Next Article in Journal
Identification of 13 Novel Loci in a Genome-Wide Association Study on Taiwanese with Hepatocellular Carcinoma
Next Article in Special Issue
Studying the Functional Potential of Ground Ivy (Glechoma hederacea L.) Extract Using an In Vitro Methodology
Previous Article in Journal
Identification and Expression Analysis of the SKP1-Like Gene Family under Phytohormone and Abiotic Stresses in Apple (Malus domestica)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Environmental Stimuli and Phytohormones in Anthocyanin Biosynthesis: A Comprehensive Review

State Key Laboratory of Plant Environmental Resilience, College of Biological Sciences, China Agricultural University, Beijing 100193, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(22), 16415; https://doi.org/10.3390/ijms242216415
Submission received: 24 October 2023 / Revised: 11 November 2023 / Accepted: 13 November 2023 / Published: 16 November 2023

Abstract

:
Anthocyanin accumulation in plants plays important roles in plant growth and development, as well as the response to environmental stresses. Anthocyanins have antioxidant properties and play an important role in maintaining the reactive oxygen species (ROS) homeostasis in plant cells. Furthermore, anthocyanins also act as a “sunscreen”, reducing the damage caused by ultraviolet radiation under high-light conditions. The biosynthesis of anthocyanin in plants is mainly regulated by an MYB-bHLH-WD40 (MBW) complex. In recent years, many new regulators in different signals involved in anthocyanin biosynthesis were identified. This review focuses on the regulation network mediated by different environmental factors (such as light, salinity, drought, and cold stresses) and phytohormones (such as jasmonate, abscisic acid, salicylic acid, ethylene, brassinosteroid, strigolactone, cytokinin, and auxin). We also discuss the potential application value of anthocyanin in agriculture, horticulture, and the food industry.

1. Introduction

Anthocyanins are a group of flavonoid pigments in plants that produce purple, pink, red, or blue colors, and play important roles in regulating plant development, growth, and the interactions between the plant and the environment [1,2,3,4,5,6,7]. For example, the accumulation of anthocyanins in plants can attract pollinators and seed distributors and can help the plant defend itself against UV-B stress, salinity, drought, and cold stress [8,9,10]. Furthermore, anthocyanin accumulation is associated with fruit veraison and ripening in many crop and fruit plants in parallel with the activation of anthocyanin-synthesis-related enzymes, such as in grape (Vitis vinifera L.), and contributes as an important nutrient for human beings [11,12,13,14,15,16].
The biosynthesis of anthocyanin in plants is mainly controlled by some anthocyanin biosynthetic structural genes, which are divided into two groups: early biosynthetic genes (EBGs, such as CHS, CHI, and F3′H) and late biosynthetic genes (LBGs, such as DFR, LDOX, UF3GT, UGT75C1, and 3AT1) [7,17,18,19]. The EBGs encode the key enzymes to synthesize the precursors common to flavonoids or other phenolics, while LBGs encode the enzymes specifically committed to anthocyanins [18] (Figure 1). For more comprehensive reviews of these enzymes’ functions and the basic biosynthetic pathway of anthocyanins, we refer readers to other reviews [2,18,20,21,22,23].
In plants, the expression of LBGs is mainly controlled by an MYB-bHLH-WD40 (MBW) complex, which consists of MYB transcriptional factors (TFs) (such as GL1, PAP1/MYB75, VvMYBA1, and VvMYBA2), bHLH TFs (such as TT8, GL3, and EGL3), and WD40 protein (such as TTG1) in Arabidopsis and grape (Vitis vinifera L.) [24,25,26]. In recent years, more and more members of the MBW complex in different plants have been identified, and different environmental stimuli or plant hormone signaling pathways can regulate the expression of MBW members, or they can regulate the assembly and the activity of MBW complex to control the anthocyanin biosynthetic gene expressions [18]. To date, many new regulators in different signals involved in MBW-mediated anthocyanin biosynthesis have been identified (Table 1).
In this review, we focus on the regulation network controlling anthocyanin biosynthesis, which is influenced by environmental stimuli and plant hormones. We also delve into the recent discoveries of key regulatory factors within the anthocyanin biosynthesis pathways in plants. Additionally, we underscore the potential applications of anthocyanins in crop breeding.

2. Environmental Stimuli and Anthocyanins

2.1. Light and Anthocyanin Biosynthesis in Plants

Light is crucial in regulating anthocyanin biosynthesis in plants [53]. Without light, the anthocyanin biosynthesis in plants is nearly blocked [27,54,55]. In Arabidopsis, ELONGATED HYPOCOTYL 5 (HY5) and its coactivators BBX20/21/22 play key roles during light-induced anthocyanin biosynthesis [17,56,57]. hy5 and bbx20 21 22 mutants accumulate lower anthocyanin levels compared with wild-type plants [56,58]. HY5 is a key component in light signaling and was shown to act as a transcriptional activator of CHS, CHI, F3′H, MYB12, and PAP1/MYB75 via its direct binding to ACEs (ACGT-containing elements), leading to the accumulation of anthocyanins in response to visible and UV-B light [17,59,60,61,62]. In contrast, VvBBX44 directly represses VvHY5 and VvMYBA1 to balance the anthocyanin concentration under light in grape (Vitis vinifera L.) [28]. Studies showed that in etiolated seedlings, CONSTITUTIVE PHOTOMORPHOGENIC1 (COP1), which is a RING-finger E3 ubiquitin ligase, promotes the polyubiquitination and subsequent degradation of HY5 and HYH (HY5 homolog) in Arabidopsis, which indicates that COP1 acts as a negative regulator of light-induced anthocyanin biosynthesis [17,59,63]. Furthermore, studies also showed that COP1 can directly target PAP1 and PAP2 and promote their degradation in the absence of light [27]. When exposed to light, COP1 is inhibited by activated photoreceptors, such as CRYPTOCHROMEs (CRYs), PHYTOCHROMEs (PHYs), and UV RESISTANCE LOCUS 8 (UVR8), thereby allowing for the accumulation of positively acting TFs [53,64].
Under high-light conditions, genes involved in anthocyanin synthesis are significantly induced, leading to the accumulation of anthocyanins in plants [29,53,65,66]. For instance, high light induces more expressions of MYB112 and PAP1/MYB75, therefore leading to more anthocyanin structural gene expressions [29,65]. It is reported that the MAPK pathway plays an important role in high-light-induced anthocyanin biosynthesis. MAP KINASE4 (MPK4) could be activated in response to light and phosphorylate PAP1/MYB75 to increase the stability of PAP1/MYB75, which is essential for light-induced anthocyanin accumulation [65]. On the other hand, the class II HD-ZIP protein HAT1 negatively regulates the high-light-induced anthocyanin accumulation through competitively interacting with MYB75 and interferes with the formation of the MBW complex, thereby repressing the LBG (such as DFR, LDOX, and UF3GT) expressions via recruiting histone deacetylase mediated by TOPLESS (TPL) [30]. There is a report indicating that under high light treatment, the light attenuation function of anthocyanins is more important than their antioxidant role in photoprotection [66] (Figure 2).

2.2. Salinity Induces Anthocyanin Biosynthesis

Salinity is one of the most widespread abiotic stresses all over the world, and it can induce secondary stress in plants, such as osmotic stress, iron stress, and oxidative stress [67,68]. It was reported that salt stress leads to the accumulation of anthocyanins, which are proposed to be antioxidants that scavenge excessive ROS induced by salinity [8,9,29,31,33,69,70]. pap1-D plants have increased anthocyanin accumulation and radical scavenging activity [71], and they exhibit an enhanced tolerance to high salinity [8]. Truong et al. (2018) reported that enhanced anthocyanin biosynthesis leads to better growth performance of plants under low-nitrate and high-salinity conditions via the regulation of nitrate metabolism [72]. Moreover, this group found that increasing the amount of anthocyanins by knocking out FLS1 in pap1-D mutant could improve the salt stress tolerance under high NO3- application [73]. Moreover, ectopic expression of AtDFR leads to a high level of anthocyanin accumulation and confers significant salinity tolerance in Brassica napus L. [70]. Furthermore, Arabidopsis UDP-glycosyltransferases UGT79B2 and UGT79B3 were shown to improve salt tolerance via the modulation of anthocyanin accumulation [9].
Salinity induces the expressions of many anthocyanin-biosynthesis-related genes, including both EBGs (CHS, CHI, F3′H) and LBGs (DFR, LDOX, UF3GT, UGT75C1, 3AT1, etc.) in plants [29,33]. To date, several transcriptional regulators are identified as key factors that regulate salt-induced structural genes. PAP1/MYB75 is the most known TF that upregulates salt-induced anthocyanins, where the salt-induced anthocyanins could be enhanced by adding sucrose, which further indicates that salt signaling may engage in cross-talk with sucrose signaling [8,33]. AtMYB112 is another TF that positively regulates salinity-induced anthocyanins. Salt stress (150 mM NaCl) can induce the expression level of AtMYB112 and lead to the upregulation of downstream genes (AtMYB7 and AtMYB32), hence promoting salt-induced anthocyanin biosynthesis in Arabidopsis [29]. Recently, research into Arabidopsis indicates that the adaptor protein EAR motif-Containing Adaptor Protein (ECAP) interacts with PAP1/MYB75 and represses its activity in normal conditions, while different levels of salinity can remove the ECAP’s repression of the PAP1/MYB75-dependent MBW complex by both jasmonate (JA) signaling and an unknown pathway (Figure 2) [33].

2.3. Drought Promotes Anthocyanin Accumulation in Plants

Drought is one of the most common environmental stresses that plants suffer [74,75,76,77]. Lots of studies found that drought promotes anthocyanin accumulation in plants, and the accumulated anthocyanins serve as an important antioxidant to scavenge ROS induced by drought stress [9,78]. Recently, there was a study showing that the ectopic overexpression of StAN1 (a key TF that regulates ANS, DFR, and UFGT from Solanum tuberosum) in tobacco plants leads to the overaccumulation of anthocyanins, and the transgenic plants have a stronger drought tolerance compared with wild-type plants [79] (Figure 2). A study of Arabidopsis indicated that different abiotic-stress-induced anthocyanins have different localizations at the tissue and organ levels [80]. ECAP was shown to mediate the drought-induced anthocyanin accumulation in Arabidopsis [31]; however, whether this process is dependent on JA signaling remains unclear.

