (flowers, fruits and foliage), and these phenolic compounds are responsible for red, blue, and purple colors, with purple being the most frequent color observed in Capsicum spp. Anthocyanins are secondary metabolites derived from the phenylpropanoid pathway [106].
Because of their antioxidant properties, the effect of anthocyanins in the prevention of chronic cardiovascular diseases and diabetes has been studied, and their anticancer, anti-obesity, antimicrobial, cardiovascular disease prevention and visual health promotion activities have also been reported [107].
Anthocyanin accumulation is regulated by the interaction of MYB, MYC and WD40 TFs with promoters of structural genes of the anthocyanin biosynthetic pathway (Table 1, Figure 2D) [108]. In C. annum, differential expression of the MybA (containing the R3R3 domain) and Myc genes was correlated with the accumulation of anthocyanins in flowers and fruits but not in foliar tissues [109]. VIGS silencing of a MYB gene in C. eximium resulted in a decrease in the expression of chalcone synthase (CHS), chalcone isomerase (CHI), flavonoid 3′,5′-hydroxylase (F3´5´H), dihydroflavonol 4-reductase (DFR), and UDP-glucose: flavonoid 3-O-glucosyltransferase (3GT) genes, whereas silencing of WD40 led to a significant reduction in CHS, F3H, F3´5´H, DFR and 3GT transcription levels [48]. WD40 proteins are a family belonging to the β-propeller protein group and characterized by the presence of a core region composed of 40 amino acids flanked by a glycine-histidine dipeptide and a tryptophan-aspartate dipeptide (WD) [110]. C. annuum L. CaMYC (belonging to the bHLH family) has been described as an important regulator of anthocyanin biosynthesis by increasing the expression of structural genes, and together with CaMYB and CaWD40, CaMYC controls the transcription of genes encoding enzymes involved in the biosynthesis of anthocyanins in chili pepper [40].

6. Transcriptional Regulation of the Ascorbic Acid (AsA) Biosynthetic Pathway

At the mature stage, Capsicum fruits contain high levels of ascorbic acid (also denominated ascorbate or vitamin C) [111], which has antioxidant properties and constitutes an important micronutrient of chili pepper [9,11,112]. AsA has been cataloged as an uncommon antioxidant agent because it donates a single reducing equivalent, and monodehydroascorbate (the radical that AsA forms) reacts with radicals instead of with nonradicals [113]. In plants, ascorbate is involved in important metabolic processes, such as photosynthesis, photoprotection, and environmental stress and pathogen attack responses; additionally, AsA has a key role in signaling during cell division and cell expansion [114]. In humans, AsA plays an important role in health promotion because it has immunomodulatory and antimicrobial effects [115] and analgesic properties [116], and its effect has been recently evaluated for the treatment and prevention of SARS-CoV-2-related disease (COVID-19) [117]. Additionally, it plays an important role in the treatment and prevention of cancer [118].
The content of AsA varies among varieties of chili pepper fruits, with a gradual increase of ascorbic acid occurring based on the state of fruit development from green to red [119] and a subsequent decrease occurring at the late ripening stage (partially red and fully red) [120]. In sweet pepper, the AsA concentration increased by 47% during the transition from green to orange fruit, and it was correlated with the conversion of chloroplasts into chromoplast, the loss of chlorophyll and the accumulation of carotenoids [121]. In the whole fruits of the serrano cultivar ‘Tampiqueño 74’, the concentration of AsA increased from 20 DPA, showed a maximum peak at 40 DPA, and was maintained at 50 and 60 DPA [122]. In plants, AsA is biosynthesized via the D-mannose/L-galactose pathway [123], and the L-galactose pathway has been reported to represent the main mechanism for the production of vitamin C in the leaves and fruits of chili pepper [122]. Previous reports indicated that the transcript levels of genes encoding the key enzymes of the AsA biosynthetic pathway were not positively correlated with the AsA concentration in the pericarp of chili pepper and suggested that these biosynthetic genes were downregulated during fruit development; nevertheless, high levels of the transcript encoding the ascorbate oxidase (AO) enzyme were reported [121,122]. Previous research has associated the degradation of AsA content with an increase in ascorbate oxidase and ascorbate peroxidase (APX) activity [124,125]. These results suggest a feedback mechanism in the regulation of AsA in chili pepper fruits [121].
A transcriptional analysis conducted in varieties of chili pepper with contrasting levels of AsA revealed that variations in the concentrations of these metabolites are related to the differential expression of key genes involved in the biosynthesis, degradation and recycling of AsA [126]. Moreover, a recent report indicated that the transcript profiles of L-galactano-1,4-lactone dehydrogenase (GLDH) (encoding an important AsA biosynthetic enzyme) were positively correlated with the CaMYB16 gene, suggesting that CaMYB16 could be a transcriptional regulator of the biosynthesis of AsA in chili pepper fruits (Table 1, Figure 2E) [47].

7. Conclusions

The abundance and diversity of bioactive compounds accumulated during different stages of development and ripening in non-climacteric fruits represent very interesting study models. Nevertheless, transcriptional regulation of the ripening process in non-climacteric fruits has been poorly understood until now. Transcriptional control of genes encoding enzymes that catalyze biosynthetic pathways during ripening has recently been studied in the case of capsaicinoid and anthocyanin production. R2R3 MYB TFs have been identified as the master regulators of capsaicinoid biosynthesis, and a MYB TF interacting with MYC and WD40 TFs has been shown to regulate anthocyanin biosynthesis in Capsicum spp.
Although carotenoids and ascorbic acid are important bioactive compounds with a diversity of nutraceutical applications, the mechanisms underlying the transcriptional regulation of chromoplast biogenesis and carotenoid biosynthesis in chili pepper fruits have not yet been established.
Comprehensive genomic, transcriptomic, and epigenomic analyses provide valuable information about the up- and down-regulation of genes encoding TFs and different structural biosynthetic enzymes. Correlation analyses between the transcriptional profiles of TFs and structural biosynthetic genes and the pattern of accumulation of metabolites across fruit development might be very useful for identifying and selecting candidates for functional gene evaluations. After selecting possible regulatory candidates and performing gene function studies by silencing or editing genes encoding TFs through VIGS and cluster regulatory interspaced palindromic repeats (CRISPR-Cas9), respectively [44,127], Y1H interaction assays could be used to establish the regulatory role of different TFs in the ripening process of chili pepper. Additionally, protein–protein interaction assays (Y2H experiments) should be performed in the future to elucidate the regulatory interactions between different proteins involved in the fruit ripening process. Moreover, ChIP-sequencing (ChIP-seq) has become an important tool for testing the interaction between chromatin and DNA, in order to identify TF targets, and to determinate how TFs could regulate signal networks that affect the plant phenotype [127]. Finally, a novel technic named reverse chromatin immunoprecipitation (R-ChIP) has allowed the characterization of chromatin composition and the isolation of upstream TFs that regulate a particular gene [128]. These experimental alternatives might corroborate the regulatory role of TF candidates on the expression of genes encoding biosynthetic enzymes of chili pepper fruits.

