Next Article in Journal
Magnetic Biochar Derived from Fenton Sludge/CMC for High-Efficiency Removal of Pb(II): Synthesis, Application, and Mechanism
Previous Article in Journal
A Comparative Analysis on the Environmental Impact of Selected Methods for Determining the Profile of Fatty Acids in Cheese
Previous Article in Special Issue
Hydrogel for the Controlled Delivery of Bioactive Components from Extracts of Eupatorium glutinosum Lam. Leaves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Classification, Biosynthesis, Biological Activities and Potential Applications of Flavonoids

1
School of Horticulture and Landscape Architecture, Yangzhou University, Yangzhou 225009, China
2
Key Laboratory of Plant Resource Conservation and Germplasm Innovation in Mountainous Region (Ministry of Education), Institute of Agro-Bioengineering, Guizhou University, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(13), 4982; https://doi.org/10.3390/molecules28134982
Submission received: 15 May 2023 / Revised: 20 June 2023 / Accepted: 20 June 2023 / Published: 25 June 2023

Abstract

:
Flavonoids represent the main class of plant secondary metabolites and occur in the tissues and organs of various plant species. In plants, flavonoids are involved in many biological processes and in response to various environmental stresses. The consumption of flavonoids has been known to reduce the risk of many chronic diseases due to their antioxidant and free radical scavenging properties. In the present review, we summarize the classification, distribution, biosynthesis pathways, and regulatory mechanisms of flavonoids. Moreover, we investigated their biological activities and discuss their applications in food processing and cosmetics, as well as their pharmaceutical and medical uses. Current trends in flavonoid research are also briefly described, including the mining of new functional genes and metabolites through omics research and the engineering of flavonoids using nanotechnology. This review provides a reference for basic and applied research on flavonoid compounds.

1. Introduction

Flavonoids, a group of important secondary metabolites, are widely distributed in widely found in fruits, vegetables, herbs, stems, cereals, nuts, flowers, and seeds [1]. So far, over 10,000 flavonoid compounds have been isolated and identified [2]. Flavonoids processed favorable biochemical and antioxidant effects on various diseases such as cardiovascular disease, cancer, and neurodegenerative disease [3]. Flavonoids are associated with a broad spectrum of health-promoting effects and are an indispensable component in a variety of nutraceutical, pharmaceutical, medicinal and cosmetic applications. This is mainly attributed to their anti-inflammatory, anti-oxidative, anti-carcinogenic, and anti-mutagenic properties as well as the ability to regulate key cellular enzyme function [4,5].
Flavonoids are synthesized through the phenylpropanoid metabolic pathway and possess 15 carbon atoms arranged in three rings (C6–C3–C6) labeled A, B, and C [6]. Flavonoids play a variety of biological activities in plants, animals, and bacteria [7]. In plants, flavonoid compounds are normally found in flowers, fruits, leaves, and seeds, which are responsible for their color, fragrance, and flavor characteristics [8]. As important secondary metabolites, plant flavonoids are involved in regulating auxin transport, male fertility, pollination, seed development, flower coloring, and allelopathy [9]. Flavonoids play protective roles against abiotic (ultraviolet radiation, cold, salt, drought, and heavy metals) and biotic (herbivores, bacteria, fungi) stresses [10]. Flavonoids are classified into various types depending on their chemical structure, degree of unsaturation, and oxidation of carbon ring, including flavones, flavanones, isoflavones, flavonols, chalcones, flavanols, and anthocyanins. Each of these flavonoids is widely distributed in nature [4,11].
Flavonoids have attracted increasing research interest in recent years. Thus, in the present review, we systematically summarize recent progress in plant flavonoid research in terms of their classification, biosynthesis, biological activity, and potential applications in food processing, cosmetics, and pharmaceutical industries [5,12,13]. We also discuss the use of omics research for mining functional genes and identifying new flavonoid compounds. In addition, we summarize recent research progress on the role of flavonoid nanoparticles in various diseases. The main purpose of this review is to provide up-to-date information on the classification, biosynthesis, biological activity, and potential applications of flavonoids. In addition, research into the therapeutic potential of flavonoid nanoparticles has highlighted their value and importance in preventing life-threatening diseases and improving global health.

2. Methodology

To perform a comprehensive review of plant flavonoids, emphasizing the classification, biosynthesis and regulatory mechanisms, biological activities, potential applications, omics research, and flavonoid nanoparticles, a thorough literature search was performed. Seven scientific electronic databases, including SCiFinder, PubMed, MEDLINE, EMBASE, Scopus, ScienceDirect, and the Google Scholar library, were used for searching in PubMed. The keywords were searched alone or in combination with other keywords. The PRISMA flow chart was used to display the methodology for the published information selection through Page’s method (Figure A1) [14,15].

3. Flavonoid Classification

Flavonoids are mainly found in plant cell vacuoles in the form of C-glycosides or O-glycosides [16]. The basic molecular structure of flavonoids depends upon their basic C6–C3–C6 skeleton (labeled A, B, and C, Figure 1). Flavonoids are classified into seven subclasses based on modifications to their basic skeletons; these subclasses include flavones, flavanones, isoflavones, flavonols, chalcones, flavanols, and anthocyanins [17,18].

3.1. Flavones

Flavones, one of the largest classes of flavonoids, consist of 4H-chromen-4-one bearing a phenyl substituent at position 2. Flavones mostly occur as 7-O-glycosides, which are found in celery, parsley, red pepper, chamomile, mint, and ginkgo [5,19,20]. Apigenin and luteolin are two common flavones (Figure A2). In nature, apigenin is usually found in a glycosylated form, with a sugar moiety attached to the tricyclic core structure via hydroxyl groups (O-glycosides) or directly to carbon (C-glycosides) [21]. The principal ingredients of apigenin are glycosylated apiin, apigenin, vitexin, isovitexin, or rhoifolin. Apigenin can scavenge free radicals and regulate antioxidant enzyme activity in pancreatic cells, and apigenin can decrease inflammation in cancer, neuroinflammation, and cardiovascular diseases [9,22].

3.2. Flavonols

Flavonols, also called 3-hydroxy flavone, can be identified by specific substitutions in their A- and B-rings, which are connected by a three-carbon chain [23]. Flavonols possess hydroxyl groups at positions 5 and 7 in the A-ring and are mainly present in epidermal cells to protect DNA against UV-induced damage [24]. Four types of flavonol compounds (quercetin, galangin, kaempferol, and myricetin) are mainly distributed in vegetables and fruits, such as asparagus, onions, lettuce, broccoli, tomato, and apples (Figure A2) [25]. Flavonols exhibit interesting biological activities, including antioxidant, antibacterial, cardioprotective, anticancer, and antiviral activities. Dietary flavonols can significantly decrease the risk of gastric cancer in smokers and in women (Figure A2).

3.3. Flavanones

Flavanones (dihydro-flavones) possess a saturated C-ring [26]. The saturated double bond between positions 2 and 3 in the C-ring represents the only structural difference between flavanones and other flavonoid compounds [27]. Flavanones are mainly distributed in citrus fruits, including oranges, lemons, mandarins, grapefruits, clementines, and limes [28]. Flavanones contain hydroxyl groups at positions 5 and 7 in the A-ring and possess hydroxyl/methoxy substituents at the C3 or C4 positions of the B-ring [29]. The defining characteristic of flavanones is a disaccharidic moiety linked to the seven positions of aglycone [30]. Depending on their structural differences, flavanones can occur in the form of naringin, naringenin, hesperidin, hesperetin, pinocembrin, likvirtin, and eriodictyol [31]. Among them, naringenin and hesperetin, as the main dietary flavanones, occur almost exclusively in citrus fruits (Figure A2) [28,32]. Naringin can increase the activity of antioxidant enzymes (CAT, PON, GPx, and SOD) and enhance the immune system. Furthermore, naringenin and hesperetin have been shown to recover impaired thyroid function in rats.

3.4. Isoflavonoids

Isoflavones have a B-ring at the C3 position of the heterocyclic C-ring of the diphenylpropane (C6–C3–C6) backbone, which represents their only chemical structural difference from other flavonoids [33]. Isoflavonoids are characteristic metabolites of leguminous plants and play essential roles in nodule induction and microbial signaling in legumes [34,35]. Isoflavones are classified into three groups: genistein, daidzein, and glycitein (Figure A2) [36]. The molecular structure of isoflavones is similar to that of animal estrogens. Isoflavones are phytoestrogens that exhibit potent estrogenic activity. Phytoestrogens are similar in structure to the human female hormone 17-β-estradiol in that they bind to estrogen receptors [37]. In addition, isoflavones possess a strong antioxidant activity, which can decrease the risk of cancers by inhibiting free radical-induced DNA damage [37].

3.5. Flavanols

Flavanols, also called catechins or flavan-3-ols, are characterized by a hydroxyl group at position 3 in the C-ring [38]. Flavanols lack a double bond between positions 2 and 3 in the C-ring [39]. Several flavanols, including catechin, gallocatechin 3-gallate, gallocatechin, epicatechin, epicatechin 3-gallate, catechin 3-gallate, and epicatechin 3-gallate, are widely distributed in many fruits (e.g., apples, bananas, pears, and blueberries) [40,41]. Flavanols can protect blood vessels against tobacco by increasing the content of NO in blood vessels. A flavanol-rich diet can facilitate the permanent improvement of endothelial function and prevent the development of cardiovascular diseases [42,43].