2.4. Cold Induces Anthocyanins Accumulation in Plants

Cold stress, which is a common abiotic factor, exerts significant impacts on plant growth, development, and overall fitness [81]. Plants have evolved complex molecular responses to counteract the detrimental effects of cold stress, and one intriguing aspect of this response is the induction of anthocyanin synthesis [4,82]. Cold stress triggers the upregulation of specific transcription factors, such as MYBs (such as BrMYB2 and AtMYB75) and bHLHs (such as BrTT8), which bind to the promoter regions of anthocyanin biosynthetic genes (such as BrDFR1, BrANS1, and BrUF3GT2), initiating the transcriptional cascade that leads to anthocyanin production [83] (Figure 2). Anthocyanins possess potent antioxidant properties to scavenge ROS generated during cold stress [84]. ROS accumulation can result in cellular damage, membrane disruption, and oxidative stress. By effectively neutralizing ROS, anthocyanins contribute to the maintenance of cellular integrity and homeostasis, reducing the potential for cold-induced damage [84,85]. In addition to ROS scavenging, anthocyanins may also play a role in photoprotection [86]. Cold stress often leads to photoinhibition due to the deregulation of photosynthesis homeostasis. Anthocyanins can act as “sunscreen” pigments, absorbing excess light energy and dissipating it as heat, thereby protecting the photosynthetic apparatus from photodamage [87,88]. In apple, MdMYB308L serves as a positive regulator of cold tolerance and anthocyanin accumulation through its interaction with MdbHLH33 and undergoes protein degradation mediated by MdMIEL1, highlighting the pivotal role of dynamic MYB-bHLH protein complexes in plant growth and development regulation [34]. In grape (Vitis vinifera L.), anthocyanin accumulation in leaves induced by the low temperature in autumn can help to enhance their cold tolerance [89].

2.5. Anthocyanins Confer Pest and Disease Resistance in Plants

Recent studies suggest that anthocyanins may contribute to plant defense against pests and diseases [4,10,90,91,92] (Figure 2). For instance, the content of anthocyanins was significantly increased in rust-infected symptomatic tissue of Malus apple, and the anthocyanin biosynthetic genes McDFR and McLOX were also upregulated [91]. Another study showed that some flavonoid glycosides in Basella alba could inhibit the growth of Spodoptera litura larvae [92]. However, the molecular mechanisms of pest- or disease-mediating anthocyanins accumulation still need further investigation.

2.6. Nutrient-Limitation-Induced Anthocyanin Accumulation in Plants

Low phosphorus and low nitrogen stresses often induce the accumulation of anthocyanins in plants. Under conditions of phosphorus and nitrogen limitation, plants adapt to environmental stress by modulating nutrient allocation and metabolic pathways, including increasing the synthesis of anthocyanins [35,36,93] (Figure 2). This phenomenon is known as nutrient-limitation-induced anthocyanin accumulation [35,94,95]. This represents a physiological response strategy of plants to environmental stress that is aimed at enhancing their resilience.
Phosphorus is a vital component in energy transfer and molecular signaling, and thus, its scarcity prompts plants to allocate resources strategically. In response, plants regulate anthocyanin biosynthesis as a part of a larger mechanism to enhance their adaptive fitness [36,96]. Under low-phosphorus conditions, increased expression of MYB transcription factors, such as PHOSPHATE STARVATION RESPONSE 1 (PHR1), was observed [39,40,97]. PHR1 can initiate a signaling cascade that directly upregulates anthocyanin-related genes, such as F3′H and LDOX in Arabidopsis [40]. Furthermore, SPX4 also controls the PAP1 protein level and affects the PAP1-mediated anthocyanin pathway under low-phosphorus stress [39]. Notably, the Gibberellin (GA)-DELLA signaling pathway also regulates the phosphate-starvation-induced anthocyanin in Arabidopsis [36].
Similarly, low-nitrogen stress triggers a sophisticated interplay of molecular mechanisms that culminate in the induction of anthocyanin accumulation. Nitrogen, as an essential component of amino acids and proteins, plays a central role in plant growth and development. In response to nitrogen scarcity, plants redistribute resources to favor metabolic pathways that improve nutrient efficiency [98,99]. This reallocation often coincides with the accumulation of anthocyanins, as observed in the model plant Arabidopsis thaliana. NLA (nitrogen limitation adaptation) plays a key role in regulating N-limitation-induced anthocyanin synthesis. The nla mutant cannot accumulate anthocyanins and instead produces an N-limitation-induced early senescence phenotype [35]. The MYB TFs also play a key role in low-N-induced anthocyanin accumulation, and the PAP1 loss-of-function mutant showed low anthocyanin accumulation and low survival rate under a low-N stress treatment [93,100]. Furthermore, a DFR-deficient mutant tt3 also showed significantly lower survival rates after N starvation compared with the wild type in Arabidopsis (Figure 2). These studies all indicate that low-N-induced anthocyanin accumulation plays a substantial role in plant tolerance to low-N stress. Moreover, studies also showed that the GA-DELLA module is involved in nitrogen-deficiency-induced anthocyanin accumulation. DELLAs could interact with PAP1 and enhance the transcriptional activity of PAP1 to promote the expressions of F3′H and DFR [37,101].

3. Plant Hormones and Anthocyanins

3.1. Strigolactone Promotes Anthocyanin Accumulation

Strigolactone (SL) is an important phytohormone that participates in regulating shoot branching, leaf shape, and metabolism in plants [41]. A report indicates that SL can promote the accumulation of anthocyanin, which further confers adaption to a low-phosphate condition in Arabidopsis [102]. Wang et al. 2020 found that the SL-mediated anthocyanin biosynthesis was triggered by the degradation of SMXL6, SMXL7, and SMXL8 proteins through the 26S proteasome pathway, further promoting the expressions of PAP1/PAP2 in Arabidopsis. The SMXLs are the key repressors and have dual functions in SL signaling. On one hand, SMXL6 can directly bind the promoters of SMXL6/7/8 and negatively regulate their transcription; on the other hand, SMXL6 can function as a transcriptional repressor to inhibit the expressions of SL-responsive genes, including PAP1/PAP2 [41] (Figure 3). However, further investigation is still needed to determine which TFs directly bind to SMXLs and function upstream of PAP1/PAP2.

3.2. JA Mediates Anthocyanin Biosynthesis

JA is a plant hormone that participates in plant defense against biotic/abiotic stress and regulates plant metabolisms [18,103]. JA was shown to have a positive effect on anthocyanin biosynthesis in plants [32,42,104]. Studies showed that multiple members of JA signaling in plants are involved in anthocyanin biosynthesis. The mutants of JA receptor gene CORONATINE INSENSITIVE1 (COI1) and JA biosynthetic gene 12-oxophytodienoate reductase 3 (OPR3) show a low anthocyanin phenotype compared with wild-type Arabidopsis seedlings [31,32,33,42]; while the key repressors of JA signaling, namely, JAZs, are negative regulators of anthocyanin accumulation [31,32,105], and ECAP acts as an adaptor protein mediating the interacting of JAZ6/8 and co-repressor TPR2 to form the JET complex and plays a negative role in anthocyanin biosynthesis in Arabidopsis [31]. It has become clear that JA regulates anthocyanin biosynthesis by affecting the stability and activity of the MBW complex [31,32]. JAZ1/8/11 can competitively inhibit the formation of the MBW complex and hence inhibit the expressions of LBGs [32]. ECAP-mediated repression is mainly achieved via histone deacetylation on the target genes of the MBW complex rather than by competitively binding with MBW members [31], while JA promotes the degradation of JAZs and allows for the activity of the MBW complex, thereby leading to the upregulation of LBGs [31,32]. In apple, a study showed that the JAZ1-TRB1-MYB9 complex dynamically modulates the JA-mediated accumulation of both anthocyanin and proanthocyanidin [43] (Figure 3). Furthermore, in grape (Vitis vinifera L.), methyl jasmonate (MeJA) treatment promotes anthocyanin accumulation by regulating a VvmIR156-VvSPL9 module in the early stage of color conversion [44].

3.3. Abscisic Acid Promotes Anthocyanin Accumulation

Abscisic acid (ABA) is a plant hormone that regulates plant growth, development, and stress response [106,107]. There are also many reports indicating that ABA mediates development-dependent anthocyanin biosynthesis in the leaves and fruits of many plants, such as apple, strawberry, sweet berry, bilberry, and lycium plants [45,108,109,110,111,112]. The key TF of ABA signaling in apple, namely, MdABI5, was found to positively regulate ABA-induced anthocyanin biosynthesis by directly upregulating the expression of MdbHLH3 and interacting with MdbHLH3 to enhance the MdMYB1–MdbHLH3 interaction, which led to the upregulation of MdDFR and MdUF3GT transcripts [45]. Elongators also play important roles in regulating ABA signaling and anthocyanin biosynthesis. The Elongator subunit (ELO1/ELP4, ELP2, and ELP6) mutants in Arabidopsis are hypersensitive to ABA and accumulate more anthocyanins than the wild type [46]. Further investigation shows that Elongator positively regulates the expression of MYBL2, which is a negative regulator of the MBW complex [46] (Figure 3).

3.4. Salicylic Acid Mediates Anthocyanin Biosynthesis

Salicylic acid (SA) is a vital plant hormone that regulates immunity against biotrophic and semi-biotrophic pathogens [113]. SA also positively regulates anthocyanin accumulation in many plants, such as grape, pomegranate, and Arabidopsis [48,114,115]. It shows that knocking out the SA receptor NPR1 leads to a low-anthocyanin-content phenotype in Arabidopsis compared with the wild type, regardless of the airborne fungus treatment [48]. Furthermore, the MBW complex also participates in airborne-fungus-induced anthocyanin biosynthesis [48]. This indicates that SA signaling mediates airborne-fungus-induced anthocyanin accumulation, though the key components of the TFs or other co-regulators are still not very clear.

3.5. Brassinosteroid and Anthocyanin Biosynthesis

Brassinosteroid (BR) is a new class of plant hormone that participates in many physiological processes, including anthocyanin biosynthesis, in many plants [49,116,117,118]. In grape, the exogenous application of BR and its analogs led to an increase in anthocyanin accumulation in its fruit [119]. However, a study of apple showed that BR treatment inhibits the synthesis of anthocyanin, and the MdBZR1-MdJa2 module plays a negative role in the control of downstream LBG expressions [49] (Figure 3). Furthermore, studies of Arabidopsis also indicated that the BR biosynthetic mutant det2 accumulates more anthocyanins than the wild type [116,120], indicating that BR acts as a negative regulator of anthocyanin biosynthesis. BR may cross talk with JA and CK signals to regulate anthocyanin biosynthesis [121,122]. It was indicated that JA-induced anthocyanin accumulation was repressed in BR mutants or the wild type treated with brassinazole, which is an inhibitor of BR biosynthesis, whereas it was induced via an application of exogenous BR. Further study showed that BR affects JA-induced anthocyanin accumulation by regulating the LBGs, and this regulation might be mediated by the WD-repeat/MYB/bHLH transcriptional complex [121].