Author Contributions

Conceptualization, M.G.V.-R. and N.O.-A.; writing—original draft preparation, M.G.V.-R. and N.O.-A.; writing—review and editing, M.G.V.-R. and N.O.-A.; visualization, M.G.V.-R.; supervision, N.O.-A.; project administration, N.O.-A.; funding acquisition, N.O.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by grant CB280755 from Conacyt (Mexico).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was funded by the Consejo Nacional de Ciencia y Tecnología (Conacyt), México, project CB280755.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Carrizo García, C.; Barfuss, M.H.; Sehr, E.M.; Barboza, G.E.; Samuel, R.; Moscone, E.A.; Ehrendorfer, F. Phylogenetic relationships, diversification and expansion of chili peppers (Capsicum, Solanaceae). Ann. Bot. 2016, 118, 35–51. [Google Scholar] [CrossRef] [Green Version]
  2. Jarret, R.L.; Barboza, G.E.; Costa Batista, F.R.D.; Berke, T.; Chou, Y.-Y.; Hulse-Kemp, A.; Ochoa-Alejo, N.; Tripodi, P.; Veres, A.; Garcia, C.C.; et al. Capsicum—An Abbreviated Compendium. J. Am. Soc. Hortic. Sci. 2019, 144, 3–22. [Google Scholar] [CrossRef] [Green Version]
  3. Dias, G.B.; Gomes, V.M.; Moraes, T.M.; Zottich, U.P.; Rabelo, G.R.; Carvalho, A.O.; Moulin, M.; Goncalves, L.S.; Rodrigues, R.; Da Cunha, M. Characterization of Capsicum species using anatomical and molecular data. Genet. Mol. Res. 2013, 12, 6488–6501. [Google Scholar] [CrossRef]
  4. Tripodi, P.; Kumar, S. The Capsicum Crop: An introduction. In The Capsicum Genome; Ramchiary, N., Kole, C., Eds.; Springer International Publishing: Basel, Switzaerland, 2019. [Google Scholar] [CrossRef]
  5. Chamikara, M.D.M.; Dissanayake, D.R.R.P.; Ishan, M.; Sooriyapathirana, S.D.S.S. Dietary, anticancer and medicinal properties of the phytochemicals in chili pepper (Capsicum spp.). Ceylon J. Sci. 2016, 45, 5–20. [Google Scholar] [CrossRef] [Green Version]
  6. Batiha, G.E.; Alqahtani, A.; Ojo, O.A.; Shaheen, H.M.; Wasef, L.; Elzeiny, M.; Ismail, M.; Shalaby, M.; Murata, T.; Zaragoza-Bastida, A.; et al. Biological properties, bioactive constituents, and pharmacokinetics of some Capsicum spp. and capsaicinoids. Int. J. Mol. Sci. 2020, 21, 5179. [Google Scholar] [CrossRef] [PubMed]
  7. Cervantes-Hernandez, F.; Alcala-Gonzalez, P.; Martinez, O.; Ordaz-Ortiz, J.J. Placenta, pericarp, and seeds of tabasco chili pepper fruits show a contrasting diversity of bioactive metabolites. Metabolites 2019, 9, 206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Wahyuni, Y.; Ballester, A.-R.; Sudarmonowati, E.; Bino, R.J.; Bovy, A.G. Secondary metabolites of Capsicum species and their importance in the human diet. J. Nat. Prod. 2013, 76, 783–793. [Google Scholar] [CrossRef]
  9. Baenas, N.; Belovic, M.; Ilic, N.; Moreno, D.A.; Garcia-Viguera, C. Industrial use of pepper (Capsicum annum L.) derived products: Technological benefits and biological advantages. Food Chem. 2019, 274, 872–885. [Google Scholar] [CrossRef]
  10. Antonio, A.S.; Wiedemann, L.S.M.; Veiga Junior, V.F. The genus Capsicum: A phytochemical review of bioactive secondary metabolites. RSC Adv. 2018, 8, 25767–25784. [Google Scholar] [CrossRef] [Green Version]
  11. De Sá Mendes, N.; Branco de Andrade Gonçalves, É.C. The role of bioactive components found in peppers. Trends Food Sci. Technol. 2020, 99, 229–243. [Google Scholar] [CrossRef]
  12. Giovannoni, J. Molecular biology of fruit maturation and ripening. Annu. Rev. Plant. Physiol. Plant Mol. Biol. 2001, 52, 725–749. [Google Scholar] [CrossRef] [PubMed]
  13. Giovannoni, J.J. Genetic regulation of fruit development and ripening. Plant Cell 2004, 16, S170–S180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bouzayen, M.; Latché, A.; Nath, P.; Pech, J.C. Mechanism of fruit ripening. In Plant Developmental Biology—Biotechnological Perspectives: Volume 1; Pua, E.C., Davey, M.R., Eds.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 319–339. [Google Scholar] [CrossRef]
  15. Paul, V.; Pandey, R.; Srivastava, G.C. The fading distinctions between classical patterns of ripening in climacteric and non-climacteric fruit and the ubiquity of ethylene-An overview. J. Food Sci. Technol. 2012, 49, 1–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Jang, Y.K.; Jung, E.S.; Lee, H.A.; Choi, D.; Lee, C.H. Metabolomic characterization of hot pepper (Capsicum annuum “CM334”) during fruit development. J. Agric. Food Chem. 2015, 63, 9452–9460. [Google Scholar] [CrossRef] [PubMed]
  17. Li, C.; Hou, X.; Qi, N.; Liu, H.; Li, Y.; Huang, D.; Wang, C.; Liao, W. Insight into ripening-associated transcription factors in tomato: A review. Sci. Hort. 2021, 288, 110363. [Google Scholar] [CrossRef]
  18. Kim, S.; Park, M.; Yeom, S.I.; Kim, Y.M.; Lee, J.M.; Lee, H.A.; Seo, E.; Choi, J.; Cheong, K.; Kim, K.T.; et al. Genome sequence of the hot pepper provides insights into the evolution of pungency in Capsicum species. Nat. Genet. 2014, 46, 270–278. [Google Scholar] [CrossRef]
  19. Lee, S.; Chung, E.J.; Joung, Y.H.; Choi, D. Non-climacteric fruit ripening in pepper: Increased transcription of EIL-like genes normally regulated by ethylene. Funct. Integr. Genomics 2010, 10, 135–146. [Google Scholar] [CrossRef]
  20. Razo-Mendivil, F.G.; Hernandez-Godínez, F.; Hayano-Kanashiro, C.; Martínez, O. Transcriptomic analysis of a wild and a cultivated varieties of Capsicum annuum over fruit development and ripening. PLoS ONE 2021, 16, e0256319. [Google Scholar] [CrossRef]