3.6. Anthocyanins

Anthocyanins, as glycosylated polyphenolic compounds, are a group of soluble vacuolar pigments that possess a range of colors, from orange, red, and purple to blue, depending on the pH of the micro-environment of the flowers, seeds, fruits, and vegetative tissues [44]. The position and number of hydroxyl and methoxyl groups present as substituents in the flavylium structure result in different anthocyanins (Figure A2). Thus, over 650 anthocyanins have been identified in many plants [45]; these are grouped into five items, including cyanidin, delphinidin, malvidin, pelargonidin, and peonidin, and their corresponding derivatives [46]. Anthocyanins are mainly found in the outer cell layer of various fruits and vegetables, such as blackcurrants, grapes, and berries [47,48]. The antioxidant ability of anthocyanins is associated with their ring orientation and the position and number of free hydroxyls around the pyrone ring. Anthocyanins play important roles in visual acuity, cholesterol decomposition, and the reduced risk of cardiovascular disease in humans [49,50]. In addition, anthocyanins are commonly used as food colorants.

3.7. Chalcones

Chalcones (1,3-diaryl-2-propen-1-ones) are natural open-chain flavonoids, carrying up to three modified or unmodified C5-, C10-, and C15-prenyl moieties on both their A and B-rings. These bioactive products are widely distributed in the Fabaceae, Moraceae, Zingiberaceae, and Cannabaceae families [9]. They exhibit a wide spectrum of pharmacological effects, including antioxidant, antibacterial, anthelmintic, antiulcer, antiviral, antiprotozoal, and anticancer effects [51]. Chalcones are precursors of flavonoids and isoflavonoids. Their structural features are easily constructed from simple aromatic compounds. Their prominent bioactivity has inspired the synthesis of chalcone analogs, as well as minor structural modifications to natural chalcones; these compounds form a large collection of bioactive chalcone derivatives [52]. Xanthohumol and isbavirachalone are two representative derivatives that exhibit abundant biological and pharmacological activity (Figure A2) [49].
Generally, the number and position of –OH groups have a great influence on flavonoid bioactivity. The –OH groups can link to the carbon atoms of the benzene ring (3,5,7, and 3′,4′-dihydroxy substitution pattern), which directly determines the bioactivity of flavonoids. Moreover, the position of the –OH group also influenced the flavonoid bioactivity. The most effective radical scavengers are flavonoids with the 3′,4′-dihydroxy substitution pattern on the B-ring and/or hydroxyl group at the C-3 position. In addition, the C2–C3 double bond is not necessary for high activity. Flavanols lacking the C2–C3 double bond displayed strong activity. The presence of a 3 –OH group significantly enhances the bioactivity.

4. Flavonoid Biosynthesis in Plants

4.1. Flavonoid Biosynthetic Pathways

Flavonoid synthesis occurs at the junction of the shikimate pathway and the acetate pathway. The former can generate p-coumaroyl-CoA, and the latter regulates C2-chain elongation [53] (Figure 2). Phenylalanine ammonia-lyase (PAL) deaminates phenylalanine to ammonia and cinnamic acid [54]. Then, C4H (cinnamic acid 4-hydroxylase) catalyzes the production of 4-coumaric acid [4], and 4CL (4-coumaric acid: CoA ligase) converts 4-coumaric acid to form 4-coumaroyl-CoA, which is a key enzyme in the phenylpropanoid metabolic pathway that regulates the biosynthesis of lignin and flavonoids [55].

4.2. Transcriptional Regulation of Flavonoid Synthesis

Flavonoid biosynthesis is tightly regulated by biosynthetic enzymes and regulatory transcription factors (TFs) [65]. Several TF families have been reported to be involved in regulating flavonoid biosynthesis in plants, including WRKY, Dof, MADS-box, bZIP, MYB, bHLH, WD40, and NAC (Table 1) [66]. Plant MYBs are characterized by a highly conserved MYB DNA-binding domain and are further classified into four groups based on the position and number of MYB repeats: 1R-MYB, 2R-MYB, 3R-MYB, and 4R-MYB [67]. Among them, R2R3-MYB TFs are involved in regulating the expression of structural genes in the flavonoid pathway [68]. For example, transgenic tobacco overexpressing NtMYB3 from Narcissus tazetta can reduce the content of flavonoids by inhibiting the expression of FLSs [69]. Transgenic Arabidopsis overexpressing GbMYB2 from Ginkgo biloba can decrease flavonoid accumulation by inhibiting the expression of some structural genes (e.g., GbPAL, GbFLS, GbANS, and GbCHI) [70]. Yan et al. revealed that soybean GmMYB100 negatively regulated flavonoid biosynthesis by inhibiting the activities of CHS and CHI promoters [71]. In addition, the overexpression of PpMYB17 in pear calli was found to bind and activate the promoters of structural genes of PpCHS, PpCHI, PpF3H, PpFLS, and PpUFGT under light conditions, which enhanced the biosynthesis of flavonoids [72]. Transgenic tobacco overexpressing FtMYB31 from Fagopyrum tataricum increased the expression of CHS, F3H, and FLS genes and promoted the accumulation of flavonoids [73]. The overexpression of SbMYB8 from Scutellaria baicalensis in transgenic tobacco promoted the expression of the SbCHS gene, increased flavonoid content, and enhanced the activities of antioxidant enzymes in transgenic tobacco [63]. Furthermore, bHLH TFs play essential roles in regulating the biosynthesis of flavonoids. CsMYC2 was able to promote flavonoid biosynthesis by increasing the expression of the UFGT gene [74]. MdbHLH3 promoted anthocyanin accumulation and fruit coloration in response to low temperatures in apples [75]. In addition, MBW complexes (MYB-bHLH-WD40) regulate flavonoid biosynthesis in different plants [64,76]. The TT2–TT8–TTG1 complex plays a major role in developing seeds and also plays an important role in regulating the expression of LBGs (DFR, LDOX, TT19, TT12, AHA10, and BAN) [77]. Moreover, the MBW complex exhibits tissue-specific regulation of the expression of the genes involved in flavonoid biosynthesis [78]. The MYB5–TT8–TTG1 complex is active in the endothelium, regulating DFR, LDOX, and TT12 expression, whereas the TT2–EGL3/GL3–TTG1 complexes regulate the expression of LDOX, BAN, AHA10, and DFR in the chalaza [78].
In addition, several TF families, including bZIP, NAC, Dof, and WRKY, play important roles in regulating flavonoid biosynthesis [79,80]. For example, VvibZIPC22 was able to bind and activate the promoters of structural genes of VviCHI and VviCHS to increase their flavonoid contents [81]. Transgenic tobacco overexpressing NtHY5 increased the expression of phenylpropanoid pathway genes, promoted the biosynthesis of flavonoids, and enhanced plant tolerance to salt stress [82]. Transgenic Arabidopsis overexpressing AtNAC078 increased the content of flavonoids under strong light conditions by upregulating the expression of CHS, F3′H, DFR, and LDO [83]. MdNAC52 promoted the biosynthesis of flavonoid compounds (anthocyanins and procyanidins) in apples by binding and activating the promoters of MdMYB9, MdMYB11, and LAR [84]. Arabidopsis AtDOF4 upregulated the expression of structural genes of DFR, LDOX, TT19, and PAP1 to increase the content of flavonoids in plants [85]. Apple callus overexpressing MdWRKY11 was able to increase the expression of F3H, FLS, DFR, ANS, and UFGT and promote the biosynthesis of flavonoids and anthocyanins [86].

4.3. Non-Coding RNA Regulates Flavonoid Biosynthesis

Non-coding RNA, including lncRNA (long non-coding RNAs) and microRNA, played important roles in regulating flavonoid biosynthesis [87]. lncRNAs may act as precursors and endogenous target mimics of miRNAs to indirectly regulate protein-coding genes (PCgenes) [87]. Two lncRNAs, XR_001591906 and MSTRG.9304, were found to regulate the expression of the CHS gene in flavonoid biosynthesis during peanut seed development [88]. miRNAs directly cleave structural genes (SG) for flavonoid synthesis, thereby negatively regulating the accumulation of flavonoids, including miR396-targeting UFGT, miR172-targeting 4CL, and miR829.1-targeting CHS [89]. The miRNA-directed cleavage of TFs involved in flavonoid synthesis through miRNA–TF–SG regulatory networks such as miR156–SPL–F3H, miR828/TAS4–MYBs–DFR, and miR858–MYBs–CHS/FLS [89,90].