3.6. Cytokinin Mediates Anthocyanin Biosynthesis

Cytokinin (CK) is an important plant hormone that controls plant organ formation, seed germination, senescence, etc. [123]. There was a study that showed that exogenous CK treatment promotes Arabidopsis accumulating more anthocyanin pigments [124], which indicates that CK may play a positive role in regulating plant anthocyanin biosynthesis. CK enhances sucrose-mediated anthocyanin pigmentation, and the CK sensors (AHK2/3/4), histidine-containing phosphotransfer proteins (AHP2/3/5), and master TFs (type B ARR1/10/12) in CK signaling mediate this process in Arabidopsis [1,125] (Figure 3). However, CK seems to play a negative role during the high-salinity-induced anthocyanin accumulation. It is reported that the CK-signaling-defective mutants ahp2,3,5 and arr1,10,12 triple mutants show more anthocyanin accumulations after a high-salinity treatment [126]. This indicates that CK-mediated anthocyanin biosynthesis is very complex and CK may cross talk with other signals to regulate anthocyanin biosynthesis and cope with the change in environment, such as high salinity. And further investigation should focus on the mechanism of how CK interacts with other signals to regulate anthocyanin biosynthesis under a stress environment.

3.7. Ethylene and Anthocyanin Biosynthesis

Ethylene participates in many biological processes, such as plant growth, senescence, fruit ripening, and stress responses [127,128]. There are also reports showing that ethylene negatively regulates anthocyanin pigmentation in many plants [47,129,130]. It was found that ethylene inhibited sucrose- and light-induced-anthocyanin accumulation in Arabidopsis [47,129,131]. The mutants of key components in Arabidopsis ethylene signaling, such as etr1-1, ein2-1, and ein3 eil1, all showed ethylene-insensitive and enhanced anthocyanin accumulation phenotypes and further investigation showed that ethylene represses anthocyanin biosynthesis by upregulating the expression of the negative TF MYBL2 while downregulating the expression of positive TFs, such as MYB75, GL3, and TT8 [47] (Figure 3). In tomato, exogenous ethylene treatment significantly repressed anthocyanin accumulation and the expression of SlAN2-like and other anthocyanin-related genes [130]. A recent study indicates an ethylene-responsive transcription factor PpERF9 inhibits anthocyanin biosynthesis through epigenetic repression of PpRAP2.4 and PpMYB114 via histone deacetylation in pear [50] (Figure 3). However, ethylene can promote anthocyanin accumulation in certain fruits by upregulating genes related to anthocyanin synthesis or increasing the activity of enzymes involved in anthocyanin metabolism. This was observed in fruits such as plum (Prunus spp.), grape (Vitis vinifera L.), and strawberry (Fragaria × ananassa) [132,133,134]. Furthermore, when dark-grown sorghum seedlings, which were treated with ethylene, were subsequently exposed to light, the anthocyanin levels increased compared with those without treatment [131]. In summary, the relationship between ethylene and anthocyanin is complex and warrants further exploration in future studies.

3.8. GA Negatively Regulates Anthocyanin Biosynthesis

GA is one of the important plant hormones that regulate a diverse range of processes associated with plant growth and development [135]. There are many studies that showed that GA acts as a negative regulator in plant anthocyanin biosynthesis [37]. For instance, a study showed that exogenous application of GA treatment could significantly reduce anthocyanin accumulation in Arabidopsis wild-type seedlings [38]. GA also negatively regulates low-temperature-induced anthocyanin accumulation in a HY5/HYH-dependent manner [136]. Zhang et al. 2017 found that a DELLA protein, namely, RGA, can strongly interact with PAP1/MYB75 and enhance its transcriptional activity in Arabidopsis, thereby leading to the upregulation of the LBGs under a nitrogen deficiency condition [37]. GA signaling may engage in cross-talk with ABA and JA signaling-mediated anthocyanin biosynthesis, where it was shown that DELLA proteins can promote anthocyanin biosynthesis through sequestering MYBL2 and JAZ suppressors of the MBW complex in Arabidopsis [38] (Figure 3).

3.9. Auxin and Anthocyanin Biosynthesis

Auxin is an important phytohormone that governs plant growth, development, and responses to environmental variations [137,138,139]. Early studies found that exogenous indole acetic acid (IAA) represses Sorghum and Brassica anthocyanin accumulation in a dose-dependent manner [140,141], indicating that auxin may play a negative role in regulating anthocyanin biosynthesis in plants. Later studies in many plants confirmed that high auxin can repress plant anthocyanin accumulation [51,52,142,143]. Recently, studies in apple showed that auxin inhibits anthocyanin biosynthesis through the Aux/IAA-ARF signaling pathway [51]. On one hand, a TF MdARF13 binds the promoter of MdDFR and inhibits its transcription; on the other hand, MdARF13 also destabilizes the MBW complex by competitively interacting with MdMYB10, which is a key member of the MBW complex in apple. When the auxin level is low, the auxin/indole-3-acetic acid (Aux/IAA) repressor MdIAA121 binds MdARF13 and restrains it from directly binding the MdDFR promoter or interacting with MdMYB10 [51] (Figure 3). Furthermore, another study in apple also found that MdIAA26 acts as a positive regulator that promotes anthocyanin accumulation, while auxin promotes the degradation of MdIAA26. However, further investigation is still needed to determine which MdARF is the target of MdIAA26 [52]. Recently, there was a study in sweet cherry that showed that the synthetic auxin 1-naphthaleneacetic acid (NAA) treatment enhances the anthocyanin pigments during the straw-color stage of fruit development, probably by regulating ethylene and ABA metabolism [144]. This indicates that auxin may cross talk with other signals to regulate anthocyanin biosynthesis, and further studies should focus on the mechanisms of how auxin interacts with other signals to control anthocyanin homeostasis in plants.

4. Unveiling the Future: Anthocyanins’ Revolutionary Role in Agriculture, Food, and Horticulture

Anthocyanins, which comprise a class of natural pigments responsible for the vibrant hues of various fruits, vegetables, and flowers, have gained significant attention due to their potential health benefits [145,146]. As scientific studies continue to unravel the multifaceted properties of anthocyanins, their applications in agriculture and food processing are emerging as promising avenues for enhancing visual appeal, nutritional content, and overall consumer satisfaction.

4.1. Crop Color Enhancement

Synthetic food colorants have raised concerns regarding their safety and impact on health. Anthocyanins offer a natural alternative to food coloring, enabling food processors to meet consumer demand for visually appealing and safe products without compromising health [147]. Anthocyanin-rich foods not only add vibrant colors but also contribute to nutritional enhancement. Incorporating these pigments into a variety of processed foods can elevate their antioxidant and phytonutrient contents, thereby improving the overall nutritional profile. Anthocyanin-rich crops have the potential to revolutionize the esthetics of agricultural landscapes. Manipulating anthocyanin biosynthesis through genetic engineering or selective breeding can result in visually appealing crops, thereby increasing consumer interest and market value (Figure 4).

4.2. Stress-Tolerant Crop Breeding

Anthocyanins have been linked to enhanced stress tolerance in plants, including resistance to various abiotic stresses and biotic stresses [2,4,93]. Incorporating these traits into crops can lead to improved resilience, increased yield stability, and sustainable agricultural practices, which may provide significant guidance to crop breeding and improvement (Figure 4).

4.3. Ornamental Plant Innovations

The utilization of anthocyanins can extend beyond edibles to ornamental plants [148]. Developing new cultivars with vibrant colors and prolonged bloom periods can significantly enhance the ornamental horticulture industry. Furthermore, anthocyanin-rich plants can contribute to urban greening initiatives to beautify cityscapes while also providing ecosystem services, such as air purification and temperature regulation (Figure 4).
The growing body of research on anthocyanins’ health benefits and diverse applications has sparked interest across the agricultural and food industries. From enhancing crop aesthetics to improving nutritional content and contributing to sustainable agricultural practices, anthocyanins hold immense promise. As scientific knowledge advances and consumer preferences shift toward natural and healthier options, the integration of anthocyanins in agriculture, food processing, and horticulture is poised to play a pivotal role in shaping the future of these industries. Thus, identifying additional regulatory and structural genes controlling anthocyanin biosynthesis in various plants and deciphering their regulatory networks holds significant scientific importance for the future molecular breeding of crops and horticultural species.