  21. Villavicencio, L.; Blankenship, S.M.; Sanders, D.C.; Swallow, W.H. Ethylene and carbon dioxide production in detached fruit of selected pepper cultivars. J. Am. Soc. Hortic. Sci. 1999, 124, 402–406. [Google Scholar] [CrossRef] [Green Version]
  22. Aizat, W.M.; Able, J.A.; Stangoulis, J.C.R.; Able, A.J. Proteomic analysis during capsicum ripening reveals differential expression of ACC oxidase isoform 4 and other candidates. Funct. Plant Biol. 2013, 40, 1115–1128. [Google Scholar] [CrossRef] [Green Version]
  23. Aizat, W.M.; Able, J.A.; Stangoulis, J.C.; Able, A.J. Characterisation of ethylene pathway components in non-climacteric Capsicum. BMC Plant Biol. 2013, 13, 191. [Google Scholar] [CrossRef] [Green Version]
  24. Chen, G.; Alexander, L.; Grierson, D. Constitutive expression of EIL-like transcription factor partially restores ripening in the ethylene-insensitive Nr tomato mutant. J. Exp. Bot. 2004, 55, 1491–1497. [Google Scholar] [CrossRef] [Green Version]
  25. Dubey, M.; Jaiswal, V.; Rawoof, A.; Kumar, A.; Nitin, M.; Chhapekar, S.S.; Kumar, N.; Ahmad, I.; Islam, K.; Brahma, V.; et al. Identification of genes involved in fruit development/ripening in Capsicum and development of functional markers. Genomics 2019, 111, 1913–1922. [Google Scholar] [CrossRef] [PubMed]
  26. Vrebalov, J.; Ruezinsky, D.; Padmanabhan, V.; White, R.; Medrano, D.; Drake, R.; Schuch, W.; Giovannoni, J. A MADS-box gene necessary for fruit ripening at the tomato ripening-inhibitor (rin) locus. Science 2002, 296, 343–346. [Google Scholar] [CrossRef] [PubMed]
  27. Vrebalov, J.; Pan, I.L.; Arroyo, A.J.; McQuinn, R.; Chung, M.; Poole, M.; Rose, J.; Seymour, G.; Grandillo, S.; Giovannoni, J.; et al. Fleshy fruit expansion and ripening are regulated by the Tomato SHATTERPROOF gene TAGL1. Plant Cell 2009, 21, 3041–3062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Manning, K.; Tör, M.; Poole, M.; Hong, Y.; Thompson, A.J.; King, G.J.; Giovannoni, J.J.; Seymour, G.B. A naturally occurring epigenetic mutation in a gene encoding an SBP-box transcription factor inhibits tomato fruit ripening. Nat. Genet. 2006, 38, 948–952. [Google Scholar] [CrossRef] [PubMed]
  29. Powell, A.; Nguyen, C.; Hill, T.; Cheng, K.L.; Figueroa-Balderas, R.; Aktas, H.; Ashrafi, H.; Pons, C.; Fernández-Muñoz, R.; Vicente, A.; et al. Uniform ripening encodes a Golden 2-like transcription factor regulating tomato fruit chloroplast development. Science 2012, 336, 1711–1715. [Google Scholar] [CrossRef] [Green Version]
  30. Lin, Z.; Hong, Y.; Yin, M.; Li, C.; Zhang, K.; Grierson, D. A tomato HD-Zip homeobox protein, LeHB-1, plays an important role in floral organogenesis and ripening. Plant J. 2008, 55, 301–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Sung, S.-K.; Moon, Y.-H.; Chung, J.-E.; Lee, S.-Y.; Park, H.G.; An, G. Characterization of MADS box genes from hot pepper. Mol. Cells 2001, 11, 352–359. [Google Scholar] [PubMed]
  32. Ng, M.; Yanofsky, M.F. Function and evolution of the plant MADS-box gene family. Nat. Rev. Genet. 2001, 2, 186–195. [Google Scholar] [CrossRef] [PubMed]
  33. Mizukami, Y.; Huang, H.; Tudor, M.; Hu, Y.; Ma, H. Functional domains of the floral regulator AGAMOUS: Characterization of the DNA binding domain and analysis of dominant negative mutations. Plant Cell 1996, 8, 831–845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Riechmann, J.L.; Krizek, B.A.; Meyerowitz, E.M. Dimerization specificity of Arabidopsis MADS domain homeotic proteins APETALA1, APETALA3, PISTILLATA, and AGAMOUS. Proc. Natl. Acad. Sci. USA 1996, 93, 4793–4798. [Google Scholar] [CrossRef] [Green Version]
  35. Fan, H.-Y.; Hu, Y.; Tudor, M.; Ma, H. Specific interactions between the K domains of AG and AGLs, members of the MADS domain family of DNA binding proteins. Plant J. 1997, 12, 999–1010. [Google Scholar] [CrossRef]
  36. Yang, Y.; Fanning, L.; Jack, T. The K domain mediates heterodimerization of the Arabidopsis floral organ identity proteins, APETALA3 and PISTILLATA. Plant J. 2003, 33, 47–59. [Google Scholar] [CrossRef] [PubMed]
  37. Dong, T.; Chen, G.; Tian, S.; Xie, Q.; Yin, W.; Zhang, Y.; Hu, Z. A non-climacteric fruit gene CaMADS-RIN regulates fruit ripening and ethylene biosynthesis in climacteric fruit. PLoS ONE 2014, 9, e95559. [Google Scholar] [CrossRef] [PubMed]
  38. Yu, C.; Zhan, Y.; Feng, X.; Huang, Z.A.; Sun, C. Identification and expression profiling of the auxin response factors in Capsicum annuum L. under abiotic stress and hormone treatments. Int. J. Mol. Sci. 2017, 18, 2719. [Google Scholar] [CrossRef] [Green Version]
  39. Liu, R.; Song, J.; Liu, S.; Chen, C.; Zhang, S.; Wang, J.; Xiao, Y.; Cao, B.; Lei, J.; Zhu, Z. Genome-wide identification of the Capsicum bHLH transcription factor family: Discovery of a candidate regulator involved in the regulation of species-specific bioactive metabolites. BMC Plant Biol. 2021, 21, 262. [Google Scholar] [CrossRef]
  40. Lu, B.-Y.; Cheng, G.-X.; Zhang, Z.; Sun, J.-T.; Ali, M.; Jia, Q.-L.; Luo, D.-X.; Gong, Z.-H.; Li, D.-W. CaMYC, a novel transcription factor, regulates anthocyanin biosynthesis in color-leaved pepper (Capsicum annuum L.). J. Plant Growth Regul. 2018, 38, 574–585. [Google Scholar] [CrossRef]