5. Pharmacological Properties of Flavonoids

5.1. Antioxidant Activity

The molecular mechanisms of antioxidant action contain the following processes (Figure 3). The antioxidant activity of flavonoids depends on their stable roles in different matrices, as well as the position and number of hydroxyl groups in their structures (Table A1) [91,92]. The mechanisms of antioxidant activity mainly comprise the following three processes: (1) Direct scavenging of ROS. In vitro, flavonoid antioxidant activity depends on the arrangement of hydroxyl groups. The B-ring, with its orthodihydroxy structure, allows flavonoid phenoxyl radicals to participate in electron delocalization, leading to the dislocation of an electron in the B-ring, which exhibits antioxidant activity. These structures enhance antioxidant activity [9]. Antioxidant activity was found to be increased in vitro through the polymerization of flavonoid monomers, such as proanthocyanidins (also called condensed tannins), and the polymers of catechins, which displayed excellent antioxidant activity due to the large number of hydroxyl groups in their molecules [93]. Moreover, free radical scavenging activity was found to be promoted by double bonds in the phenolic ring and hydroxyl side chains and through the glycosylation of anthocyanidins [9]. (2) The activation of antioxidant enzymes. Some flavonoids can suppress the activities of free radical-generating enzymes such as nitric-oxide synthase and xanthine oxidase [94]. Furthermore, flavonoids activate antioxidant enzymes, including the induction of phase II detoxifying enzymes (e.g., NAD(P)H-quinone oxidoreductase, glutathione S-transferase, and UDP-glucuronosyltransferase), which are the major defense enzymes against electrophilic toxicants and oxidative stress. The electrophile-responsive element (EpRE) is distributed in the 5′-regulatory region of a number of genes encoding phase II enzymes. Flavonoids have been found to activate the EpRE-mediated gene expression and display redox properties [95]. (3) The inhibition of oxidases. Flavonoids exert antioxidant activity by inhibiting the activities of xanthine oxidase (XO) and protein kinase C, which catalyze the production of superoxide anion. Zeng et al. reported that fisetin reduced oxidative damage by inhibiting XO activity [96]. Furthermore, flavonoids can inhibit the activities of cyclooxygenase, NADH oxidase, microsomal monooxygenase, and lipoxygenase [10,97]. (4) Metal-chelating activity. Hydroxyflavones may generate complexes with metal cations, and their chelating properties differ significantly depending on the number and position of the hydroxyl substituents [98]. Specific flavonoids can decrease the toxicity of redox-active transition metal ions [99]. Lesjak et al. revealed that quercetin, catechin, and rutin displayed high antioxidant activity against Fe (III) [100]. Kaempferol can chelate Cu(II) ions, suppress the formation of oxidants and radicals, and possess strong antioxidant properties [101]. Numerous flavonoids can generate stable metal complexes between the carbonyl moiety and their OH groups. Quercetin possesses three potential bidentate binding sites (α-hydroxy-carbonyl, β-hydroxy-carbonyl, and catechol), which can generate stable metallic complexes. In addition, catechin can chelate a large number of metal ions, including Mn (VI), Fe(II), Fe(III), Cu(II), Zn(II), and Al(III) [102]. (5) Increasing the level of α-tocopherol radicals. The oxidation of low-density lipoprotein (LDL) represents a free radical-driven lipid peroxidation process, which is associated with human health [103]. Flavonoids act as hydrogen donors to α-tocopherol radicals and have great potential to delay LDL oxidation through interaction with the α-tocopherol radicals. Torello et al. (2021) reported that several flavonoids (e.g., quercetin, epigallocatechin gallate, and naringin) showed strong inhibitory activity against LDL oxidation in vitro [104]. (6) Mitigating oxidative stress caused by nitric oxide. Nitric oxide (NO) plays an important role in maintaining the dilation of blood vessels, and a loss of NO causes oxidative stress in the vasculature [105]. The interaction of NO and O2− forms the cytotoxic product peroxynitrite, which is a powerful oxidant capable of causing pathological damage. Flavonoids can inhibit the production of NO in several lipopolysaccharide-activated cells by inhibiting inducible NOS expression [106]. Among flavonoids, quercetin can suppress the production of NO and vascular endothelial growth factor (VEGF) [107].
Table 1. Pharmacological activities of flavonoids.
Table 1. Pharmacological activities of flavonoids.
FlavonoidsClassificationPharmacological ActivitySources of PlantReferences
Proanthocyanidinsanthocyaninsantioxidant, anti-inflammatory, antibacterial, antifungal and anti-cardiovasculargrapes, apples, sorghum, cherries, and other natural plant[93]
Cyanidinanthocyaninsanti-inflammatory, antiviral, and anticancerblack rice, black beans, purple potatoes, blueberries[108]
Curcumincurcuminoidsanti-inflammatory and anticancerCurcuma longa[109]
Methyl chalconechalconesanti-inflammatory and anticancerapple, citrus, soybean, ginger, mulberry[110]
Trans-chalconechalconesanti-inflammatory and anticancerapple, citrus, soybean, ginger, mulberry[110]
Xanthohumolchalconesanti-cardiovascular and antiviralHumulus lupulus[111]
Licochalconechalconesantibacterial and antifungalGlycyrrhiza uralensis[112]
Catechinflavanolsantioxidant, anti-inflammatory, antiviral, and anti-cardiovascularCamellia sinensis[101,102,113]
Epigallocatechin gallateflavanolsantioxidant, antibacterial, antifungal, anti-cardiovascular, and antiviralCamellia sinensis[104,114,115]
Naringinflavanonesantioxidant, anti-inflammatory, anti-cardiovascular, and antivirallemons, oranges, grapefruits, citrus[104,110,116,117,118,119]
Hesperidinflavanonesanti-inflammatory, anti-cardiovascular, and antivirallemons, limes, oranges, grapefruits, citrus[116,117,120,121]
Diosminflavanonesanti-inflammatorycitrus fruits[122]
Orientinflavanonesanti-inflammatoryTrollius chinensis, Cajanus cajan, Crataegus laevigata[123]
Vitexinflavanonesantioxidant, anti-inflammatory, and anticancerFicus deltoid, Spirodela polyrhiza[123]
Acacetinflavanonesanti-cardiovascular, anticancer, and antiviralAcacia farnesiana[124,125]
Silymarinflavanonesantioxidant, anti-cardiovascular, and antiviralSilybum marianum[126,127]
Liquiritigeninflavanonesanti-inflammatory, antiviral, and anticancerGlycyrrhiza uralensis[128]
Isorhamnetinflavanonesantiviral and anticancerGinkgo biloba, Hippophae rhamnoides[125]
Apigeninflavonesantibacterial, antifungal, and antiviralApium graveolens[129,130,131,132]
Morinflavonesantioxidant and anti-inflammatoryCudrania cochinchinensis, Maclura pomifera[133]
BaicalinflavonesAnti-cardiovascular, antibacterial, and antifungalScutellaria baicalensis[114,134]
Luteolinflavonesanti-inflammatory, anti-cardiovascular, and antiviralDracocephalum integrifolium, Lonicera japonica, Capsicum annuum[132,135]
Fisetinflavonolsantioxidantstrawberry, apple, onion, cucumber, and other fruits and vegetables[96]
Quercetinflavonolsantioxidant, anti-inflammatory, anti-cardiovascular, antibacterial, and antifungalvegetables, fruit, seeds, nuts, tea, and red wine[100,102,107,120,136,137,138]
Rutinflavonolsantioxidant, anti-inflammatory, and antiviralrue, tobacco, jujube, apricot, orange, tomato, buckwheat, and citrus fruits[101,120,126,127]
Kaempferolflavonolsantioxidant, anti-inflammatory, antibacterial, antiviral, and anticancerfruits, vegetables, herbs, and other natural plants[101,133,139]
Myricetinflavonolsantioxidant, anti-inflammatory, and anti-cardiovascularMyrica rubra[133,140,141]
Glabrolisoflavaneantibacterial and antifungalGlycyrrhiza uralensis[112]
Genisteinisoflavoneantioxidant, antifungal, antiviral, and anticancersoybeans and other plants[120,142,143]

5.2. Anti-Inflammatory Action

Inflammation has many causes, including infections, injuries, and diseases [144]. Chronic inflammation is a common pathological basis for age-associated diseases such as cardiovascular disease, diabetes, cancer, and Alzheimer’s disease [145]. Among these flavonoids, apigenin can reduce the steady-state mRNA levels induced by TNF-α and can downregulate the expression of intercellular adhesion molecule-1 (ICAM-1), E-selectin, and vascular cell adhesion molecule-1 (VCAM-1) to endothelial cells. Yang et al. reported that many flavonoids reduced the expression of pro-inflammatory cytokines (e.g., TNF-α, IL-1β, IL-8, IL-6, and monocyte chemoattractant protein-1 (MCP-1)) in Jurkat T cells, peripheral blood mononuclear cells, and RAW macrophages. Quercetin and catechins were found to increase IL-10 production through the combined inhibition of TNF-α and IL-1β [113]. Quercetin suppressed the activity of heat shock factor (HSF) and reduced heat-induced damage [136]. Furthermore, several flavonols (e.g., morin, quercetin, kaempferol, and myricetin) could inhibit the activity of lipoxygenase [133]. Many flavonoids could decrease the production of arachidonic acid and suppress the activities of phospholipase A2, cyclooxygenase, and NOS, which further decreased the production of key inflammatory substances (prostaglandins, leukotrienes, and NO) [146]. In addition, flavonoids could inhibit the production of arachidonic acid metabolites and chemokines and decrease leukocyte infiltration and edema [147]. Moreover, flavonoids could chelate metal iron, suppress the activation of the complement system, and lower inflammation [148]. Genistein reduced airway hyper-responsiveness, ovalbumin-induced bronchoconstriction, and pulmonary eosinophilia in a guinea pig model of asthma and also suppressed the inflammatory response and joint destruction in collagen-induced arthritic mice. Quercetin, rutin, and hesperidin decreased chronic inflammation in an experimental model, and the results further showed that rutin played a key role in the chronic phase [120]. Diosmin and hesperidin could inhibit leukotriene B4 (LTB4) biosynthesis, which reduces colitis in a trinitrobenzenesulfonic acid (TNBS)-induced colitis rat model [122]. Naringenin displayed anti-inflammatory activity by inhibiting pro-inflammatory cytokines and reducing leukocyte infiltration. Several chalcones (e.g., trans-chalcone, hesperidin, and methyl chalcone) displayed inhibitory activity against pro-inflammatory cytokines [110]. Zhong et al. revealed that the flavonoids (vitexin, orientin, and rutin) from Tartary buckwheat sprout could reduce NO production by inhibiting iNOS and COX2 expression in lipopolysaccharide (LPS)-induced RAW cells and in male BALB/c mice [123]. Quercetin and myricetin protected 661 W cells and the cone photoreceptor cell line from the toxic effects of H2O2. They also enhanced the gene expression of M and S opsins under oxidative stress [140].