Author Contributions

Y.F. and C.L. proposed the concept and content. C.L., L.S. and X.L. drafted the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This project was funded by the National Natural Science Foundation of China (grant no. 32230030 to Y.F. and grant no. 32200261 to C.L.) and the China Postdoctoral Science Foundation grant (2018M641532 to C.L.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We apologize to the many authors whose work we were not able to cite here due to the limited space.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Das, P.K.; Shin, D.H.; Choi, S.B.; Park, Y.I. Sugar-hormone cross-talk in anthocyanin biosynthesis. Mol. Cells 2012, 34, 501–507. [Google Scholar] [CrossRef]
  2. Li, Z.; Ahammed, G.J. Plant stress response and adaptation via anthocyanins: A review. Plant Stress 2023, 10, 100230. [Google Scholar] [CrossRef]
  3. Li, Z.; Ahammed, G.J. Hormonal regulation of anthocyanin biosynthesis for improved stress tolerance in plants. Plant Physiol. Biochem. 2023, 201, 107835. [Google Scholar] [CrossRef] [PubMed]
  4. Kaur, S.; Tiwari, V.; Kumari, A.; Chaudhary, E.; Sharma, A.; Ali, U.; Garg, M. Protective and defensive role of anthocyanins under plant abiotic and biotic stresses: An emerging application in sustainable agriculture. J. Biotechnol. 2023, 361, 12–29. [Google Scholar] [CrossRef] [PubMed]
  5. Vidana Gamage, G.C.; Lim, Y.Y.; Choo, W.S. Anthocyanins from Clitoria ternatea flower: Biosynthesis, extraction, stability, antioxidant activity, and applications. Front. Plant Sci. 2021, 12, 792303. [Google Scholar] [CrossRef] [PubMed]
  6. Choo, W.S. Fruit pigment changes during ripening. In Encyclopedia of Food Chemistry; Melton, L., Shahidi, F., Varelis, P., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 117–123. ISBN 9780128140451. [Google Scholar] [CrossRef]
  7. He, F.; Mu, L.; Yan, G.L.; Liang, N.N.; Pan, Q.H.; Wang, J.; Reeves, M.J.; Duan, C.Q. Biosynthesis of anthocyanins and their regulation in colored grapes. Molecules 2010, 15, 9057. [Google Scholar] [CrossRef] [PubMed]
  8. Oh, J.E. Enhanced level of anthocyanin leads to increased salt tolerance in Arabidopsis PAP1-D plants upon sucrose treatment. J. Korean Soc. Appl. Biol. Chem. 2011, 54, 79–88. [Google Scholar] [CrossRef]
  9. Li, P.; Li, Y.J.; Zhang, F.J.; Zhang, G.Z.; Jiang, X.Y.; Yu, H.M.; Hou, B.K. The Arabidopsis UDP-glycosyltransferases UGT79B2 and UGT79B3, contribute to cold, salt and drought stress tolerance via modulating anthocyanin accumulation. Plant J. 2016, 89, 85–103. [Google Scholar] [CrossRef]
  10. Lev-Yadun, S.; Gould, K.S. Role of anthocyanins in plant defence. In Anthocyanins; Winefield, C., Davies, K., Gould, K., Eds.; Springer: New York, NY, USA, 2008; pp. 22–28. ISBN 978-0-387-77334-6. [Google Scholar] [CrossRef]
  11. Li, W.F.; Mao, J.; Yang, S.J.; Guo, Z.G.; Ma, Z.H.; Dawuda, M.M.; Zuo, C.W.; Chu, M.Y.; Chen, B.H. Anthocyanin accumulation correlates with hormones in the fruit skin of ‘Red Delicious’ and its four generation bud sport mutants. BMC Plant Biol. 2018, 18, 363. [Google Scholar] [CrossRef]
  12. Yan, S.; Chen, N.; Huang, Z.; Li, D.; Zhi, J.; Yu, B.; Liu, X.; Cao, B.; Qiu, Z. Anthocyanin fruit encodes an R2R3-MYB transcription factor, SlAN2-like, activating the transcription of SlMYBATV to fine-tune anthocyanin content in tomato fruit. New Phytol. 2019, 225, 2048–2063. [Google Scholar] [CrossRef]
  13. Sun, C.; Deng, L.; Du, M.; Zhao, J.; Chen, Q.; Huang, T.; Jiang, H.; Li, C.B.; Li, C. A transcriptional network promotes anthocyanin biosynthesis in tomato flesh. Mol. Plant 2020, 13, 42–58. [Google Scholar] [CrossRef] [PubMed]
  14. Butelli, E.; Titta, L.; Giorgio, M.; Mock, H.-P.; Matros, A.; Peterek, S.; Schijlen, E.G.W.M.; Hall, R.D.; Bovy, A.G.; Luo, J.; et al. Enrichment of tomato fruit with health-promoting anthocyanins by expression of select transcription factors. Nat. Biotechnol. 2008, 26, 1301–1308. [Google Scholar] [CrossRef] [PubMed]
  15. Salama, A.M.; Abdelsalam, M.A.; Rehan, M.; Elansary, M.; El-Shereif, A. Anthocyanin accumulation and its corresponding gene expression, total phenol, antioxidant capacity, and fruit quality of ‘Crimson Seedless’ Grapevine (Vitis vinifera L.) in response to grafting and pre-harvest applications. Horticulturae 2023, 9, 1001. [Google Scholar] [CrossRef]
  16. Kuhn, N.; Guan, L.; Dai, Z.W.; Wu, B.H.; Lauvergeat, V.; Gomès, E.; Li, S.H.; Godoy, F.; Arce-Johnson, P.; Delrot, S. Berry ripening: Recently heard through the grapevine. J. Exp. Bot. 2014, 65, 4543–4559. [Google Scholar] [CrossRef]
  17. Shin, D.H.; Choi, M.; Kim, K.; Bang, G.; Cho, M.; Choi, S.B.; Choi, G.; Park, Y.I. HY5 regulates anthocyanin biosynthesis by inducing the transcriptional activation of the MYB75/PAP1 transcription factor in Arabidopsis. FEBS Lett. 2013, 587, 1543–1547. [Google Scholar] [CrossRef]
  18. Lacchini, E.; Goossens, A. Combinatorial control of plant specialized metabolism: Mechanisms, functions, and consequences. Annu. Rev. Cell. Dev. Biol. 2020, 36, 291–313. [Google Scholar] [CrossRef]
  19. Li, C.; Yu, W.; Xu, J.; Lu, X.; Liu, Y. Anthocyanin biosynthesis induced by MYB transcription factors in plants. Int. J. Mol. Sci. 2022, 23, 11701. [Google Scholar] [CrossRef]
  20. Yan, H.; Pei, X.; Zhang, H.; Li, X.; Zhang, X.; Zhao, M.; Chiang, V.L.; Sederoff, R.R.; Zhao, X. MYB-mediated regulation of anthocyanin biosynthesis. Int. J. Mol. Sci. 2021, 22, 3103. [Google Scholar] [CrossRef]
  21. Saigo, T.; Wang, T.; Watanabe, M.; Tohge, T. Diversity of anthocyanin and proanthocyanin biosynthesis in land plants. Curr. Opin. Plant Biol. 2020, 55, 93–99. [Google Scholar] [CrossRef]
  22. Mouradov, A.; Spangenberg, G. Flavonoids: A metabolic network mediating plants adaptation to their real estate. Front. Plant Sci. 2014, 5, 620. [Google Scholar] [CrossRef]
  23. Lacchini, E.; Venegas-Molina, J.; Goossens, A. Structural and functional diversity in plant specialized metabolism signals and products: The case of oxylipins and triterpenes. Curr. Opin. Plant Biol. 2023, 74, 102371. [Google Scholar] [CrossRef] [PubMed]
  24. Xu, W.; Grain, D.; Bobet, S.; Le Gourrierec, J.; Thevenin, J.; Kelemen, Z.; Lepiniec, L.; Dubos, C. Complexity and robustness of the flavonoid transcriptional regulatory network revealed by comprehensive analyses of MYB-bHLH-WDR complexes and their targets in Arabidopsis seed. New Phytol. 2014, 202, 132–144. [Google Scholar] [CrossRef] [PubMed]
  25. Jiu, S.; Guan, L.; Leng, X.; Zhang, K.; Haider, M.S.; Yu, X.; Zhu, X.; Zheng, T.; Ge, M.; Wang, C.; et al. The role of VvMYBA2r and VvMYBA2w alleles of the MYBA2 locus in the regulation of anthocyanin biosynthesis for molecular breeding of grape (Vitis spp.) skin coloration. Plant Biotechnol. J. 2021, 19, 1216–1239. [Google Scholar] [CrossRef]
  26. Niu, T.Q.; Gao, Z.D.; Zhang, P.F.; Zhang, X.J.; Gao, M.Y.; Ji, W.; Fan, W.X.; Wen, P.F. MYBA2 gene involved in anthocyanin and flavonol biosynthesis pathways in grapevine. Genet. Mol. Res. 2016, 15, gmr15048922. [Google Scholar] [CrossRef]
  27. Maier, A.; Schrader, A.; Kokkelink, L.; Falke, C.; Welter, B.; Iniesto, E.; Rubio, V.; Uhrig, J.F.; Hülskamp, M.; Hoecker, U. Light and the E3 ubiquitin ligase COP1/SPA control the protein stability of the MYB transcription factors PAP1 and PAP2 involved in anthocyanin accumulation in Arabidopsis. Plant J. 2013, 74, 638–651. [Google Scholar] [CrossRef]
  28. Liu, W.; Mu, H.; Yuan, L.; Li, Y.; Li, Y.; Li, S.; Ren, C.; Duan, W.; Fan, P.; Dai, Z.; et al. VvBBX44 and VvMYBA1 form a regulatory feedback loop to balance anthocyanin biosynthesis in grape. Hortic. Res. 2023, 10, uhad176. [Google Scholar] [CrossRef]
  29. Lotkowska, M.E.; Tohge, T.; Fernie, A.R.; Xue, G.P.; Balazadeh, S.; Mueller-Roeber, B. The Arabidopsis transcription factor MYB112 promotes anthocyanin formation during salinity and under high light stress. Plant Physiol. 2015, 169, 1862–1880. [Google Scholar] [CrossRef]
  30. Zheng, T.; Tan, W.; Yang, H.; Zhang, L.; Li, T.; Liu, B.; Zhang, D.; Lin, H. Regulation of anthocyanin accumulation via MYB75/HAT1/TPL-mediated transcriptional repression. PLoS Genet. 2019, 15, e1007993. [Google Scholar] [CrossRef] [PubMed]
  31. Li, C.; Shi, L.; Wang, Y.; Li, W.; Chen, B.; Zhu, L.; Fu, Y. Arabidopsis ECAP is a new adaptor protein that connects JAZ repressors with the TPR2 co-repressor to suppress jasmonate-responsive anthocyanin accumulation. Mol. Plant 2020, 13, 246–265. [Google Scholar] [CrossRef]