  41. Ohme-Takagi, M.; Shinshi, H. Ethylene-inducible DNA binding proteins that interact with an ethylene-responsive element. Plant Cell 1995, 7, 173–182. [Google Scholar] [CrossRef] [Green Version]
  42. Song, J.; Chen, C.; Zhang, S.; Wang, J.; Huang, Z.; Chen, M.; Cao, B.; Zhu, Z.; Lei, J. Systematic analysis of the Capsicum ERF transcription factor family: Identification of regulatory factors involved in the regulation of species-specific metabolites. BMC Genom. 2020, 21, 573. [Google Scholar] [CrossRef]
  43. Jin, H.; Martin, C. Multifunctionality and diversity within the plant MYB-gene family. Plant Mol. Biol. 1999, 41, 577–585. [Google Scholar] [CrossRef] [PubMed]
  44. Arce-Rodriguez, M.L.; Ochoa-Alejo, N. An R2R3-MYB transcription factor regulates capsaicinoid biosynthesis. Plant Physiol. 2017, 174, 1359–1370. [Google Scholar] [CrossRef] [Green Version]
  45. Sun, B.; Zhu, Z.; Chen, C.; Chen, G.; Cao, B.; Chen, C.; Lei, J. Jasmonate-inducible R2R3-MYB transcription factor regulates capsaicinoid biosynthesis and stamen development in Capsicum. J. Agric. Food Chem. 2019, 67, 10891–10903. [Google Scholar] [CrossRef] [PubMed]
  46. Sun, B.; Zhou, X.; Chen, C.; Chen, C.; Chen, K.; Chen, M.; Liu, S.; Chen, G.; Cao, B.; Cao, F.; et al. Coexpression network analysis reveals an MYB transcriptional activator involved in capsaicinoid biosynthesis in hot peppers. Hortic. Res. 2020, 7, 162. [Google Scholar] [CrossRef] [PubMed]
  47. 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]
  48. Aguilar-Barragán, A.; Ochoa-Alejo, N. Virus-induced silencing of MYB and WD40 transcription factor genes affects the accumulation of anthocyanins in chilli pepper fruit. Biol. Plant. 2014, 58, 567–574. [Google Scholar] [CrossRef]
  49. Li, Q.-H.; Yang, S.-P.; Yu, Y.-N.; Khan, A.; Feng, P.-L.; Ali, M.; Shao, D.-K.; Wang, Y.-Y.; Zhang, R.-X.; Gai, W.-X.; et al. Comprehensive transcriptome-based characterization of differentially expressed genes involved in carotenoid biosynthesis of different ripening stages of Capsicum. Sci. Hort. 2021, 288, 110311. [Google Scholar] [CrossRef]
  50. Ulmasov, T.; Hagen, G.; Guilfoyle, T.J. Activation and repression of transcription by auxin-response factors. Proc. Natl. Acad. Sci. USA 1999, 96, 5844–5849. [Google Scholar] [CrossRef] [Green Version]
  51. Guilfoyle, T.J.; Hagen, G. Auxin response factors. Curr. Opin. Plant Biol. 2007, 10, 453–560. [Google Scholar] [CrossRef]
  52. Cheng, Y.; Ahammed, G.J.; Yu, J.; Yao, Z.; Ruan, M.; Ye, Q.; Li, Z.; Wang, R.; Feng, K.; Zhou, G.; et al. Putative WRKYs associated with regulation of fruit ripening revealed by detailed expression analysis of the WRKY gene family in pepper. Sci. Rep. 2016, 6, 39000. [Google Scholar] [CrossRef] [Green Version]
  53. Rushton, P.J.; Somssich, I.E.; Ringler, P.; Shen, Q.J. WRKY transcription factors. Trends Plant Sci. 2010, 15, 247–258. [Google Scholar] [CrossRef]
  54. Zuo, J.; Wang, Y.; Zhu, B.; Luo, Y.; Wang, Q.; Gao, L. Network analysis of noncoding RNAs in pepper provides insights into fruit ripening control. Sci. Rep. 2019, 9, 8734. [Google Scholar] [CrossRef]
  55. Thiele, R.; Mueller-Seitz, E.; Petz, M. Chili pepper fruits: Presumed precursors of fatty acids characteristic for capsaicinoids. J. Agric. Food Chem. 2008, 56, 4219–4224. [Google Scholar] [CrossRef] [PubMed]
  56. Chapa-Oliver, A.M.; Mejia-Teniente, L. Capsaicin: From plants to a cancer-suppressing agent. Molecules 2016, 21, 931. [Google Scholar] [CrossRef] [Green Version]
  57. Naves, E.R.; de Avila Silva, L.; Sulpice, R.; Araujo, W.L.; Nunes-Nesi, A.; Peres, L.E.P.; Zsogon, A. Capsaicinoids: Pungency beyond Capsicum. Trends Plant Sci. 2019, 24, 109–120. [Google Scholar] [CrossRef] [PubMed]
  58. Vázquez-Espinosa, M.; Olguín-Rojas, J.A.; Fayos, O.; González-de-Peredo, A.V.; Espada-Bellido, E.; Ferreiro-González, M.; Barroso, C.G.; Barbero, G.F.; Garcés-Claver, A.; Palma, M. Influence of fruit ripening on the total and individual capsaicinoids and capsiate content in naga jolokia peppers (Capsicum chinense Jacq.). Agronomy 2020, 10, 252. [Google Scholar] [CrossRef] [Green Version]
  59. Arce-Rodriguez, M.L.; Ochoa-Alejo, N. Biochemistry and molecular biology of capsaicinoid biosynthesis: Recent advances and perspectives. Plant Cell Rep. 2019, 38, 1017–1030. [Google Scholar] [CrossRef] [PubMed]
  60. Burgos-Valencia, E.; Echevarría-Machado, I.; Narváez-Zapata, J.A.; Martínez-Estévez, M. Gene expression related to the capsaicinoids biosynthesis in the Capsicum genus: Molecular and transcriptomic studies. Braz. J. Bot. 2020, 43, 201–212. [Google Scholar] [CrossRef]
  61. Tanaka, Y.; Watachi, M.; Nemoto, W.; Goto, T.; Yoshida, Y.; Yasuba, K.I.; Ohno, S.; Doi, M. Capsaicinoid biosynthesis in the pericarp of chili pepper fruits is associated with a placental septum-like transcriptome profile and tissue structure. Plant Cell Rep. 2021, 40, 1859–1874. [Google Scholar] [CrossRef]
  62. Aza-Gonzalez, C.; Nunez-Palenius, H.G.; Ochoa-Alejo, N. Molecular biology of capsaicinoid biosynthesis in chili pepper (Capsicum spp.). Plant Cell Rep. 2011, 30, 695–706. [Google Scholar] [CrossRef]