5.3. Cardiovascular Action

Flavonoids used as cardioprotective agents can regulate oxidative stress and inflammation, commonly cause vasodilation, and regulate endothelial cell apoptosis [149]. Recent studies have revealed that dietary flavonoid intake can effectively reduce the risk of cardiovascular death in adult Americans [150]. In vitro and in animal models, some flavonoids have vasodilatory effects, improve endothelial dysfunction and insulin resistance, exert antiplatelet aggregation and atherosclerotic protective effects, and lower blood pressure [151]. Flavonoids were found to prevent hepatic steatosis, dyslipidemia, and insulin sensitivity, primarily by inhibiting hepatic fatty acid synthesis and increasing fatty acid oxidation [152]. Several studies have demonstrated that naringenin, quercetin, and hesperetin possess vasodilator properties, and their results further showed that naringenin could reduce high blood pressure by promoting vasodilation [116,117]. Isoflavones seem to protect against inflammatory vascular diseases by inhibiting monocyte-endothelial cell adhesion [153]. Quercetin exerts cardioprotective effects of quercetin against ischemia-reperfusion injury in the heart and also exhibits atheroprotective activity [137]. Baicalin can improve cardiac dysfunction and inhibit apoptosis in the heart. Chrysin has been reported to possess antiplatelet activity [134]. Anthocyanins reduce the risk of myocardial infarction (MI) in humans, improve systolic blood pressure, and reduce the content of triglycerides, total cholesterol, and low-density lipoprotein cholesterol [154,155]. Several studies have demonstrated that acacetin plays an important role in regulating human arrhythmia [124]. Sun et al. (NIDAN) found that xanthohumol could increase PTEN expression and inhibit AKT/mTOR phosphorylation in isoprenaline-treated mice and exert a protective effect against ISO-induced myocardial hypertrophy and fibrosis [111]. Flavonoids extracted from Myrica rubra exerted cardioprotective effects by modulating the PI3K/Akt/GSK3β pathway to attenuate oxidative damage and cardiomyocyte apoptosis [141]. The sirtuin 1 (SIRT1) enzyme plays a key role in the regulation of many physiological functions, and its reduced expression often causes aging-related diseases such as myocardial hypertrophy, myocardial infarction, and endothelial dysfunction. A study by Testai et al. found that the long-term administration of the citrus flavonoid naringenin (NAR) (100 mg/kg/day) in mice resulted in enhanced SIRT1 expression, significantly reduced ROS production in myocardial tissue and significantly lowered levels of the cardiovascular inflammatory markers TNF-α and IL6, revealing that nutritional therapy with NAR may help improve myocardial aging and protect cardiac function [118]. Yang et al. also revealed that Oxytropis falcata flavonoid extracts protect against myocardial ischemia-reperfusion injury by downregulating the ROS-mediated JNK/p38MAPK/NFκB pathway to regulate inflammatory responses, oxidative stress, and apoptosis [156].

5.4. Antibacterial and Antifungal Action

Flavonoids exert their antibacterial activity through several mechanisms, including bacterial membrane disruption, the inhibition of biofilm formation, the suppression of nucleic acid and ATP biosynthesis, and the disruption of electron transport (Table A1) [6,157]. Quercetin, apigenin, naringenin, chrysin, genistein, kaempferol, daidzin, and daidzein were found to block the biofilm formation, and the results further showed that quercetin, myricetin, baicalein, and luteolin suppressed the DNA replication in bacteria [8,158]. The epigallocatechin gallate and baicalein suppressed the biosynthesis of ATP in bacteria [114]. Wu et al. [112] found that the flavonoids (e.g., glabrol, licochalcone A, licochalcone C, and licochalcone E) from licorice showed high efficiency against MRSA and displayed low cytotoxicity to mammalian cells. Among these flavonoids, glabrol exerts a bactericidal effect by increasing the permeability of the cell membrane and collapsing the proton motive force. Zhang et al. [159] also found that flavonoid extracts from Coriolus versicolors showed strong antibacterial activity against Escherichia coli, Staphylococcus aureus, and Bacillus subtilis in an experimental study. Yuan et al. [160] revealed that the cell membrane is the main site of flavonoids acting on Gram-positive bacteria, which likely involves the damage to phospholipid bilayers, the inhibition of the respiratory chain, or the ATP synthesis.
Various flavonoids have been extracted and studied to determine their antifungal activity and are expected to be promising, efficient, economical, and harmless drugs for inhibiting fungal infections in humans [161]. Several mechanisms have been proposed to explain the antifungal effects of flavonoids, including the plasma membrane disruption [162], induction of mitochondrial dysfunction [163], inhibition of cell wall generation [164], the suppression of cell division, the inhibition of RNA/protein biosynthesis, and the inhibition of efflux-mediated pumping systems [165]. Baicalein and apigenin were used as antifungal agents as they regulate ROS species, decrease lipid peroxidation, and block membrane disruption [50,130]. Glabridin inhibited the biosynthesis of the main components of fungi cell walls (β-glucans and chitin) [114]. Quercetin can inhibit oxidative phosphorylation and suppress the production of ROS [138]. Apigenin can control the cell cycle of fungi, while myricetin, quercetin, kaempferol, naringenin, genistein, and luteolin suppress the biosynthesis of DNA, RNA, and protein [166]. Li et al. [167] revealed that some flavonoids (glcyrrhiza, glabra, isoflavones, and chalcone) induced significant bactericidal effects by regulating the expression level of phosphatidylserine decarboxylase, which caused a loss of mitochondrial membrane potential and cell membrane disruption. Furthermore, baicalein and wogonin extracted from Scutellaria roots could induce apoptosis-like programmed cell death through the overproduction of ROS. Baicalein displayed a strong antifungal activity against Trichophyton rubrum, Trichophyton mentagrophytes, Aspergillus fumigatus, and Candida albicans, and wogonin inhibited all fungi except Candida albicans.