  32. Qi, T.; Song, S.; Ren, Q.; Wu, D.; Huang, H.; Chen, Y.; Fan, M.; Peng, W.; Ren, C.; Xie, D. The Jasmonate-ZIM-domain proteins interact with the WD-Repeat/bHLH/MYB complexes to regulate jasmonate-mediated anthocyanin accumulation and trichome initiation in Arabidopsis thaliana. Plant Cell 2011, 23, 1795–1814. [Google Scholar] [CrossRef]
  33. Li, C.; Shi, L.; Li, X.; Wang, Y.; Bi, Y.; Li, W.; Ma, H.; Chen, B.; Zhu, L.; Fu, Y. ECAP is a key negative regulator mediating different pathways to modulate salt stress-induced anthocyanin biosynthesis in Arabidopsis. New Phytol. 2022, 233, 2216–2231. [Google Scholar] [CrossRef] [PubMed]
  34. An, J.P.; Wang, X.F.; Zhang, X.W.; Xu, H.F.; Bi, S.Q.; You, C.X.; Hao, Y.J. An apple MYB transcription factor regulates cold tolerance and anthocyanin accumulation and undergoes MIEL1-mediated degradation. Plant Biotechnol. J. 2019, 18, 337–353. [Google Scholar] [CrossRef] [PubMed]
  35. Peng, M.; Hudson, D.; Schofield, A.; Tsao, R.; Yang, R.; Gu, H.; Bi, Y.M.; Rothstein, S.J. Adaptation of Arabidopsis to nitrogen limitation involves induction of anthocyanin synthesis which is controlled by the NLA gene. J. Exp. Bot. 2008, 59, 2933–2944. [Google Scholar] [CrossRef]
  36. Jiang, C.; Gao, X.; Liao, L.; Harberd, N.P.; Fu, X. Phosphate starvation root architecture and anthocyanin accumulation responses are modulated by the Gibberellin-DELLA signaling pathway in Arabidopsis. Plant Physiol. 2007, 145, 1460–1470. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Liu, Z.; Liu, J.; Lin, S.; Wang, J.; Lin, W.; Xu, W. GA-DELLA pathway is involved in regulation of nitrogen deficiency-induced anthocyanin accumulation. Plant Cell Rep. 2017, 36, 557–569. [Google Scholar] [CrossRef]
  38. Xie, Y.; Tan, H.; Ma, Z.; Huang, J. DELLA proteins promote anthocyanin biosynthesis via sequestering MYBL2 and JAZ suppressors of the MYB/bHLH/WD40 complex in Arabidopsis thaliana. Mol. Plant 2016, 9, 711–721. [Google Scholar] [CrossRef] [PubMed]
  39. He, Y.; Zhang, X.; Li, L.; Sun, Z.; Li, J.; Chen, X.; Hong, G. SPX4 interacts with both PHR1 and PAP1 to regulate critical steps in phosphorus-status-dependent anthocyanin biosynthesis. New Phytol. 2020, 230, 205–217. [Google Scholar] [CrossRef]
  40. Liu, Z.; Wu, X.; Wang, E.; Liu, Y.; Wang, Y.; Zheng, Q.; Han, Y.; Chen, Z.; Zhang, Y. PHR1 positively regulates phosphate starvation-induced anthocyanin accumulation through direct upregulation of genes F3′H and LDOX in Arabidopsis. Planta 2022, 256, 42. [Google Scholar] [CrossRef]
  41. Wang, L.; Wang, B.; Yu, H.; Guo, H.; Lin, T.; Kou, L.; Wang, A.; Shao, N.; Ma, H.; Xiong, G.; et al. Transcriptional regulation of strigolactone signalling in Arabidopsis. Nature 2020, 583, 277–281. [Google Scholar] [CrossRef]
  42. Shan, X.; Zhang, Y.; Peng, W.; Wang, Z.; Xie, D. Molecular mechanism for jasmonate-induction of anthocyanin accumulation in Arabidopsis. J. Exp. Bot. 2009, 60, 3849–3860. [Google Scholar] [CrossRef]
  43. An, J.P.; Xu, R.R.; Liu, X.; Zhang, J.C.; Wang, X.F.; You, C.X.; Hao, Y.J. Jasmonate induces biosynthesis of anthocyanin and proanthocyanidin in apple by mediating the JAZ1–TRB1–MYB9 complex. Plant J. 2021, 106, 1414–1430. [Google Scholar] [CrossRef] [PubMed]
  44. Su, Z.; Wang, X.; Xuan, X.; Sheng, Z.; Jia, H.; Emal, N.; Liu, Z.; Zheng, T.; Wang, C.; Fang, J. Characterization and action mechanism analysis of VvmiR156b/c/d-VvSPL9 module responding to multiple-hormone signals in the modulation of grape berry color formation. Foods 2021, 10, 896. [Google Scholar] [CrossRef] [PubMed]
  45. An, J.P.; Zhang, X.W.; Liu, Y.J.; Wang, X.F.; You, C.X.; Hao, Y.J.; Hancock, R. ABI5 regulates ABA-induced anthocyanin biosynthesis by modulating the MYB1-bHLH3 complex in apple. J. Exp. Bot. 2021, 72, 1460–1472. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, X.; Hua, D.; Chen, Z.; Zhou, Z.; Gong, Z. Elongator mediates ABA responses, oxidative stress resistance and anthocyanin biosynthesis in Arabidopsis. Plant J. 2009, 60, 79–90. [Google Scholar] [CrossRef]
  47. Jeong, S.W.; Das, P.K.; Jeoung, S.C.; Song, J.Y.; Lee, H.K.; Kim, Y.K.; Kim, W.J.; Park, Y.I.; Yoo, S.D.; Choi, S.B.; et al. Ethylene suppression of sugar-induced anthocyanin pigmentation in Arabidopsis. Plant Physiol. 2010, 154, 1514–1531. [Google Scholar] [CrossRef]
  48. Liu, Y.; Li, M.; Li, T.; Chen, Y.; Zhang, L.; Zhao, G.; Zhuang, J.; Zhao, W.; Gao, L.; Xia, T. Airborne fungus-induced biosynthesis of anthocyanins in Arabidopsis thaliana via jasmonic acid and salicylic acid signaling. Plant Sci. 2020, 300, 110635. [Google Scholar] [CrossRef]
  49. Su, M.; Wang, S.; Liu, W.; Yang, M.; Zhang, Z.; Wang, N.; Chen, X. MdJa2 participates in the brassinosteroid signaling pathway to regulate the synthesis of anthocyanin and proanthocyanidin in red-fleshed apple. Front. Plant Sci. 2022, 13, 830349. [Google Scholar] [CrossRef]
  50. Ni, J.; Wang, S.; Yu, W.; Liao, Y.; Pan, C.; Zhang, M.; Tao, R.; Wei, J.; Gao, Y.; Wang, D.; et al. The ethylene-responsive transcription factor PpERF9 represses PpRAP2.4 and PpMYB114 via histone deacetylation to inhibit anthocyanin biosynthesis in pear. Plant Cell 2023, 35, 2271–2292. [Google Scholar] [CrossRef]
  51. Wang, Y.C.; Wang, N.; Xu, H.F.; Jiang, S.H.; Fang, H.C.; Su, M.Y.; Zhang, Z.Y.; Zhang, T.L.; Chen, X.S. Auxin regulates anthocyanin biosynthesis through the Aux/IAA–ARF signaling pathway in apple. Hortic. Res. 2018, 5, 59. [Google Scholar] [CrossRef]
  52. Wang, C.K.; Han, P.L.; Zhao, Y.W.; Ji, X.L.; Yu, J.Q.; You, C.X.; Hu, D.G.; Hao, Y.J. Auxin regulates anthocyanin biosynthesis through the auxin repressor protein MdIAA26. Biochem. Biophys. Res. Commun. 2020, 533, 717–722. [Google Scholar] [CrossRef]
  53. Araguirang, G.E.; Richter, A.S. Activation of anthocyanin biosynthesis in high light—What is the initial signal? New Phytol. 2022, 236, 2037–2043. [Google Scholar] [CrossRef]
  54. Zhang, Y.; Xu, S.; Cheng, Y.; Peng, Z.; Han, J. Transcriptome profiling of anthocyanin-related genes reveals effects of light intensity on anthocyanin biosynthesis in red leaf lettuce. PeerJ 2018, 6, e4607. [Google Scholar] [CrossRef] [PubMed]
  55. Zoratti, L.; Karppinen, K.; Luengo Escobar, A.; Häggman, H.; Jaakola, L. Light-controlled flavonoid biosynthesis in fruits. Front. Plant Sci. 2014, 5, 534. [Google Scholar] [CrossRef]
  56. Bursch, K.; Toledo-Ortiz, G.; Pireyre, M.; Lohr, M.; Braatz, C.; Johansson, H. Identification of BBX proteins as rate-limiting cofactors of HY5. Nat. Plants 2020, 6, 921–928. [Google Scholar] [CrossRef] [PubMed]
  57. Xu, D.; Jiang, Y.; Li, J.; Lin, F.; Holm, M.; Deng, X.W. BBX21, an Arabidopsis B-box protein, directly activates HY5 and is targeted by COP1 for 26S proteasome-mediated degradation. Proc. Natl. Acad. Sci. USA 2016, 113, 7655–7660. [Google Scholar] [CrossRef] [PubMed]
  58. Ponnu, J.; Riedel, T.; Penner, E.; Schrader, A.; Hoecker, U. Cryptochrome 2 competes with COP1 substrates to repress COP1 ubiquitin ligase activity during Arabidopsis photomorphogenesis. Proc. Natl. Acad. Sci. USA 2019, 116, 27133–27141. [Google Scholar] [CrossRef]
  59. Shin, J.; Park, E.; Choi, G. PIF3 regulates anthocyanin biosynthesis in an HY5-dependent manner with both factors directly binding anthocyanin biosynthetic gene promoters in Arabidopsis. Plant J. 2007, 49, 981–994. [Google Scholar] [CrossRef] [PubMed]
  60. Stracke, R.; Favory, J.J.; Gruber, H.; Bartelniewoehner, L.; Bartels, S.; Binkert, M.; Funk, M.; Weisshaar, B.; Ulm, R. The Arabidopsis bZIP transcription factor HY5 regulates expression of the PFG1/MYB12 gene in response to light and ultraviolet-B radiation. Plant Cell Environ. 2010, 33, 88–103. [Google Scholar] [CrossRef]
  61. Abbas, N.; Maurya, J.P.; Senapati, D.; Gangappa, S.N.; Chattopadhyay, S. Arabidopsis CAM7 and HY5 physically interact and directly bind to the HY5 promoter to regulate its expression and thereby promote photomorphogenesis. Plant Cell 2014, 26, 1036–1052. [Google Scholar] [CrossRef]
  62. Zhou, T.; Meng, L.; Ma, Y.; Liu, Q.; Zhang, Y.; Yang, Z.; Yang, D.; Bian, M. Overexpression of sweet sorghum cryptochrome 1a confers hypersensitivity to blue light, abscisic acid and salinity in Arabidopsis. Plant Cell Rep. 2017, 37, 251–264. [Google Scholar] [CrossRef]
  63. Holm, M.; Ma, L.G.; Qu, L.J.; Deng, X.W. Two interacting bZIP proteins are direct targets of COP1-mediated control of light-dependent gene expression in Arabidopsis. Genes Dev. 2002, 16, 1247–1259. [Google Scholar] [CrossRef] [PubMed]
  64. Podolec, R.; Ulm, R. Photoreceptor-mediated regulation of the COP1/SPA E3 ubiquitin ligase. Curr. Opin. Plant Biol. 2018, 45, 18–25. [Google Scholar] [CrossRef]