  63. Zhang, Z.X.; Zhao, S.N.; Liu, G.F.; Huang, Z.M.; Cao, Z.M.; Cheng, S.H.; Lin, S.S. Discovery of putative capsaicin biosynthetic genes by RNA-Seq and digital gene expression analysis of pepper. Sci. Rep. 2016, 6, 34121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Du, H.; Zhang, L.; Liu, L.; Tang, X.F.; Yang, W.J.; Wu, Y.M.; Huang, Y.B.; Tang, Y.X. Biochemical and molecular characterization of plant MYB transcription factor family. Biochemistry 2009, 74, 1–11. [Google Scholar] [CrossRef]
  65. Ambawat, S.; Sharma, P.; Yadav, N.R.; Yadav, R.C. MYB transcription factor genes as regulators for plant responses: An overview. Physiol. Mol. Biol. Plants 2013, 19, 307–321. [Google Scholar] [CrossRef] [Green Version]
  66. Han, K.; Jang, S.; Lee, J.H.; Lee, D.G.; Kwon, J.K.; Kang, B.C. A MYB transcription factor is a candidate to control pungency in Capsicum annuum. Theor. Appl. Genet. 2019, 132, 1235–1246. [Google Scholar] [CrossRef] [PubMed]
  67. Zhu, Z.; Sun, B.; Cai, W.; Zhou, X.; Mao, Y.; Chen, C.; Wei, J.; Cao, B.; Chen, C.; Chen, G.; et al. Natural variations in the MYB transcription factor MYB31 determine the evolution of extremely pungent peppers. New Phytol. 2019, 223, 922–938. [Google Scholar] [CrossRef]
  68. Kondo, F.; Hatakeyama, K.; Sakai, A.; Minami, M.; Nemoto, K.; Matsushima, K. The pungent-variable sweet chili pepper ‘Shishito’ (Capsicum annuum) provides insights regarding the relationship between pungency, the number of seeds, and gene expression involving capsaicinoid biosynthesis. Mol. Genet. Genom. 2021, 296, 591–603. [Google Scholar] [CrossRef] [PubMed]
  69. Wang, J.; Liu, Y.; Tang, B.; Dai, X.; Xie, L.; Liu, F.; Zou, X. Genome-wide identification and capsaicinoid biosynthesis-related expression analysis of the R2R3-MYB gene family in Capsicum annuum L. Front. Genet. 2020, 11, 598183. [Google Scholar] [CrossRef] [PubMed]
  70. Galego, L.; Almeida, J. Role of DIVARICATA in the control of dorsoventral asymmetry in Antirrhinum flowers. Genes Dev. 2002, 16, 880–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Nakano, T.; Suzuki, K.; Fujimura, T.; Shinshi, H. Genome-wide analysis of the ERF Gene Family in Arabidopsis and rice. Plant Physiol. 2006, 140, 411–432. [Google Scholar] [CrossRef] [Green Version]
  72. Huang, P.-Y.; Catinot, J.; Zimmerli, L. Ethylene response factors in Arabidopsis immunity. J. Exp. Bot. 2016, 67, 1231–1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Keyhaninejad, N.; Curry, J.; Romero, J.; O’Connell, M.A. Fruit specific variability in capsaicinoid accumulation and transcription of structural and regulatory genes in Capsicum fruit. Plant Sci. 2014, 215–216, 59–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Toledo-Ortiz, G.; Huq, E.; Quail, P.H. The Arabidopsis Basic/Helix-Loop-Helix transcription factor Family. Plant Cell 2003, 15, 1749–1770. [Google Scholar] [CrossRef] [Green Version]
  75. Guzman, I.; Bosland, P.W.; O’Connell, M.A. Heat, color, and flavor compounds in Capsicum fruit. In The Biological Activity of Phytochemicals. Recent Advances in Phytochemistry; Springer: New York, NY, USA, 2011; Volume 41, pp. 109–126. [Google Scholar] [CrossRef]
  76. Rodriguez-Concepcion, M.; Avalos, J.; Bonet, M.L.; Boronat, A.; Gomez-Gomez, L.; Hornero-Mendez, D.; Limon, M.C.; Meléndez-Martínez, A.J.; Olmedilla-Alonso, B.; Palou, A.; et al. A global perspective on carotenoids: Metabolism, biotechnology, and benefits for nutrition and health. Progr. Lipid Res. 2018, 70, 62–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Rosas-Saavedra, C.; Stange, C. Biosynthesis of carotenoids in plants: Enzymes and color. In Carotenoids in Nature: Biosynthesis, Regulation and Function; Stange, C., Ed.; Springer International Publishing: Cham, Switzerland, 2016; pp. 35–69. [Google Scholar] [CrossRef]
  78. Villa-Rivera, M.G.; Ochoa-Alejo, N. Chili pepper carotenoids: Nutraceutical properties and mechanisms of action. Molecules 2020, 25, 5573. [Google Scholar] [CrossRef]
  79. Wahyuni, Y.; Ballester, A.-R.; Sudarmonowati, E.; Bino, R.J.; Bovy, A.G. Metabolite biodiversity in pepper (Capsicum) fruits of thirty-two diverse accessions: Variation in health-related compounds and implications for breeding. Phytochemistry 2011, 72, 1358–1370. [Google Scholar] [CrossRef] [PubMed]
  80. Mohd Hassan, N.; Yusof, N.A.; Yahaya, A.F.; Mohd Rozali, N.N.; Othman, R. Carotenoids of Capsicum fruits: Pigment profile and health-promoting functional attributes. Antioxidants 2019, 8, 469. [Google Scholar] [CrossRef] [Green Version]
  81. Brand, A.; Borovsky, Y.; Hill, T.; Rahman, K.A.; Bellalou, A.; Van Deynze, A.; Paran, I. CaGLK2 regulates natural variation of chlorophyll content and fruit color in pepper fruit. Theor. Appl. Genet. 2014, 127, 2139–2148. [Google Scholar] [CrossRef]
  82. Borovsky, Y.; Monsonego, N.; Mohan, V.; Shabtai, S.; Kamara, I.; Faigenboim, A.; Hill, T.; Chen, S.; Stoffel, K.; Van Deynze, A.; et al. The zinc-finger transcription factor CcLOL1 controls chloroplast development and immature pepper fruit color in Capsicum chinense and its function is conserved in tomato. Plant J. 2019, 99, 41–55. [Google Scholar] [CrossRef]
  83. Hurtado-Hernandez, H.; Smith, P.G. Inheritance of mature fruit color in Capsicum annuum L. J. Hered. 1985, 76, 211–213. [Google Scholar] [CrossRef]
  84. Jeong, H.B.; Jang, S.J.; Kang, M.Y.; Kim, S.; Kwon, J.K.; Kang, B.C. Candidate gene analysis reveals that the fruit color locus C1 corresponds to PRR2 in pepper (Capsicum frutescens). Front. Plant Sci. 2020, 11, 399. [Google Scholar] [CrossRef]