5.5. Antiviral Action

Flavonoids are known to be effective antivirals as they block virus attachment, preventing viruses from entering host cells and interfering with the replication, transcription, and translation of virus genomic DNA [168]. Roschek et al. [169] revealed that flavonoids can attach themselves to the surface proteins of viruses, inhibit viruses from entering the host cells, modulate the immune system, and reduce viral load. In vitro and in vivo studies have revealed that apigenin exhibits a wide range of antiviral effects against RNA and DNA viruses, such as herpes simplex virus type 1 and type 2, African swine fever virus, hepatitis B virus, and hepatitis C virus [131]. For example, baicalein can block the replication of the avian influenza H5N1 virus in humans [170]. Luteolin can inhibit HIV-1 reactivation, and genistein suppresses HIV-1 infection in CD4 + T cells and macrophages [135]. Kaempferol can block the HIV-1 replication in host cells and prevent herpes simplex virus types 1 and 2 from entering them [139]. Badshah et al. [171] comprehensively reviewed many flavonoids that showed antiviral activities in different testing environments, such as in vitro, in vivo (mouse model), and in silico.
The anti-inflammatory activity of flavonoids also influences cancer, carcinogen inactivation, anti-proliferation, cell cycle arrest, the induction of apoptosis, and the inhibition of angiogenesis [172]. Epidemiological studies revealed that dietary flavonoid intake could decrease the risk of breast, lung, colon, prostate, and pancreas tumors [173]. Several flavonoid compounds, including quercetin, rutin, hesperetin, silymarin, xanthohumol, 7,7′-dimethoxyagastisflavone, chrysoplenetin and chrysophanol D, formononetin, genistein, cyaniding and peonidin isolated from many medicinal plants, exert genoprotective, cytotoxic, anti-proliferative and/or proapoptotic actions in different tumoural cell lines [126,127]. Hirchaud et al. revealed that liquiritigenin isolated from licorice promoted apoptosis in HeLa cells by increasing the p53 and Bax gene expression, decreasing Bcl-2 gene expression, releasing cytochrome c, and elevating the activity of caspase-9 and -3 [128]. Flavonoids inhibit angiogenesis and metastasis by regulating the expression levels of VEGF and TGF-b1 genes [131]. In addition, flavonoids isolated from Dimorphandra mollis and Croton betulaster inhibited the proliferation of human glioblastoma cells and reduced the expression levels of VEGF and TGF-b1 genes [174]. Lewis et al. found that the therapy combining SAHA and EGCG not only significantly reduced the expression levels of miR-221/222 but also increased p27 and ER in carcinogenic α cells and the expression level of tumor suppressor genes [175]. Moreover, flavonoids have the ability to regulate non-coding microRNAs (miRNAs). The ability to alter miRNA levels through different mechanisms, either by inhibiting oncogenic miRNAs or activating tumor-suppressive miRNAs or by affecting transcription and epigenetic miRNA processing in TNBC [176].
There are two flavonoid compounds, isorhamnetin, and acacetin, that suppress the proliferation of human breast cancer cells [125]. The isoflavone genistein promotes breast cancer cell arrest at the G2/M phase and subsequent ROS-dependent apoptosis. Kaempferol has been shown to induce apoptosis via cell cycle arrest in human breast cancer MDA-MB-453 cells [142]. Genistein acts as a potential anticancer agent against multiple cancers such as breast cancer, prostate cancer, and ovarian cancer by inducing cell cycle G2/M phase arrest [143]. Naringenin can inhibit ROS formation, improve antioxidant enzyme activity such as SOD, CAT, and GSH, and further inhibit the proliferation and migration of MG-63 human osteosarcoma cells [119]. Hesperidin suppresses MG-63 cell cycle progression and activates the apoptosis of cancer cells [121]. Epigallocatechin-3-gallate (ECGG) causes growth arrest and cell death in prostate cancer cells [177]. In in vitro research, quercetin caused cell_cycle arrest, DNA damage, and cell growth inhibition in several cancer cell lines in vitro research, such as leukemia cancer, colon cancer, breast cancer, and ovarian cancer [115]. Apigenin and luteolin cause alterations to ROS signaling and induce apoptosis in several ovarian cancer cell lines, such as A2780, OVCAR-3, and SKOV-3 [132]. Cyanidin blocks cancer cell proliferation and induces apoptosis in the human epithelial colorectal adenocarcinoma cell line (Caco-2) [108]. In one study, resveratrol combined with grape seed proanthocyanidins significantly reduced the DNMT and HDAC activities of breast cancer cell lines [178]. In addition, resveratrol and procyanidins can synergistically inhibit the growth of breast cancer cells [179]. Flavonoids and other polyphenols act as epigenetic modifiers in breast cancer. Moreover, Mirza et al. reported that EGCG, genistein, curcumin, resveratrol, lutein A and gingival sterone exhibit cancer prevention effects through the epigenetic regulation of tumor suppressor genes [109].

6. Applications of Flavonoids in Cosmetics and Foods

6.1. Applications of Flavonoids in Cosmetics

Oxidative stress can cause changes in skin pigmentation changes and facial aging [180]. Thus, antioxidant supplements are considered the best approach to both the prevention and treatment of the above-mentioned issues [181]. Flavonoids are widely used in common cosmetics primarily due to their antioxidant and soothing properties [182]. The cosmetic applications of flavonoids comprise three aspects: sun protection, antiaging effects, and anti-inflammatory effects [183]. (1) Sun protection. The chromophores of flavonoids can increase light absorption, absorb light in the UV/blue spectral region, and reduce oxidative stress damage from sunlight [184]. Silymarin, a mixture of flavonolignans, has antioxidant, anti-inflammatory, and immunomodulatory properties and has led to the prevention of photocarcinogenesis in mice. Linarin, a flavone glycoside from Buddleja cardioids, was tested on guinea pigs and shown to have a sun protection factor (SPF) value of 9, thus providing remarkable protection from UV damage [185]. The quercetin and rutin from Moringa oleifera exhibited an SPF value of 2 [186]. The O/W formulation with anthocyanin from raspberry and blackberry obtained a high SPF value of 15.8 [187]. Anthocyanins activated transcription factor Nrf2 and induced the production of various antioxidant enzymes [188]. Finally, acacetin can prevent UVB-induced MMP-1 expression, which leads to skin photoaging, and may therefore have therapeutic potential as an anti-wrinkle agent to improve skin health [189]. (2) Antiaging effects. Flavonoids play important roles in protecting against the signs of aging, especially in the skin [190]. Micek et al. [191] reported that 3% taxifolin cream improved the viscoelasticity of aging skin with a better penetration rate. Water/ethanolic extracts from various parts of Nymphaea lotus are rich in flavonoids, which are usually used for homemade self-care products [189]. Water lily extracts contain excess flavonoids, which inhibit matrix metalloproteinase-1 and cause skin aging via extracellular matrix decompositions [192]. (3) Anti-inflammatory effects. Flavonoids inhibit phospholipase and cyclooxygenase activities and regulate immune cell migration and cytokines production [193]. A previous study revealed that the polymethoxyflavones from the Kaempferia parviflora inhibited pro-inflammatory mediators induced by TNF-α, such as cyclooxygenase-2 (COX-2), IL-6, interleukin (IL)-1β, mitogen-activated protein kinase (MAPK), nuclear factor-kappa B (NF-κB), and activator protein 1 (AP-1) [194,195]. Galangin, a specific flavonoid compound distributed in Alpinia galanga, has protective properties against human skin fibroblasts, which also exhibit an anti-inflammation effect by inhibiting NF-κB and increasing heme oxygenase-1 [189,196].

6.2. Application of Flavonoids in Foods

Food additives, derived from either plants or animals or produced via artificial synthesis, are added to foods and related products to enhance their flavor, taste, and freshness [197]. However, the excessive consumption of synthetic additives may cause potential risks to human health. Flavonoids, as the most important group of phenolics, are widely distributed in many vegetables and fruits and are used in food processing to decrease the use of synthetic chemicals and improve human health [198]. The content of flavonoids varies between plants and between organs of the same plant. Several horticultural crops, cabbage, carrots, spinach, mushrooms, peaches, strawberries, orange juice, and white wine, contain relatively low flavonoid content (less than 10 mg/kg) [198]. The content of flavonoids is lower than 50 mg/kg in lettuce, beans, red pepper, tomato, grapes, and tea and is higher than 50 mg/kg in broccoli, kale, French beans, celery, and cranberries [199]. The highest content is found in fruits such as berries (33.63 mg/100 g anthocyanins in strawberries and 13.52 mg/100 g anthocyanins in blueberries), apples (184 mg/200 g quercetin and 180 mg/200 g epicatechin), and citrus fruit (292 mg/500 mL hesperidin in orange juice) [198]. Moreover, flavonoids are used to extend the shelf life of many foods and preserve many foods due to their antimicrobial and antioxidant properties [200]. With the development of the economy and the improvements in people’s living standards, natural compounds have shown greater activity than synthetic chemicals because the body can accept natural compounds [201]. Flavonoids are used as food preservatives to prevent fat and oil oxidation, feed animals as supplements, protect various enzymes and vitamins, and inhibit microbial growth in foods [198,202]. A previous study revealed that flavonoids could inhibit lipid oxidation in red meats and poultry and retard spoilage due to microorganism growth in meats thanks to their antioxidant and antimicrobial properties [200,203].

7. Omics Research and Flavonoid Nanoparticles

Omics research attempts to comprehensively understand the biological molecules in an organism at a particular functional level, such as the genome, transcriptome, or proteome [204]. Complete genomic sequences and annotations provided important information, including promoter sequences, gene sequences, gene structure, and predicted gene functions. Recently, a combination of transcriptomic and metabolomic analyses has represented an effective approach to exploring the functions of genes associated with metabolism [205]. Yuan et al. showed that metabolome analysis detected 124 different flavonoid metabolites, and 30 different genes involved in flavonoid biosynthesis were identified through transcriptome analysis [206]. Integrated transcriptomic and metabolomic data revealed that anthocyanin biosynthetic genes exhibited the differential expression patterns between purple- and green-skinned fruit of Ficus carica [207]. Similarly, the accumulation of malvidin 3-O-glucoside and delphinidin 3-O-glucoside was associated with the reddening of the jujube peel, which may be correlated with an increase in the expression levels of three UFGT genes [208]. Combined fruit transcriptomic and metabolomic analyses of fruit have uncovered six candidate genes (AaF3H, AaLDOX, AaUFGT, AaMYB, AabHLH, and AaHB2) and seven flavonoid compounds that are closely associated with the pigmentation of red- and green-fleshed cultivars of Actinidia arguta [209]. Thus, the combination of transcriptomic (proteomics) and metabolomic analyses could be used to investigate the spatial distributions and dynamics of flavonoid metabolites and their corresponding spatiotemporal gene expression.
Although flavonoids are thought to have beneficial effects on human health, such as antioxidant, anti-inflammatory, and anticancer properties, the use of flavonoids in disease treatment is not satisfactory due to low solubility, poor absorption, and rapid metabolism [210]. Nanotechnology can solve the issue with flavonoids, namely, low water solubility, which plays a major role in low bioavailability. For example, fisetin, as a hydrophobic dietary flavone found in strawberries, apples, cucumbers, and onions, possessed low aqueous solubility (less than 1 mg/mL) and resulted in low bioavailability, and has limited its use [211]. Moreover, different types of flavonoid nanoparticles have different therapeutic effects on different diseases. Dobrzynska et al. summarized the effects of the different flavonoid nanoparticles on cancer therapy by enhancing their anti-tumor effects or reducing the systemic toxicity of drugs [212]. Siddiqui et al. showed that the polylactic acid-polyethylene glycol (PLA-PEG) with encapsulated EGCG nanoparticles were more than 10 times more effective against 22Rr1 prostate cancer cells than free EGCG [213]. Quercetin-loaded PEG nanoparticles were found to prolong the circulation time of quercetin in the bloodstream and increase its solubility and stability. Tan et al. revealed that the PEG-derivatized phosphatidylethanolamine nanomicelles increased the anticancer activity of quercetin and were more effective against A549 lung cancer cells than free-quercetin [214]. Genistein-loaded TPGS-b-PCL (d-α-tocopheryl polyethylene glycol 1000 succinate-poly(ε-caprolactone)) was found to inhibit HeLa cervical tumor cells growth and possessed a higher level of cytotoxicity in comparison with genistein-loaded PCL nanoparticles [212]. In addition, apigenin nanoparticles, including apigenin loaded in PLGA nanoparticles and apigenin encapsulated in PLGA nanoparticles, showed an anti-proliferative effect against A475 skin cancer cells and delayed the development of hepatocellular carcinoma in rats [212]. Nanonargenin works by inhibiting both the PI3K and MAPK paths and by restricting ER alpha to the cytoplasm to lessen the proliferation of Tam-RMC cells [215].