  65. Li, S.; Wang, W.; Gao, J.; Yin, K.; Wang, R.; Wang, C.; Petersen, M.; Mundy, J.; Qiu, J.L. MYB75 phosphorylation by MPK4 is required for light-induced anthocyanin accumulation in Arabidopsis. Plant Cell 2016, 28, 2866–2883. [Google Scholar] [CrossRef] [PubMed]
  66. Zheng, X.T.; Yu, Z.C.; Tang, J.W.; Cai, M.L.; Chen, Y.L.; Yang, C.W.; Chow, W.S.; Peng, C.L. The major photoprotective role of anthocyanins in leaves of Arabidopsis thaliana under long-term high light treatment: Antioxidant or light attenuator? Photosynth. Res. 2020, 149, 25–40. [Google Scholar] [CrossRef]
  67. Yang, Y.; Guo, Y. Unraveling salt stress signaling in plants. J. Integr. Plant. Biol. 2018, 60, 796–804. [Google Scholar] [CrossRef] [PubMed]
  68. Van Zelm, E.; Zhang, Y.; Testerink, C. Salt Tolerance Mechanisms of Plants. Annu. Rev. Plant Biol. 2020, 71, 403–433. [Google Scholar] [CrossRef]
  69. Dabravolski, S.A.; Isayenkov, S.V. The role of anthocyanins in plant tolerance to drought and salt stresses. Plants 2023, 12, 2558. [Google Scholar] [CrossRef]
  70. Kim, J.; Lee, W.J.; Vu, T.T.; Jeong, C.Y.; Hong, S.W.; Lee, H. High accumulation of anthocyanins via the ectopic expression of AtDFR confers significant salt stress tolerance in Brassica napus L. Plant Cell Rep. 2017, 36, 1215–1224. [Google Scholar] [CrossRef]
  71. Tohge, T.; Matsui, K.; Ohme-Takagi, M.; Yamazaki, M.; Saito, K. Enhanced radical scavenging activity of genetically modified Arabidopsis seeds. Biotechnol. Lett. 2005, 27, 297–303. [Google Scholar] [CrossRef]
  72. Truong, H.A.; Lee, W.J.; Jeong, C.Y.; Trịnh, C.S.; Lee, S.; Kang, C.S.; Cheong, Y.K.; Hong, S.W.; Lee, H. Enhanced anthocyanin accumulation confers increased growth performance in plants under low nitrate and high salt stress conditions owing to active modulation of nitrate metabolism. J. Plant Physiol. 2018, 231, 41–48. [Google Scholar] [CrossRef]
  73. Lee, Y.J.; Lee, W.J.; Le, Q.T.; Hong, S.W.; Lee, H. Growth performance can be increased under high nitrate and high salt stress through enhanced nitrate reductase activity in Arabidopsis anthocyanin over-producing mutant plants. Front. Plant Sci. 2021, 12, 644455. [Google Scholar] [CrossRef] [PubMed]
  74. Zhu, J.K. Abiotic stress signaling and responses in plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [PubMed]
  75. Gong, Z.; Xiong, L.; Shi, H.; Yang, S.; Herrera-Estrella, L.R.; Xu, G.; Chao, D.Y.; Li, J.; Wang, P.Y.; Qin, F.; et al. Plant abiotic stress response and nutrient use efficiency. Sci. China Life Sci. 2020, 63, 635–674. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, X.; Ding, Y.; Yang, Y.; Song, C.; Wang, B.; Yang, S.; Guo, Y.; Gong, Z. Protein kinases in plant responses to drought, salt, and cold stress. J. Integr. Plant Biol. 2021, 63, 53–78. [Google Scholar] [CrossRef]
  77. Zhang, H.; Zhu, J.; Gong, Z.; Zhu, J.K. Abiotic stress responses in plants. Nat. Rev. Genet. 2021, 23, 104–119. [Google Scholar] [CrossRef] [PubMed]
  78. Naing, A.H.; Kim, C.K. Abiotic stress-induced anthocyanins in plants: Their role in tolerance to abiotic stresses. Physiol. Plant. 2021, 172, 1711–1723. [Google Scholar] [CrossRef]
  79. Cirillo, V.; D’Amelia, V.; Esposito, M.; Amitrano, C.; Carillo, P.; Carputo, D.; Maggio, A. Anthocyanins are key regulators of drought stress tolerance in Tobacco. Biology 2021, 10, 139. [Google Scholar] [CrossRef]
  80. Kovinich, N.; Kayanja, G.; Chanoca, A.; Otegui, M.S.; Grotewold, E. Abiotic stresses induce different localizations of anthocyanins in Arabidopsis. Plant Signal. Behav. 2015, 10, e1027850. [Google Scholar] [CrossRef]
  81. Ding, Y.; Shi, Y.; Yang, S. Molecular regulation of plant responses to environmental temperatures. Mol. Plant 2020, 13, 544–564. [Google Scholar] [CrossRef]
  82. Dar, N.A.; Mir, M.A.; Mir, J.I.; Mansoor, S.; Showkat, W.; Parihar, T.J.; Haq, S.A.U.; Wani, S.H.; Zaffar, G.; Masoodi, K.Z. MYB-6 and LDOX-1 regulated accretion of anthocyanin response to cold stress in purple black carrot (Daucus carota L.). Mol. Biol. Rep. 2022, 49, 5353–5364. [Google Scholar] [CrossRef]
  83. He, Q.; Ren, Y.; Zhao, W.; Li, R.; Zhang, L. Low temperature promotes anthocyanin biosynthesis and related gene expression in the seedlings of purple head Chinese Cabbage (Brassica rapa L.). Genes 2020, 11, 81. [Google Scholar] [CrossRef]
  84. Catalá, R.; Medina, J.; Salinas, J. Integration of low temperature and light signaling during cold acclimation response in Arabidopsis. Proc. Natl. Acad. Sci. USA 2011, 108, 16475–16480. [Google Scholar] [CrossRef] [PubMed]
  85. Cerqueira, J.V.A.; Andrade, M.T.d.; Rafael, D.D.; Zhu, F.; Martins, S.V.C.; Nunes-Nesi, A.; Benedito, V.; Fernie, A.R.; Zsögön, A. Anthocyanins and reactive oxygen species: A team of rivals regulating plant development? Plant Mol. Biol. 2023, 112, 213–223. [Google Scholar] [CrossRef] [PubMed]
  86. Harvaux, M.; Kloppstech, K. The protective functions of carotenoid and flavonoid pigments against excess visible radiation at chilling temperature investigated in Arabidopsis npq and tt mutants. Planta 2001, 213, 953–966. [Google Scholar] [CrossRef]
  87. Calzadilla, P.I.; Vilas, J.M.; Escaray, F.J.; Unrein, F.; Carrasco, P.; Ruiz, O.A. The increase of photosynthetic carbon assimilation as a mechanism of adaptation to low temperature in Lotus japonicus. Sci. Rep. 2019, 9, 863. [Google Scholar] [CrossRef] [PubMed]
  88. Renner, S.S.; Zohner, C.M. The occurrence of red and yellow autumn leaves explained by regional differences in insolation and temperature. New Phytol. 2019, 224, 1464–1471. [Google Scholar] [CrossRef] [PubMed]
  89. Pang, Q.; Yu, W.; Sadeghnezhad, E.; Chen, X.; Hong, P.; Pervaiz, T.; Ren, Y.; Zhang, Y.; Dong, T.; Jia, H.; et al. Omic analysis of anthocyanin synthesis in wine grape leaves under low-temperature. Sci. Hortic. 2023, 307, 111483. [Google Scholar] [CrossRef]
  90. Yang, Z.Q.; Chen, H.; Tan, J.H.; Xu, H.L.; Jia, J.; Feng, Y.H. Cloning of three genes involved in the flavonoid metabolic pathway and their expression during insect resistance in Pinus massoniana Lamb. Genet. Mol. Res. 2016, 15, gmr15049332. [Google Scholar] [CrossRef]
  91. Lu, Y.; Chen, Q.; Bu, Y.; Luo, R.; Hao, S.; Zhang, J.; Tian, J.; Yao, Y. Flavonoid accumulation plays an important role in the rust resistance of malus plant leaves. Front. Plant Sci. 2017, 8, 1286. [Google Scholar] [CrossRef]
  92. Aboshi, T.; Ishiguri, S.; Shiono, Y.; Murayama, T. Flavonoid glycosides in Malabar spinach Basella alba inhibit the growth of Spodoptera litura larvae. Biosci. Biotechnol. Biochem. 2018, 82, 9–14. [Google Scholar] [CrossRef]
  93. Liang, J.; He, J. Protective role of anthocyanins in plants under low nitrogen stress. Biochem. Biophys. Res. Commun. 2018, 498, 946–953. [Google Scholar] [CrossRef] [PubMed]
  94. Peng, M.; Hannam, C.; Gu, H.; Bi, Y.M.; Rothstein, S.J. A mutation in NLA, which encodes a RING-type ubiquitin ligase, disrupts the adaptability of Arabidopsis to nitrogen limitation. Plant J. 2007, 50, 320–337. [Google Scholar] [CrossRef] [PubMed]
  95. Diaz, C.; Saliba-Colombani, V.; Loudet, O.; Belluomo, P.; Moreau, L.; Daniel-Vedele, F.; Morot-Gaudry, J.F.; Masclaux-Daubresse, C. Leaf yellowing and anthocyanin accumulation are two genetically independent strategies in response to nitrogen limitation in Arabidopsis thaliana. Plant Cell Physiol. 2006, 47, 74–83. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, Y.; Chen, Y.F.; Wu, W.H. Potassium and phosphorus transport and signaling in plants. J. Integr. Plant Biol. 2021, 63, 34–52. [Google Scholar] [CrossRef] [PubMed]
  97. Ecker, J.R.; Bustos, R.; Castrillo, G.; Linhares, F.; Puga, M.I.; Rubio, V.; Pérez-Pérez, J.; Solano, R.; Leyva, A.; Paz-Ares, J. A Central regulatory system largely controls transcriptional activation and repression responses to phosphate starvation in Arabidopsis. PLoS Genet. 2010, 6, e1001102. [Google Scholar] [CrossRef]
  98. Liu, K.H.; Liu, M.H.; Lin, Z.W.; Wang, Z.F.; Chen, B.Q.; Liu, C.; Guo, A.; Konishi, M.; Yanagisawa, S.; Wagner, G.; et al. NIN-LIKE PROTEIN 7 transcription factor is a plant nitrate sensor. Science 2022, 377, 1419–1425. [Google Scholar] [CrossRef]
  99. Wang, R.; Cresswell, T.; Johansen, M.P.; Harrison, J.J.; Jiang, Y.; Keitel, C.; Cavagnaro, T.R.; Dijkstra, F.A. Reallocation of nitrogen and phosphorus from roots drives regrowth of grasses and sedges after defoliation under deficit irrigation and nitrogen enrichment. J. Ecol. 2021, 109, 4071–4080. [Google Scholar] [CrossRef]
  100. Lea, U.S.; Slimestad, R.; Smedvig, P.; Lillo, C. Nitrogen deficiency enhances expression of specific MYB and bHLH transcription factors and accumulation of end products in the flavonoid pathway. Planta 2006, 225, 1245–1253. [Google Scholar] [CrossRef]