  85. Pan, Y.; Bradley, G.; Pyke, K.; Ball, G.; Lu, C.; Fray, R.; Marshall, A.; Jayasuta, S.; Baxter, C.; van Wijk, R.; et al. Network inference analysis identifies an APRR2-like gene linked to pigment accumulation in tomato and pepper fruits. Plant Physiol. 2013, 161, 1476–1485. [Google Scholar] [CrossRef] [Green Version]
  86. Makino, S.; Kiba, T.; Imamura, A.; Hanaki, N.; Nakamura, A.; Suzuki, T.; Taniguchi, M.; Ueguchi, C.; Sugiyama, T.; Mizuno, T. Genes encoding pseudo-response regulators: Insight into His-to-Asp phosphorelay and circadian rhythm in Arabidopsis thaliana. Plant Cell Physiol. 2000, 41, 791–803. [Google Scholar] [CrossRef] [Green Version]
  87. Sun, T.; Yuan, H.; Cao, H.; Yazdani, M.; Tadmor, Y.; Li, L. Carotenoid metabolism in plants: The role of plastids. Mol. Plant. 2018, 11, 58–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Camara, B.; Brangeon, J. Carotenoid metabolism during chloroplast to chromoplast transformation in Capsicum annuum fruit. Planta 1981, 151, 359–364. [Google Scholar] [CrossRef] [PubMed]
  89. Hornero-Méndez, D.; Mínguez-Mosquera, M.I. Xanthophyll esterification accompanying carotenoid overaccumulation in chromoplast of Capsicum annuum ripening fruits is a constitutive process and useful for ripeness index. J. Agric. Food Chem. 2000, 48, 1617–1622. [Google Scholar] [CrossRef] [PubMed]
  90. Ytterberg, A.J.; Peltier, J.B.; van Wijk, K.J. Protein profiling of plastoglobules in chloroplasts and chromoplasts. A surprising site for differential accumulation of metabolic enzymes. Plant Physiol. 2006, 140, 984–997. [Google Scholar] [CrossRef] [Green Version]
  91. Siddique, M.A.; Grossmann, J.; Gruissem, W.; Baginsky, S. Proteome analysis of bell pepper (Capsicum annuum L.) Chromoplasts. Plant Cell Physiol. 2006, 47, 1663–1673. [Google Scholar] [CrossRef] [Green Version]
  92. Wang, Y.Q.; Yang, Y.; Fei, Z.; Yuan, H.; Fish, T.; Thannhauser, T.W.; Mazourek, M.; Kochian, L.V.; Wang, X.; Li, L. Proteomic analysis of chromoplasts from six crop species reveals insights into chromoplast function and development. J. Exp. Bot. 2013, 64, 949–961. [Google Scholar] [CrossRef] [Green Version]
  93. Gómez-García, M.D.R.; Ochoa-Alejo, N. Biochemistry and molecular biology of carotenoid biosynthesis in chili peppers (Capsicum spp.). Int. J. Mol. Sci. 2013, 14, 19025–19053. [Google Scholar] [CrossRef] [Green Version]
  94. Sathasivam, R.; Radhakrishnan, R.; Kim, J.K.; Park, S.U. An update on biosynthesis and regulation of carotenoids in plants. S. Afr. J. Bot. 2021, 140, 290–302. [Google Scholar] [CrossRef]
  95. Martínez-López, L.A.; Ochoa-Alejo, N.; Martinez, O. Dynamics of the chili pepper transcriptome during fruit development. BMC Genom. 2014, 15, 143. [Google Scholar] [CrossRef] [Green Version]
  96. Camara, B.; Bardat, F.; Moneger, R. Sites of biosynthesis of carotenoids in Capsicum chromoplasts. Eur. J. Biochem. 1982, 127, 255–258. [Google Scholar] [CrossRef] [PubMed]
  97. Cazzonelli, C.I.; Pogson, B.J. Source to sink: Regulation of carotenoid biosynthesis in plants. Trends Plant Sci. 2010, 15, 266–274. [Google Scholar] [CrossRef]
  98. Stanley, L.; Yuan, Y.W. Transcriptional regulation of carotenoid biosynthesis in plants: So many regulators, so little consensus. Front. Plant Sci. 2019, 10, 1017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Toledo-Ortiz, G.; Huq, E.; Rodriguez-Concepcion, M. Direct regulation of phytoene synthase gene expression and carotenoid biosynthesis by phytochrome-interacting factors. Proc. Natl. Acad. Sci. USA 2010, 107, 11626–11631. [Google Scholar] [CrossRef] [Green Version]
  100. Huq, E.; Al-Sady, B.; Hudson, M.; Kim, C.; Apel, K.; Quail, P.H. Phytochrome-interacting factor 1 is a critical bHLH regulator of chlorophyll biosynthesis. Science 2004, 305, 1937–1941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Bou-Torrent, J.; Toledo-Ortiz, G.; Ortiz-Alcaide, M.; Cifuentes-Esquivel, N.; Halliday, K.J.; Martinez-Garcia, J.F.; Rodriguez-Concepcion, M. Regulation of carotenoid biosynthesis by shade relies on specific subsets of antagonistic transcription factors and cofactors. Plant Physiol. 2015, 169, 1584–1594. [Google Scholar] [CrossRef] [Green Version]
  102. Grassi, S.; Piro, G.; Lee, J.M.; Zheng, Y.; Fei, Z.; Dalessandro, G.; Giovannoni, J.J.; Lenucci, M.S. Comparative genomics reveals candidate carotenoid pathway regulators of ripening watermelon fruit. BMC Genom. 2013, 14, 781. [Google Scholar] [CrossRef] [Green Version]
  103. Hou, B.Z.; Li, C.L.; Han, Y.Y.; Shen, Y.Y. Characterization of the hot pepper (Capsicum frutescens) fruit ripening regulated by ethylene and ABA. BMC Plant Biol. 2018, 18, 162. [Google Scholar] [CrossRef]
  104. Zhu, F.; Luo, T.; Liu, C.; Wang, Y.; Yang, H.; Yang, W.; Zheng, L.; Xiao, X.; Zhang, M.; Xu, R.; et al. An R2R3-MYB transcription factor represses the transformation of alpha- and beta-branch carotenoids by negatively regulating expression of CrBCH2 and CrNCED5 in flavedo of Citrus reticulate. New Phytol. 2017, 216, 178–192. [Google Scholar] [CrossRef] [Green Version]
  105. Lu, S.; Zhang, Y.; Zhu, K.; Yang, W.; Ye, J.; Chai, L.; Xu, Q.; Deng, X. The citrus transcription factor CsMADS6 modulates carotenoid metabolism by directly regulating carotenogenic genes. Plant Physiol. 2018, 176, 2657–2676. [Google Scholar] [CrossRef] [Green Version]