8. Conclusions

Flavonoids represent the main class of plant secondary metabolites and occur in the different tissues and organs of various plant species. The elucidation of their biosynthetic pathways, as well as their regulation by transcription factors and non-coding RNA, has enabled researchers to employ metabolic engineering to synthesize diverse flavonoids with valuable applications. Flavonoid compounds are classified into seven subgroups due to modifications to their basic skeletons, including flavones, flavanones, isoflavones, flavonols or catechins, and anthocyanins. Flavonoids, as potential candidates for bioactive compounds, are widely used in food processing and the cosmetics and pharmaceutical industries to improve human health due to their antioxidant and free radical scavenging properties. Moreover, the biological activities of flavonoid compounds, including their antioxidant effects, antimicrobial effects, anticancer effects, cardioprotective effects, anti-inflammatory effects, and skin protective effects, were addressed in this review. In addition, we summarized the mining of new functional genes metabolites through omics research, and the engineering of flavonoids using nanotechnology was also summarized in this review. This review provides a reference for basic and applied research on flavonoid compounds.

9. Future Perspectives

9.1. Mining of Functional Genes

Plant genomes and transcriptomes were widely used to investigate gene information, including gene locations, gene structure, and gene expression patterns. Integrated transcriptome (proteome) and metabolome analyses have revealed the dynamic changes in flavonoid compounds and the corresponding functional genes and transcription factors. The yeast one-hybrid and yeast two-hybrid systems are considered valuable and straightforward techniques for studying the interactions between transcription factors and the promotion of functional genes and transcription factors. DNA molecular markers and a genome-wide association study (GWAS) were used to identify the relationships between genotypes and phenotypes and explore the variation in genomic loci associated with the important agronomic traits and detected key genes.

9.2. Extraction and Utilization of Bioactive Ingredients

Sample extraction techniques severely block the isolation and extraction of individual flavonoid compounds in plants, which severely restricts the development of medicines. In addition, the pharmacological mechanisms of flavonoids and their derivatives in many plants are still unclear due to the lack of animal studies and clinical trials. With the innovations of new extraction techniques and the development of molecular biology, studies on flavonoids and their derivatives mainly focus on their isolation and extraction, metabolic pathway analysis, molecular regulatory mechanisms, and potential applications in pharmacy and health services research using nanotechnology.

Author Contributions

Designed the experiment, X.C. and S.C.; wrote the manuscript, S.C. and X.W.; revised the manuscript, S.C., H.G. and Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (32102306). China Postdoctoral Science Foundation (2021M702765).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. PRISMA flow chart presenting literature data selection.
Figure A1. PRISMA flow chart presenting literature data selection.
Molecules 28 04982 g0a1
Figure A2. The Basic Structure of Flavonoids.
Figure A2. The Basic Structure of Flavonoids.
Molecules 28 04982 g0a2
Table A1. The Biological Activity and Mechanism of Action of Flavonoids.
Table A1. The Biological Activity and Mechanism of Action of Flavonoids.
FlavonoidsBiological ActivityMechanism of ActionReferences
Proanthocyanidinsantioxidant activitypolymerizing flavonoid monomer and Catechin polymer[93]
Fisetinantioxidant activityinhibiting the activities of oxidase[10,96,97]
Hydroxyflavonesantioxidant activitygenerating complexes with metal cations[98]
Kaempferolantioxidant activitychelate Cu(II) ions[101]
Catecholantioxidant activitygenerating stable metallic complexes [102]
Quercetinantioxidant activityinteracting with α-tocopherol radicals[104,106,107]
Epigallocatechin
Gallate
Naringin
antioxidant activityinteracting with α-tocopherol radicals[104]
Apigeninantiinflammatory actionreducing the steady-state mRNA levels induced by TNF-α[145]
Quercetin
catechins
antiinflammatory actionreducing the expression of pro-inflammatory cytokines[113,133]
Kaempferol
Myricetin
antiinflammatory actioninhibit the activity of lipoxygenase[133]
Diosmin
Hesperidin
antiinflammatory actioninhibiting leukotriene B4 biosynthesis[122]
Chalconesantiinflammatory actioninhibiting proinflammatory cytokines activity[110]
Vitexin
Orientin
Rutin
antiinflammatory actioninhibiting iNOS and COX2 expression[123]
Naringenin
Hesperetin
anticardiovascular actionreducing high blood pressure by promoting vasodilation[116,117]
Quercetinanticardiovascular actionprotecting the heart against ischemia-reperfusion injury via the cardioprotective effects of quercetin.[116,117,137]
Isoflavonesanticardiovascular actioninhibiting monocyte-endothelial cell adhesion[153]
Anthocyaninsanticardiovascular actionimproving systolic blood pressure, and reduce the content of triglycerides, total cholesterol, and low-density lipoprotein cholesterol[154,155]
Xanthohumolanticardiovascular actionincreasing PTEN expression and inhibited AKT/mTOR phosphorylation[111]
Flavonoidanticardiovascular actionmodulating the PI3K/Akt/GSK3β pathway or downregulating the ROS-mediated JNK/p38MAPK/NFκB pathway[141,156]
the citrus flavonoid naringeninanticardiovascular actionenhancing SIRT1 expression, reduced ROS production[118]
Quercetinantibacterial actionblocking the biofilm formation and suppressed the DNA replication[8,138,158,166]
Apigenin
Naringenin
Chrysin
Genistein
Kaempferol
Daidzin
Antibacterial
action
blocking the biofilm formation or regulate ROS species, decrease lipid peroxidation[8,129,130,158,166]
Baicalein
Luteolin
Myricetin
antibacterial actionsuppressing the DNA replication or suppressed the biosynthesis of ATP[8,129,130,158]
Epigallocatechin gallateantibacterial actionsuppressing the biosynthesis of ATP[114]
Glabrolantibacterial actionincreasing the permeability of the cell membrane and collapsing the proton motive force[112]
Glabridinantifungal actioninhibiting the biosynthesis of the main components of fungi cell walls[114]
Myricetin
Kaempferol
Naringenin
Genistein
Luteolin
antifungal action
suppressing the biosynthesis of DNA, RNA, and protein[166]
Glcyrrhiza
Glabra
Isoflavones
Chalcone
antifungal actionregulating the expression level of phosphatidylserine decarboxylase[167]
Apigeninantiviral actioninhibiting viruses from entering the host cells, modulate the immune system, and reduce viral load[131,169]
Baicaleinantiviral actionblocking the replication of the avian influenza H5N1 virus[170]
Luteolinantiviral actioninhibiting HIV-1 reactivation[135]
Kaempferolantiviral actionblocking the HIV-1 replication[139]
Liquiritigeninpromoted apoptosisincreasing the p53 and Bax gene expression, decreasing Bcl-2 gene expression [128]
Isoflflavone genisteinsuppress the proliferationpromoting breast cancer cell arrest at the G2/M phase and subsequent ROS-dependent apoptosis[125,143]
Kaempferolsuppress the proliferationinducing apoptosis via cell cycle arrest MDA-MB-453 cells[125,142]
Naringenininhibit the proliferation and migrationinhibiting ROS formation[119]
Apigenin
Luteolin
induce apoptosiscausing the alterations to ROS signaling[132]