  101. Sun, H.; Cui, H.; Zhang, J.; Kang, J.; Wang, Z.; Li, M.; Yi, F.; Yang, Q.; Long, R. Gibberellins inhibit flavonoid biosynthesis and promote nitrogen metabolism in Medicago truncatula. Int. J. Mol. Sci. 2021, 22, 9291. [Google Scholar] [CrossRef]
  102. Ito, S.; Nozoye, T.; Sasaki, E.; Imai, M.; Shiwa, Y.; Shibata-Hatta, M.; Ishige, T.; Fukui, K.; Ito, K.; Nakanishi, H.; et al. Strigolactone regulates anthocyanin accumulation, acid phosphatases production and plant growth under low phosphate condition in Arabidopsis. PLoS ONE 2015, 10, e0119724. [Google Scholar] [CrossRef]
  103. Nguyen, T.H.; Goossens, A.; Lacchini, E. Jasmonate: A hormone of primary importance for plant metabolism. Curr. Opin. Plant Biol. 2022, 67, 102197. [Google Scholar] [CrossRef] [PubMed]
  104. Sakamoto, M.; Suzuki, T. Methyl Jasmonate and salinity increase anthocyanin accumulation in Radish Sprouts. Horticulturae 2019, 5, 62. [Google Scholar] [CrossRef]
  105. Guo, Q.; Yoshida, Y.; Major, I.T.; Wang, K.; Sugimoto, K.; Kapali, G.; Havko, N.E.; Benning, C.; Howe, G.A. JAZ repressors of metabolic defense promote growth and reproductive fitness in Arabidopsis. Proc. Natl. Acad. Sci. USA 2018, 115, E10768–E10777. [Google Scholar] [CrossRef] [PubMed]
  106. Brunetti, C.; Sebastiani, F.; Tattini, M. Review: ABA, flavonols, and the evolvability of land plants. Plant Sci. 2019, 280, 448–454. [Google Scholar] [CrossRef] [PubMed]
  107. Sun, Y.; Oh, D.H.; Duan, L.; Ramachandran, P.; Ramirez, A.; Bartlett, A.; Tran, K.N.; Wang, G.; Dassanayake, M.; Dinneny, J.R. Divergence in the ABA gene regulatory network underlies differential growth control. Nat. Plants 2022, 8, 549–560. [Google Scholar] [CrossRef] [PubMed]
  108. Jia, H.F.; Chai, Y.M.; Li, C.L.; Lu, D.; Luo, J.J.; Qin, L.; Shen, Y.Y. Abscisic acid plays an important role in the regulation of strawberry fruit ripening. Plant Physiol. 2011, 157, 188–199. [Google Scholar] [CrossRef]
  109. Shen, X.; Zhao, K.; Liu, L.; Zhang, K.; Yuan, H.; Liao, X.; Wang, Q.; Guo, X.; Li, F.; Li, T. A role for PacMYBA in ABA-regulated anthocyanin biosynthesis in Red-Colored Sweet Cherry cv. Hong Deng (Prunus avium L.). Plant Cell Physiol. 2014, 55, 862–880. [Google Scholar] [CrossRef]
  110. Karppinen, K.; Tegelberg, P.; Häggman, H.; Jaakola, L. Abscisic acid regulates anthocyanin biosynthesis and gene expression associated with cell wall modification in Ripening Bilberry (Vaccinium myrtillus L.) Fruits. Front. Plant Sci. 2018, 9, 1259. [Google Scholar] [CrossRef]
  111. González-Villagra, J.; Cohen, J.D.; Reyes-Díaz, M.M. Abscisic acid is involved in phenolic compounds biosynthesis, mainly anthocyanins, in leaves of Aristotelia chilensis plants (Mol.) subjected to drought stress. Physiol. Plant. 2018, 165, 855–866. [Google Scholar] [CrossRef]
  112. Li, G.; Zhao, J.; Qin, B.; Yin, Y.; An, W.; Mu, Z.; Cao, Y. ABA mediates development-dependent anthocyanin biosynthesis and fruit coloration in Lycium plants. BMC Plant Biol. 2019, 19, 317. [Google Scholar] [CrossRef]
  113. Peng, Y.; Yang, J.; Li, X.; Zhang, Y. Salicylic acid: Biosynthesis and signaling. Annu. Rev. Plant Biol. 2021, 72, 761–791. [Google Scholar] [CrossRef] [PubMed]
  114. Obinata, N.; Yamakawa, T.; Takamiya, M.; Tanaka, N.; Ishimaru, K.; Kodama, T. Effects of salicylic acid on the production of procyanidin and anthocyanin in cultured grape cells. Plant Biotechnol. 2003, 20, 105–111. [Google Scholar] [CrossRef]
  115. García-Pastor, M.E.; Zapata, P.J.; Castillo, S.; Martínez-Romero, D.; Guillén, F.; Valero, D.; Serrano, M. The effects of salicylic acid and its derivatives on increasing pomegranate fruit quality and bioactive compounds at harvest and during storage. Front. Plant Sci. 2020, 11, 668. [Google Scholar] [CrossRef]
  116. Sävenstrand, H.; Brosché, M.; Strid, Å. Ultraviolet-B signalling: Arabidopsis brassinosteroid mutants are defective in UV-B regulated defence gene expression. Plant Physiol. Biochem. 2004, 42, 687–694. [Google Scholar] [CrossRef] [PubMed]
  117. Liang, T.; Shi, C.; Peng, Y.; Tan, H.; Xin, P.; Yang, Y.; Wang, F.; Li, X.; Chu, J.; Huang, J.; et al. Brassinosteroid-activated BRI1-EMS-SUPPRESSOR 1 inhibits flavonoid biosynthesis and coordinates growth and UV-B stress responses in plants. Plant Cell 2020, 32, 3224–3239. [Google Scholar] [CrossRef]
  118. Yang, N.; Zhou, Y.; Wang, Z.; Zhang, Z.; Xi, Z.; Wang, X. Emerging roles of brassinosteroids and light in anthocyanin biosynthesis and ripeness of climacteric and non-climacteric fruits. Crit. Rev. Food Sci. Nutr. 2021, 63, 4541–4553. [Google Scholar] [CrossRef] [PubMed]
  119. Vergara, A.E.; Díaz, K.; Carvajal, R.; Espinoza, L.; Alcalde, J.A.; Pérez-Donoso, A.G. Exogenous applications of brassinosteroids improve color of red table grape (Vitis vinifera L. Cv. “Redglobe”) berries. Front. Plant Sci. 2018, 9, 363. [Google Scholar] [CrossRef]
  120. Chory, J.; Nagpal, P.; Peto, C.A. Phenotypic and genetic analysis of det2, a new mutant that affects light-regulated seedling development in Arabidopsis. Plant Cell 1991, 3, 445–459. [Google Scholar] [CrossRef]
  121. Peng, Z.; Han, C.; Yuan, L.; Zhang, K.; Huang, H.; Ren, C. Brassinosteroid enhances jasmonate-induced anthocyanin accumulation in Arabidopsis Seedlings. J. Integr. Plant Biol. 2011, 53, 632–640. [Google Scholar] [CrossRef]
  122. Yuan, L.B.; Peng, Z.H.; Zhi, T.T.; Zho, Z.; Liu, Y.; Zhu, Q.; Xiong, X.Y.; Ren, C.M. Brassinosteroid enhances cytokinin-induced anthocyanin biosynthesis in Arabidopsis seedlings. Biol. Plant. 2015, 59, 99–105. [Google Scholar] [CrossRef]
  123. Hwang, I.; Sheen, J.; Müller, B. Cytokinin signaling networks. Annu. Rev. Plant Biol. 2012, 63, 353–380. [Google Scholar] [CrossRef]
  124. Deikman, J.; Hammer, P.E. Induction of anthocyanin accumulation by cytokinins in Arabidopsis thaliana. Plant Physiol. 1995, 108, 47–57. [Google Scholar] [CrossRef] [PubMed]
  125. Das, P.K.; Shin, D.H.; Choi, S.B.; Yoo, S.D.; Choi, G.; Park, Y.I. Cytokinins enhance sugar-induced anthocyanin biosynthesis in Arabidopsis. Mol. Cells 2012, 34, 93–101. [Google Scholar] [CrossRef]
  126. Abdelrahman, M.; Nishiyama, R.; Tran, C.D.; Kusano, M.; Nakabayashi, R.; Okazaki, Y.; Matsuda, F.; Chávez Montes, R.A.; Mostofa, M.G.; Li, W.; et al. Defective cytokinin signaling reprograms lipid and flavonoid gene-to-metabolite networks to mitigate high salinity in Arabidopsis. Proc. Natl. Acad. Sci. USA 2021, 118, e2105021118. [Google Scholar] [CrossRef] [PubMed]
  127. Dubois, M.; den Broeck, L.V.; Inzé, D. The pivotal role of ethylene in plant growth. Trends Plant Sci. 2018, 23, 311–323. [Google Scholar] [CrossRef]
  128. Binder, B.M. Ethylene signaling in plants. J. Biol. Chem. 2020, 295, 7710–7725. [Google Scholar] [CrossRef] [PubMed]
  129. Kwon, Y.; Oh, J.E.; Noh, H.; Hong, S.W.; Bhoo, S.H.; Lee, H. The ethylene signaling pathway has a negative impact on sucrose-induced anthocyanin accumulation in Arabidopsis. J. Plant Res. 2011, 124, 193–200. [Google Scholar] [CrossRef]
  130. Xu, Y.; Liu, X.; Huang, Y.; Xia, Z.; Lian, Z.; Qian, L.; Yan, S.; Cao, B.; Qiu, Z. Ethylene inhibits anthocyanin biosynthesis by repressing the R2R3-MYB regulator SlAN2-like in tomato. Int. J. Mol. Sci. 2022, 23, 7648. [Google Scholar] [CrossRef]
  131. Craker, L.E.; Wetherbee, P.J. Ethylene, light, and anthocyanin synthesis. Plant Physiol. 1973, 51, 436–438. [Google Scholar] [CrossRef]
  132. El-Kereamy, A.; Chervin, C.; Roustan, J.P.; Cheynier, V.; Souquet, J.M.; Moutounet, M.; Raynal, J.; Ford, C.; Latché, A.; Pech, J.C.; et al. Exogenous ethylene stimulates the long-term expression of genes related to anthocyanin biosynthesis in grape berries. Physiol. Plant. 2003, 119, 175–182. [Google Scholar] [CrossRef]
  133. Villarreal, N.M.; Bustamante, C.A.; Civello, P.M.; Martínez, G.A. Effect of ethylene and 1-MCP treatments on strawberry fruit ripening. J. Sci. Food Agric. 2010, 90, 683–689. [Google Scholar] [CrossRef]
  134. Cheng, Y.; Liu, L.; Yuan, C.; Guan, J. Molecular characterization of ethylene-regulated anthocyanin biosynthesis in plums during fruit ripening. Plant Mol. Biol. Report. 2015, 34, 777–785. [Google Scholar] [CrossRef]
  135. Xu, H.; Liu, Q.; Yao, T.; Fu, X. Shedding light on integrative GA signaling. Curr. Opin. Plant Biol. 2014, 21, 89–95. [Google Scholar] [CrossRef] [PubMed]
  136. Zhang, Y.; Liu, Z.; Liu, R.; Hao, H.; Bi, Y. Gibberellins negatively regulate low temperature-induced anthocyanin accumulation in a HY5/HYH-dependent manner. Plant Signal. Behav. 2011, 6, 632–634. [Google Scholar] [CrossRef]