  106. Aza-González, C.; Núñez-Palenius, H.; Ochoa-Alejo, N. Molecular biology of chili pepper anthocyanin biosynthesis. Rev. Soc. Quím. Méx. 2012, 56, 93–98. [Google Scholar] [CrossRef]
  107. Khoo, H.E.; Azlan, A.; Tang, S.T.; Lim, S.M. Anthocyanidins and anthocyanins: Colored pigments as food, pharmaceutical ingredients, and the potential health benefits. Food Nutr. Res. 2017, 61, 1361779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Ramsay, N.A.; Glover, B.J. MYB–bHLH–WD40 protein complex and the evolution of cellular diversity. Trends Plant Sci. 2005, 10, 63–70. [Google Scholar] [CrossRef] [PubMed]
  109. Stommel, J.R.; Lightbourn, G.J.; Winkel, B.S.; Griesbach, R.J. Transcription factor families regulate the anthocyanin biosynthetic pathway in Capsicum annuum. J. Am. Soc. Hortic. Sci. 2009, 134, 244–251. [Google Scholar] [CrossRef] [Green Version]
  110. Smith, T.F.; Gaitatzes, C.; Saxena, K.; Neer, E.J. The WD repeat: A common architecture for diverse functions. Trends Biochem. Sci. 1999, 24, 181–185. [Google Scholar] [CrossRef]
  111. Howard, L.R.; Smith, R.T.; Wagner, A.B.; Villalon, B.; Burns, E.E. Provitamin A and ascorbic acid content of fresh pepper cultivars (Capsicum annuum) and processed jalapeños. J. Food Sci. 1994, 59, 362–365. [Google Scholar] [CrossRef]
  112. Campos, M.R.S.; Gómez, K.R.; Ordoñez, Y.M.; Ancona, D.B. Polyphenols, ascorbic acid and carotenoids contents and antioxidant properties of habanero pepper (Capsicum chinense) fruit. Food Nutr. Sci. 2013, 04, 47–54. [Google Scholar] [CrossRef] [Green Version]
  113. Njus, D.; Kelley, P.M.; Tu, Y.-J.; Schlegel, H.B. Ascorbic acid: The chemistry underlying its antioxidant properties. Free Radic. Biol. Med. 2020, 159, 37–43. [Google Scholar] [CrossRef]
  114. Smirnoff, N. Vitamin C. In Biosynthesis of Vitamins in Plants Part B; Rébeillé, F., Douce, R., Eds.; Academic Press: Cambridge, MA, USA, 2011; Volume 59, pp. 107–177. [Google Scholar] [CrossRef]
  115. Mousavi, S.; Bereswill, S.; Heimesaat, M.M. Immunomodulatory and Antimicrobial Effects of Vitamin C. Eur. J. Microbiol. Immunol. 2019, 9, 73–79. [Google Scholar] [CrossRef]
  116. Chaitanya, N.C.; Muthukrishnan, A.; Krishnaprasad, C.M.S.; Sanjuprasanna, G.; Pillay, P.; Mounika, B. An insight and update on the analgesic properties of vitamin C. J. Pharm. Bioallied Sci. 2018, 10, 119–125. [Google Scholar] [CrossRef]
  117. Colunga Biancatelli, R.M.L.; Berrill, M.; Catravas, J.D.; Marik, P.E. Quercetin and vitamin C: An experimental, synergistic therapy for the prevention and treatment of SARS-CoV-2 related disease (COVID-19). Front. Immunol. 2020, 11, 1451. [Google Scholar] [CrossRef]
  118. Vissers, M.C.M.; Das, A.B. Potential mechanisms of action for vitamin C in cancer: Reviewing the evidence. Front. Physiol. 2018, 9, 809. [Google Scholar] [CrossRef] [Green Version]
  119. Khadi, B.; Goud, J.V.; Patil, V. Variation in ascorbic acid and mineral content in fruits of some varieties of chilli (Capsicum annuum L.). Plant Foods Hum. Nutr. 1987, 37, 9–15. [Google Scholar] [CrossRef]
  120. Kumar, O.; Tata, S. Ascorbic acid contents in chili peppers (Capsicum L.). Not. Sci. Biol. 2009, 1, 50–52. [Google Scholar] [CrossRef] [Green Version]
  121. Alos, E.; Rodrigo, M.J.; Zacarias, L. Transcriptomic analysis of genes involved in the biosynthesis, recycling and degradation of L-ascorbic acid in pepper fruits (Capsicum annuum L.). Plant Sci. 2013, 207, 2–11. [Google Scholar] [CrossRef]
  122. Gómez-García, M.D.R.; 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. 2015, 39, 157–168. [Google Scholar] [CrossRef]
  123. Smirnoff, N. Ascorbic acid metabolism and functions: A comparison of plants and mammals. Free Radic. Biol. Med. 2018, 122, 116–129. [Google Scholar] [CrossRef]
  124. Yahia, E.M.; Contreras-Padilla, M.; Gonazalez-Aguilar, G. Ascorbic acid content in relation to ascorbic acid oxidase activity and polyamine content in tomato and bell pepper fruits during development, maturation and senescence. LWT-Food Sci. Technol. 2001, 34, 452–457. [Google Scholar] [CrossRef] [Green Version]
  125. Barata-Soares, A.D.; Gomez, M.L.P.A.; Mesquita, C.H.D.; Lajolo, F.M. Ascorbic acid biosynthesis: A precursor study on plants. Braz. J. Plant Physiol. 2004, 16, 147–154. [Google Scholar] [CrossRef]
  126. Chiaiese, P.; Corrado, G.; Minutolo, M.; Barone, A.; Errico, A. Transcriptional regulation of ascorbic acid during fruit ripening in pepper (Capsicum annuum) varieties with low and high antioxidants content. Plants 2019, 8, 206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Chen, F.; Hu, Y.; Vannozzi, A.; Wu, K.; Cai, H.; Qin, Y.; Mullis, A.; Lin, Z.; Zhang, L. The WRKY transcription factor family in model plants and crops. Crit. Rev. Plant Sci. 2017, 36, 311–335. [Google Scholar] [CrossRef]
  128. Wen, X.; Wang, J.; Zhang, D.; Ding, Y.; Ji, X.; Tan, Z.; Wang, Y. Reverse Chromatin Immunoprecipitation (R-ChIP) enables investigation of the upstream regulators of plant genes. Commun. Biol. 2020, 3, 770. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Ripening process of Capsicum annuum cv. ‘Tampiqueño 74’ fruits.
Figure 1. Ripening process of Capsicum annuum cv. ‘Tampiqueño 74’ fruits.