References

  1. Santos, E.L.; Maia, B.; Ferriani, A.P.; Teixeira, S.D. Flavonoids: Classification, biosynthesis and chemical ecology. In Flavonoids: From Biosynthesis to Human Health; IntechOpen: London, UK, 2017; Volume 13, pp. 78–94. [Google Scholar] [CrossRef] [Green Version]
  2. Ullah, A.; Munir, S.; Badshah, S.L.; Khan, N.; Ghani, L.; Poulson, B.G.; Emwas, A.-H.; Jaremko, M. Important Flavonoids and Their Role as a Therapeutic Agent. Molecules 2020, 25, 5243. [Google Scholar] [CrossRef]
  3. García-Lafuente, A.; Guillamón, E.; Villares, A.; Rostagno, M.A.; Martínez, J.A. Flavonoids as anti-inflammatory agents: Implications in cancer and cardiovascular disease. Inflamm. Res. 2009, 58, 537–552. [Google Scholar] [CrossRef] [PubMed]
  4. Karak, P. Biological activities of flavonoids: An overview. Int. J. Pharm. Sci. Res. 2019, 10, 1567–1574. [Google Scholar] [CrossRef]
  5. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. do Nascimento, R.P.; dos Santos, B.L.; Amparo, J.A.O.; Soares, J.R.P.; da Silva, K.C.; Santana, M.R.; Almeida, Á.M.A.N.; da Silva, V.D.A.; Costa, M.d.F.D.; Ulrich, H.; et al. Neuroimmunomodulatory Properties of Flavonoids and Derivates: A Potential Action as Adjuvants for the Treatment of Glioblastoma. Pharmaceutics 2022, 14, 116. [Google Scholar] [CrossRef] [PubMed]
  7. Buer, C.S.; Imin, N.; Djordjevic, M.A. Flavonoids: New roles for old molecules. J. Integr. Plant. Biol. 2010, 52, 98–111. [Google Scholar] [CrossRef]
  8. Dias, M.C.; Pinto, D.C.G.A.; Silva, A.M.S. Plant flavonoids: Chemical characteristics and biological activity. Molecules 2021, 26, 5377. [Google Scholar] [CrossRef]
  9. Vicente, O.; Boscaiu, M. Flavonoids: Antioxidant compounds for plant defence and for a healthy human diet. Not. Bot. Horti Agrobot. Cluj. 2018, 46, 14–21. [Google Scholar] [CrossRef] [Green Version]
  10. Zhuang, W.-B.; Li, Y.-H.; Shu, X.-C.; Pu, Y.-T.; Wang, X.-J.; Wang, T.; Wang, Z. The Classification, Molecular Structure and Biological Biosynthesis of Flavonoids, and Their Roles in Biotic and Abiotic Stresses. Molecules 2023, 28, 3599. [Google Scholar] [CrossRef]
  11. Pietta, P.; Minoggio, M.; Bramati, L. Plant polyphenols: Structure, occurrence and bioactivity. Stud. Nat. Prod. Chem. 2003, 28, 257–312. [Google Scholar] [CrossRef]
  12. Jan, R.; Asaf, S.; Numan, M.; Lubna Kim, K.M. Plant secondary metabolite biosynthesis and transcriptional regulation in response to biotic and abiotic stress conditions. Agronomy 2021, 11, 968. [Google Scholar] [CrossRef]
  13. Sircaik, S.; Dhiman, K.; Gambhir, G.; Kumar, P.; Srivastava, D.K. Transgenic Implications for Biotic and Abiotic Stress Tolerance in Agricultural Crops. In Agricultural Biotechnology: Latest Research and Trends; Springer: Berlin/Heidelberg, Germany, 2021; pp. 185–221. [Google Scholar] [CrossRef]
  14. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. Int. J. Surg. 2021, 88, 105906. [Google Scholar] [CrossRef]
  15. Khuntia, A.; Martorell, M.; Ilango, K.; Bungau, S.G.; Radu, A.F.; Behl, T.; Sharifi-Rad, J. Theoretical evaluation of Cleome species’ bioactive compounds and therapeutic potential: A literature review. Biomed. Pharmacother. 2022, 151, 113161. [Google Scholar] [CrossRef] [PubMed]
  16. Hussain, T.; Tan, B.; Murtaza, G.; Liu, G.; Yin, Y. Flavonoids and type 2 diabetes: Evidence of efficacy in clinical and animal studies and delivery strategies to enhance their therapeutic efficacy. Pharm. Res. 2020, 152, 104629. [Google Scholar] [CrossRef] [PubMed]
  17. Rehan, M. Biosynthesis of Diverse Class Flavonoids via Shikimate and Phenylpropanoid Pathway. In Bioactive Compounds-Biosynthesis, Characterization and Applications; IntechOpen: London, UK, 2021; p. 75392. [Google Scholar]
  18. Zakaryan, H.; Arabyan, E.; Oo, A.; Zandi, K. Flavonoids: Promising natural compounds against viral infections. Arch. Virol. 2017, 162, 2539–2551. [Google Scholar] [CrossRef] [PubMed]
  19. Luca, S.V.; Macovei, I.; Bujor, A.; Miron, A.; Skalicka-Woźniak, K.; Aprotosoaie, A.C.; Trifan, A. Bioactivity of dietary polyphenols: The role of metabolites. Crit. Rev. Food Sci. 2020, 60, 626–659. [Google Scholar] [CrossRef] [PubMed]
  20. Bose, S.; Sarkar, D.; Bose, A.; Mandal, S.C. Natural flavonoids and its pharmaceutical importance. Pharma Rev. 2018, 94, 61–75. [Google Scholar]
  21. Tang, D.; Chen, K.; Huang, L.; Li, J. Pharmacokinetic properties and drug interactions of apigenin, a natural flavone. Expert. Opin. Drug Metab. Toxi. 2017, 13, 323–330. [Google Scholar] [CrossRef]
  22. Ginwala, R.; Bhavsar, R.; Chigbu, D.G.I.; Jain, P.; Khan, Z.K. Potential role of flavonoids in treating chronic inflammatory diseases with a special focus on the anti-inflammatory activity of apigenin. Antioxidants 2019, 8, 35. [Google Scholar] [CrossRef] [Green Version]
  23. Proteggente, A.R.; Pannala, A.S.; Paganga, G.; Buren, L.V.; Wagner, E.; Wiseman, S.; Van De Put, F.; Dacombe, C.; Rice-Evans, C.A. The antioxidant activity of regularly consumed fruit and vegetables reflects their phenolic and vitamin C composition. Free Radic. Res. 2002, 36, 217–233. [Google Scholar] [CrossRef]
  24. Martínez-Lüscher, J.; Brillante, L.; Kurtural, S.K. Flavonol profile is a reliable indicator to assess canopy architecture and the exposure of red wine grapes to solar radiation. Front. Plant. Sci. 2019, 10, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ali, A.; Parisi, A.; Normanno, G. Polyphenols as emerging antimicrobial agents. In Emerging Modalities in Mitigation of Antimicrobial Resistance; Springer International Publishing: Cham, Switzerland, 2022; pp. 219–259. [Google Scholar] [CrossRef]
  26. Vandercook, C.E.; Stephenson, R.G. Lemon juice composition. Identification of major phenolic compounds and estimation by paper chromatography. J. Agr. Food Chem. 1966, 14, 450–454. [Google Scholar] [CrossRef]
  27. Wang, Y.; Liu, X.J.; Chen, J.B.; Cao, J.P.; Sun, C.D. Citrus flavonoids and their antioxidant evaluation. Crit. Rev. Food Sci. 2022, 62, 3833–3854. [Google Scholar] [CrossRef] [PubMed]
  28. Najmanová, I.; Vopršalová, M.; Saso, L.; Mladěnka, P. The pharmacokinetics of flavanones. Crit. Rev. Food Sci. 2020, 60, 3155–3171. [Google Scholar] [CrossRef] [PubMed]
  29. Roy, M.; Datta, A. Fundamentals of phytochemicals. In Cancer Genetics and Therapeutics; Springer: Singapore, 2019; pp. 49–81. [Google Scholar] [CrossRef]
  30. Galleano, M.; Calabro, V.; Prince, P.D.; Litterio, M.C.; Piotrkowski, B.; Vazquez-Prieto, M.A.; Miatello, R.M.; Oteiza, P.I.; Fraga, C.G. Flavonoids and metabolic syndrome. Ann. Acad. Sci. 2012, 1259, 87–94. [Google Scholar] [CrossRef] [PubMed]
  31. Liu, Y.; Qian, J.; Li, J.; Xing, M.; Grierson, D.; Sun, C.; Xu, C.; Li, X.; Chen, K. Hydroxylation decoration patterns of flavonoids in horticultural crops: Chemistry, bioactivity, and biosynthesis. Hortic. Res. 2022, 9, uhab068. [Google Scholar] [CrossRef]
  32. Shah, A.; Smith, D.L. Flavonoids in agriculture: Chemistry and roles in, biotic and abiotic stress responses, and microbial associations. Agronomy 2020, 10, 1209. [Google Scholar] [CrossRef]
  33. Kopečná-Zapletalová, M.; Krasulová, K.; Anzenbacher, P.; Hodek, P.; Anzenbacherová, E. Interaction of isoflavonoids with human liver microsomal cytochromes P450: Inhibition of CYP enzyme activities. Xenobiotica 2017, 47, 324–331. [Google Scholar] [CrossRef]