  137. Zhao, Y. Auxin biosynthesis and its role in plant development. Annu. Rev. Plant Biol. 2010, 61, 49–64. [Google Scholar] [CrossRef]
  138. Mazzoni-Putman, S.M.; Brumos, J.; Zhao, C.; Alonso, J.M.; Stepanova, A.N. Auxin interactions with other hormones in plant development. Cold Spring Harb. Perspect. Biol. 2021, 13, a039990. [Google Scholar] [CrossRef] [PubMed]
  139. Leyser, O. Auxin signaling. Plant Physiol. 2018, 176, 465–479. [Google Scholar] [CrossRef] [PubMed]
  140. Kang, B.G.; Burg, S.P. Role of ethylene in phytochrome-induced anthocyanin synthesis. Planta 1973, 110, 227–235. [Google Scholar] [CrossRef] [PubMed]
  141. Vince, D. Growth and anthocyanin synthesis in excised Sorghum internodes: I. Effects of growth regulating substances. Planta 1968, 82, 261–279. [Google Scholar] [CrossRef]
  142. Zhou, L.L.; Zeng, H.N.; Shi, M.Z.; Xie, D.Y. Development of tobacco callus cultures over expressing Arabidopsis PAP1/MYB75 transcription factor and characterization of anthocyanin biosynthesis. Planta 2008, 229, 37–51. [Google Scholar] [CrossRef]
  143. Liu, Z.; Shi, M.Z.; Xie, D.Y. Regulation of anthocyanin biosynthesis in Arabidopsis thaliana red pap1-D cells metabolically programmed by auxins. Planta 2013, 239, 765–781. [Google Scholar] [CrossRef] [PubMed]
  144. Clayton-Cuch, D.; Yu, L.; Shirley, N.; Bradley, D.; Bulone, V.; Böttcher, C. Auxin treatment enhances anthocyanin production in the non-climacteric sweet cherry (Prunus avium L.). Int. J. Mol. Sci. 2021, 22, 10760. [Google Scholar] [CrossRef] [PubMed]
  145. Ahmed, M.; Bose, I.; Goksen, G.; Roy, S. Himalayan sources of anthocyanins and its multifunctional applications: A review. Foods 2023, 12, 2203. [Google Scholar] [CrossRef] [PubMed]
  146. Câmara, J.S.; Locatelli, M.; Pereira, J.A.M.; Oliveira, H.; Arlorio, M.; Fernandes, I.; Perestrelo, R.; Freitas, V.; Bordiga, M. Behind the scenes of anthocyanins—From the health benefits to potential applications in food, pharmaceutical and cosmetic fields. Nutrients 2022, 14, 5133. [Google Scholar] [CrossRef] [PubMed]
  147. Denish, P.R.; Fenger, J.A.; Powers, R.; Sigurdson, G.T.; Grisanti, L.; Guggenheim, K.G.; Laporte, S.; Li, J.; Kondo, T.; Magistrato, A.; et al. Discovery of a natural cyan blue: A unique food-sourced anthocyanin could replace synthetic brilliant blue. Sci. Adv. 2021, 7, eabe7871. [Google Scholar] [CrossRef] [PubMed]
  148. Monniaux, M. Unusual suspects in flower color evolution. Science 2023, 379, 534–535. [Google Scholar] [CrossRef]
Figure 1. Plants accumulate anthocyanins in response to biological or abiotic stresses for better growth. When exposed to stresses, the expressions of anthocyanin biosynthetic genes (EBGs/LBGs) are upregulated, subsequently leading to an increase in anthocyanin accumulation within the plants. This heightened accumulation assists the plants in defending against stress by means of scavenging excessive reactive oxygen species (ROS) and reallocating nitrogen resources, among other mechanisms. Red arrow indicates up-regulation of EBGs or LBGs.
Figure 1. Plants accumulate anthocyanins in response to biological or abiotic stresses for better growth. When exposed to stresses, the expressions of anthocyanin biosynthetic genes (EBGs/LBGs) are upregulated, subsequently leading to an increase in anthocyanin accumulation within the plants. This heightened accumulation assists the plants in defending against stress by means of scavenging excessive reactive oxygen species (ROS) and reallocating nitrogen resources, among other mechanisms. Red arrow indicates up-regulation of EBGs or LBGs.
Ijms 24 16415 g001
Figure 2. Mechanism of environmental-stress-related anthocyanin biosynthesis. Environmental stresses can promote anthocyanin biosynthesis by inducing the expressions of anthocyanin biosynthesis regulatory genes (such as MYBs, WD40, and bHLHs). Black arrows indicate positive regulation. This model is modified from Araguirang et al. (2022) [53].
Figure 2. Mechanism of environmental-stress-related anthocyanin biosynthesis. Environmental stresses can promote anthocyanin biosynthesis by inducing the expressions of anthocyanin biosynthesis regulatory genes (such as MYBs, WD40, and bHLHs). Black arrows indicate positive regulation. This model is modified from Araguirang et al. (2022) [53].
Ijms 24 16415 g002
Figure 3. Mechanism of plant-hormone-related anthocyanin biosynthesis in Arabidopsis, apple, and pear. The major regulation network between MBW complex and plant hormones (such as Auxin, ABA, JA, GA, and BR) shows that plant hormones can promote or repress anthocyanin biosynthesis via positively or negatively regulating MBW complex or directly regulating LBG expressions. Black arrows indicate positive regulation and perpendicular lines indicate negative regulation.
Figure 3. Mechanism of plant-hormone-related anthocyanin biosynthesis in Arabidopsis, apple, and pear. The major regulation network between MBW complex and plant hormones (such as Auxin, ABA, JA, GA, and BR) shows that plant hormones can promote or repress anthocyanin biosynthesis via positively or negatively regulating MBW complex or directly regulating LBG expressions. Black arrows indicate positive regulation and perpendicular lines indicate negative regulation.
Ijms 24 16415 g003
Figure 4. Anthocyanins’ potential applications in agriculture, food, and horticulture. Anthocyanins have various potential applications in food, crops, and horticulture due to their health benefits, enhanced stress tolerance, and bright colors.
Figure 4. Anthocyanins’ potential applications in agriculture, food, and horticulture. Anthocyanins have various potential applications in food, crops, and horticulture due to their health benefits, enhanced stress tolerance, and bright colors.
Ijms 24 16415 g004
Table 1. List of genes related to MBW complex and anthocyanin accumulation in plants.
Table 1. List of genes related to MBW complex and anthocyanin accumulation in plants.
ProteinPlant
Species
Environmental Stimuli and
Phytohormones
Function Regarding
Anthocyanin
References
HY5Arabidopsis
(Col-0, Ler, No-0)
LightUpregulation[17]
COP1Arabidopsis
(Col-0)
LightDownregulation[27]
VvBBX44Grape (Vitis vinifera L.)LightDownregulation[28]
MYB112Arabidopsis
(Col-0)
Light and
salinity
Upregulation[29]
HAT1Arabidopsis
(Col-0)
LightDownregulation[30]
TPLArabidopsis
(Col-0)
LightDownregulation[30]
JAZ1/6/8/11Arabidopsis
(Col-0)
JasmonateDownregulation[31,32]
ECAPArabidopsis
(Col-0)
Jasmonate,
salinity, and drought
Downregulation[31,33]
MdMYB308LApple (Malus domestica)ColdUpregulation[34]
NLAArabidopsis
(Col-0)
Low nitrogenUpregulation[35]
DELLAArabidopsis
(Col-0, Ler)
Gibberellin, low phosphorus, and low nitrogenUpregulation[36,37,38]
SPX4Arabidopsis
(Col-0)
Low phosphorusDownregulation[39]
PHR1Arabidopsis
(Col-0)
Low phosphorusUpregulation[40]
SMXL6/7/8Arabidopsis (Col-0)StrigolactoneDownregulation[41]
COI1Arabidopsis
(Col-0)
JasmonateUpregulation[42]
MdMYB9Apple (Malus domestica)JasmonateUpregulation[43]
MdTRB1Apple (Malus domestica)JasmonateUpregulation[43]
MdJAZ1Apple (Malus domestica)JasmonateDownregulation[43]
VvSPL9Grape (Vitis vinifera L.)JasmonateDownregulation[44]
MdABI5Apple (Malus domestica)Abscisic acidUpregulation[45]
MYBL2Arabidopsis
(Col-0)
Ethylene,
Cytokinin and
Abscisic acid
Downregulation[46,47]
ElongatorArabidopsis
(Col-0)
Abscisic acidDownregulation[46]
NPR1Arabidopsis
(Col-0)
Salicylic acidUpregulation[48]
MdJa2Apple (Malus domestica)BrassinosteroidDownregulation[49]
EIN2Arabidopsis
(Col-0, WS)
EthyleneDownregulation[47]
EIN3/EIL1Arabidopsis
(Col-0, WS)
EthyleneDownregulation[47]
ETR1Arabidopsis
(Col-0, WS)
EthyleneDownregulation[47]
PpERF9Pear (Pyrus spp.)EthyleneDownregulation[50]
MdARF13Apple (Malus domestica)AuxinDownregulation[51]
MdIAA26Apple (Malus domestica)AuxinUpregulation[52]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shi, L.; Li, X.; Fu, Y.; Li, C. Environmental Stimuli and Phytohormones in Anthocyanin Biosynthesis: A Comprehensive Review. Int. J. Mol. Sci. 2023, 24, 16415. https://doi.org/10.3390/ijms242216415

AMA Style

Shi L, Li X, Fu Y, Li C. Environmental Stimuli and Phytohormones in Anthocyanin Biosynthesis: A Comprehensive Review. International Journal of Molecular Sciences. 2023; 24(22):16415. https://doi.org/10.3390/ijms242216415

Chicago/Turabian Style

Shi, Lei, Xing Li, Ying Fu, and Changjiang Li. 2023. "Environmental Stimuli and Phytohormones in Anthocyanin Biosynthesis: A Comprehensive Review" International Journal of Molecular Sciences 24, no. 22: 16415. https://doi.org/10.3390/ijms242216415

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