Ijms 22 12151 g001
Figure 2. Transcription factors involved in the ripening process of Capsicum spp. (A) Chili pepper fruit development. Days post-anthesis (DPA). (B) Capsaicinoid biosynthetic pathway. Phenylalanine ammonia lyase (PAL), cinnamate 4-hydroxylase (Ca4H), 4-coumaroyl-CoAligase (4CL), hydroxycinnamoyl transferase (HCT), coumaroyl shikimate/quinate3-hydroxylase (C3H), caffeic acid O-methyltransferase (COMT), caffeoyl-CoA 3-O-methyltransferase (HCLH), aminotransferase (pAMT), branched-chain amino acid transferase (BCAT), isovalerate dehydrogenase (BCKDH), ketoacyl-ACPsynthase (KAS), acyl carrier protein (ACL), acyl-ACP thioesterase (FAT), acyl-CoA synthetase (ACS), and capsaicinoid synthase (CS). (C) Carotenoid biosynthetic pathway. Phytoene synthase (PSY), phytoene desaturase (PDS), ζ-carotene isomerase (Z-ISO), ζ-carotene desaturase (ZDS), carotene isomerase (CRTISO), β-lycopene cyclase (LCYB), ε-lycopene cyclase (LCYE), β-carotene hydroxylase (BCH), β-carotene hydroxylase cytochrome 450 type A and B (CYP), zeaxanthin epoxidase (ZEP), violaxanthin epoxidase (VDE), capsanthin-capsorubin synthase (CCS), neoxanthin synthase (NSY), and 9-cis-epoxycarotenoid dioxygenase (NCED). (D) Anthocyanin biosynthetic pathway. Chalcone synthase (CHS), chalcone isomerase (CHI), flavanone 3-hydroxylase (F3H), flavonoid 3′-hydroxylase (F3′H), flavonoid-3′,5′-hydroxylase (F3′5′H), dihydroflavonol 4-reductase (DFR), anthocyanidin synthase (ANS), and UFGT: UDP-Glc-flavonoid 3-O-glucosyl transferase (3GT). (E) L-ascorbic acid biosynthetic pathway. GDP-D-mannose pyrophosphorylase (GMP), GDP-D-mannose-3′,5′-epimerase (GME), GDP-L-galactose phosphorylase (GGP/VTC2), L-galactose-1-dehydrogenase (GalDH), and L-galactano-1,4-lactone dehydrogenase (GLDH).
Figure 2. Transcription factors involved in the ripening process of Capsicum spp. (A) Chili pepper fruit development. Days post-anthesis (DPA). (B) Capsaicinoid biosynthetic pathway. Phenylalanine ammonia lyase (PAL), cinnamate 4-hydroxylase (Ca4H), 4-coumaroyl-CoAligase (4CL), hydroxycinnamoyl transferase (HCT), coumaroyl shikimate/quinate3-hydroxylase (C3H), caffeic acid O-methyltransferase (COMT), caffeoyl-CoA 3-O-methyltransferase (HCLH), aminotransferase (pAMT), branched-chain amino acid transferase (BCAT), isovalerate dehydrogenase (BCKDH), ketoacyl-ACPsynthase (KAS), acyl carrier protein (ACL), acyl-ACP thioesterase (FAT), acyl-CoA synthetase (ACS), and capsaicinoid synthase (CS). (C) Carotenoid biosynthetic pathway. Phytoene synthase (PSY), phytoene desaturase (PDS), ζ-carotene isomerase (Z-ISO), ζ-carotene desaturase (ZDS), carotene isomerase (CRTISO), β-lycopene cyclase (LCYB), ε-lycopene cyclase (LCYE), β-carotene hydroxylase (BCH), β-carotene hydroxylase cytochrome 450 type A and B (CYP), zeaxanthin epoxidase (ZEP), violaxanthin epoxidase (VDE), capsanthin-capsorubin synthase (CCS), neoxanthin synthase (NSY), and 9-cis-epoxycarotenoid dioxygenase (NCED). (D) Anthocyanin biosynthetic pathway. Chalcone synthase (CHS), chalcone isomerase (CHI), flavanone 3-hydroxylase (F3H), flavonoid 3′-hydroxylase (F3′H), flavonoid-3′,5′-hydroxylase (F3′5′H), dihydroflavonol 4-reductase (DFR), anthocyanidin synthase (ANS), and UFGT: UDP-Glc-flavonoid 3-O-glucosyl transferase (3GT). (E) L-ascorbic acid biosynthetic pathway. GDP-D-mannose pyrophosphorylase (GMP), GDP-D-mannose-3′,5′-epimerase (GME), GDP-L-galactose phosphorylase (GGP/VTC2), L-galactose-1-dehydrogenase (GalDH), and L-galactano-1,4-lactone dehydrogenase (GLDH).
Ijms 22 12151 g002
Table 1. Transcription factors involved in the fruit ripening of Capsicum spp.
Table 1. Transcription factors involved in the fruit ripening of Capsicum spp.
Transcription FactorConsensus MotifGeneFunction during RipeningReference
ARF
(Auxin-response factor)
TGTCTC
TGTCCC
TGTCAC
CaARF4,
CaARF5,
CaARF18
Flower and fruit development[38]
bHLH
(basic Helix-Loop-Helix)
E-box
CANNTG
CabHLH007, CabHLH009, CabHLH026, CabHLH063 CabHLH086Capsaicinoid biosynthesis regulation[39]
CabHLH009
CabHLH032
CabHLH048
CabHLH095
CabHLH100
Carotenoid accumulation
CaMYCAnthocyanin biosynthesis[40]
ERFGCC box
TAAGAGCCGCC
CaERF102, CaERF53, CaERF111, CaERF92Capsaicinoid biosynthesis[41,42]
CaERF82, CaERF97, CaERF66, CaERF107, CaERF101Carotenoid biosynthesis[42]
MADS boxCArGCaMADS
CaMADS6
CaMADS-RIN
Ethylene signaling[25,31,37]
MYBTAACTAACCaMYB31
CaMYB48
CaMYB108 CaMYB103
CaMYB115
CaDIV14
Capsaicinoid biosynthesis and accumulation[43,44,45,46,47]
CaMYB3R-5
CaDIV1
Carotenoid biosynthesis
Putative regulator of the CCS gene
[47]
MYBAnthocyanin biosynthesis[48]
CaMYB16Ascorbic acid biosynthesis[47]
F-box protein SKIP23 Carotenoid biosynthesis.
Putative regulators of the CCS gene
[49]
GATA transcription factor 26
U-box domain-containing protein 52
zinc finger family FYVE/PHD-type
RING/FYVE/PHD-type
CONSTANS-LIKE 9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Villa-Rivera, M.G.; Ochoa-Alejo, N. Transcriptional Regulation of Ripening in Chili Pepper Fruits (Capsicum spp.). Int. J. Mol. Sci. 2021, 22, 12151. https://doi.org/10.3390/ijms222212151

AMA Style

Villa-Rivera MG, Ochoa-Alejo N. Transcriptional Regulation of Ripening in Chili Pepper Fruits (Capsicum spp.). International Journal of Molecular Sciences. 2021; 22(22):12151. https://doi.org/10.3390/ijms222212151

Chicago/Turabian Style

Villa-Rivera, Maria Guadalupe, and Neftalí Ochoa-Alejo. 2021. "Transcriptional Regulation of Ripening in Chili Pepper Fruits (Capsicum spp.)" International Journal of Molecular Sciences 22, no. 22: 12151. https://doi.org/10.3390/ijms222212151

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