  34. Hummelova, J.; Rondevaldova, J.; Balastikova, A.; Lapcik, O.; Kokoska, L. The relationship between structure and in vitro antibacterial activity of selected isoflavones and their metabolites with special focus on antistaphylococcal effect of demethyltexasin. Lett. Appl. Microbiol. 2015, 60, 242–247. [Google Scholar] [CrossRef]
  35. Lim, Y.J.; Jeong, H.Y.; Gil, C.S.; Kwon, S.J.; Na, J.K.; Lee, C.; Eom, S.H. Isoflavone accumulation and the metabolic gene expression in response to persistent UV-B irradiation in soybean sprouts. Food Chem. 2020, 303, 125376. [Google Scholar] [CrossRef]
  36. Meng, N.; Yu, B.J.; Guo, J.S. Ameliorative effects of inoculation with Bradyrhizobium japonicum on Glycine max and Glycine soja seedlings under salt stress. Plant. Growth Regul. 2016, 80, 137–147. [Google Scholar] [CrossRef]
  37. Mori-Yasumoto, K.; Hashimoto, Y.; Agatsuma, Y.; Fuchino, H.; Yasumoto, K.; Shirota, O.; Satake, M.; Sekita, S. Leishmanicidal phenolic compounds derived from Dalbergia cultrata. Nat. Prod. Res. 2021, 35, 4907–4915. [Google Scholar] [CrossRef]
  38. Dutta, M.S.; Mahapatra, P.; Ghosh, A.; Basu, S. Estimation of the reducing power and electrochemical behavior of few flavonoids and polyhydroxybenzophenones substantiated by bond dissociation energy: A comparative analysis. Mol. Divers. 2022, 26, 1101–1113. [Google Scholar] [CrossRef] [PubMed]
  39. Pico, J.; Xu, K.; Guo, M.; Mohamedshah, Z.; Ferruzzi, M.G.; Martinez, M.M. Manufacturing the ultimate green banana flour: Impact of drying and extrusion on phenolic profile and starch bioaccessibility. Food Chem. 2019, 297, 124990. [Google Scholar] [CrossRef]
  40. Ding, T.; Cao, K.; Fang, W.; Zhu, G.; Chen, C.; Wang, X.; Wang, L. Evaluation of phenolic components (anthocyanins, flavanols, phenolic acids, and flavonols) and their antioxidant properties of peach fruits. Sci. Hortic. 2020, 268, 109365. [Google Scholar] [CrossRef]
  41. Yang, S.; Mi, L.; Wu, J.; Liao, X.; Xu, Z. Strategy for anthocyanins production: From efficient green extraction to novel microbial biosynthesis. Crit. Rev. Food Sci. 2022, 1–16. [Google Scholar] [CrossRef] [PubMed]
  42. Guven, H.; Arici, A.; Simsek, O. Flavonoids in our foods: A short review. J. Basic. Clin. Health Sci. 2019, 3, 96–106. [Google Scholar] [CrossRef]
  43. Di Pietro, N.; Baldassarre, M.P.A.; Cichelli, A.; Pandolfi, A.; Formoso, G.; Pipino, C. Role of polyphenols and carotenoids in endothelial dysfunction: An overview from classic to innovative biomarkers. Oxid. Med. Cell. Longev. 2020, 2020, 6381380. [Google Scholar] [CrossRef]
  44. Fang, J. Classification of fruits based on anthocyanin types and relevance to their health effects. Nutrition 2015, 31, 1301–1306. [Google Scholar] [CrossRef]
  45. 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] [Green Version]
  46. Chaves-Silva, S.; Dos Santos, A.L.; Chalfun-Júnior, A.; Peres, L.E.P.; Zhao, J.; Benedito, V.A. Understanding the genetic regulation of anthocyanin biosynthesis in plants–tools for breeding purple varieties of fruits and vegetables. Phytochemistry 2018, 153, 11–27. [Google Scholar] [CrossRef]
  47. Tohge, T.; de Souza, L.P.; Fernie, A.R. Current understanding of the pathways of flavonoid biosynthesis in model and crop plants. J. Exp. Bot. 2017, 68, 4013–4028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Kim, J.I.; Hidalgo-Shrestha, C.; Bonawitz, N.D.; Franke, R.B.; Clint, C. Spatio-temporal control of phenylpropanoid biosynthesis by inducible complementation of a cinnamate 4-hydroxylase mutant. J. Exp. Bot. 2021, 72, 3061–3073. [Google Scholar] [CrossRef]
  49. Shen, N.; Wang, T.; Gan, Q.; Liu, S.; Wang, L.; Jin, B. Plant flavonoids: Classification, distribution, biosynthesis, and antioxidant activity. Food Chem. 2022, 383, 132531. [Google Scholar] [CrossRef]
  50. Gonzali, S.; Perata, P. Anthocyanins from purple tomatoes as novel antioxidants to promote human health. Antioxidants 2020, 9, 1017. [Google Scholar] [CrossRef]
  51. Goyal, K.; Kaur, R.; Goyal, A.; Awasthi, R. Chalcones: A review on synthesis and pharmacological activities. J. Appl. Pharm. Sci. 2021, 11, 001–014. [Google Scholar] [CrossRef]
  52. Jasim, H.A.; Nahar, L.; Jasim, M.A.; Moore, S.A.; Ritchie, K.J.; Sarker, S.D. Chalcones: Synthetic chemistry follows where nature leads. Biomolecules 2021, 11, 1203. [Google Scholar] [CrossRef]
  53. Li, M.; Guo, L.; Wang, Y.; Li, Y.; Jiang, X.; Liu, Y.; Xie, D.; Gao, L.; Xia, T. Molecular and biochemical characterization of two 4-coumarate: CoA ligase genes in tea plant (Camellia sinensis). Plant Mol. Biol. 2022, 109, 579–593. [Google Scholar] [CrossRef]
  54. Sun, T.; Li, S.; Song, X.; Pei, G.; Diao, J.; Cui, J.; Shi, M.; Chen, L.; Zhang, W. Re-direction of carbon flux to key precursor malonyl-CoA via artificial small RNAs in photosynthetic Synechocystis sp. PCC 6803. Biotechnol. Biofuels. 2018, 11, 1–16. [Google Scholar] [CrossRef] [Green Version]
  55. Zhao, C.; Liu, X.; Gong, Q.; Cao, J.; Shen, W.; Yin, X.; Grierson, D.; Zhang, B.; Xu, C.; Li, X.; et al. Three AP2/ERF family members modulate flavonoid synthesis by regulating type IV chalcone isomerase in citrus. Plant Biotechnol. J. 2021, 19, 671–688. [Google Scholar] [CrossRef] [PubMed]
  56. Cheng, A.X.; Han, X.J.; Wu, Y.F.; Lou, H.X. The function and catalysis of 2-oxoglutarate-dependent oxygenases involved in plant flavonoid biosynthesis. Int. J. Mol. Sci. 2014, 15, 1080–1095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Dastmalchi, M.; Dhaubhadel, S. Soybean chalcone isomerase: Evolution of the fold, and the differential expression and localization of the gene family. Planta 2015, 241, 507–523. [Google Scholar] [CrossRef] [PubMed]
  58. Richter, A.S. Tansley insight. New Phytol. 2022, 236, 2037–2043. [Google Scholar] [CrossRef]
  59. Baba, S.A.; Ashraf, N. Functional characterization of flavonoid 3′-hydroxylase, CsF3′ H, from Crocus sativus L: Insights into substrate specificity and role in abiotic stress. Arch. Biochem. Biophys. 2019, 667, 70–78. [Google Scholar] [CrossRef]
  60. Li, H.; Tian, J.; Yao, Y.; Zhang, J.; Song, T.; Li, K.; Yao, Y. Identification of leucoanthocyanidin reductase and anthocyanidin reductase genes involved in proanthocyanidin biosynthesis in Malus crabapple plants. Plant Physiol. Biochem. 2019, 139, 141–151. [Google Scholar] [CrossRef]
  61. Ma, S.; Hu, R.; Ma, J.; Fan, J.; Wu, F.; Wang, Y.; Huang, L.; Feng, G.; Li, D.; Nie, G.; et al. Integrative analysis of the metabolome and transcriptome provides insights into the mechanisms of anthocyanins and proanthocyanidins biosynthesis in Trifolium repens. Ind. Crop. Prod. 2022, 187, 115529. [Google Scholar] [CrossRef]
  62. Rauf, A.; Imran, M.; Abu-Izneid, T.; Haq, I.U.; Pate, S.; Pan, X.; Naz, S.; Silva, A.S.; Saeed, F.; Suleria, H.A.R. Proanthocyanidins: A comprehensive review. Biomed. Pharmacother. 2019, 116, 108999. [Google Scholar] [CrossRef]
  63. Ni, J.; Zhao, Y.; Tao, R.; Yin, L.; Gao, L.; Strid, Å.; Qian, M.; Li, J.; Li, Y.; Shen, J.; et al. Ethylene mediates the branching of the jasmonate-induced flavonoid biosynthesis pathway by suppressing anthocyanin biosynthesis in red Chinese pear fruits. Plant Biotechnol. J. 2020, 18, 1223–1240. [Google Scholar] [CrossRef] [Green Version]
  64. Kim, M.J.; Paramanantham, A.; Lee, W.S.; Yun, J.W.; Chang, S.H.; Kim, D.C.; Park, H.S.; Choi, Y.H.; Kim, G.S.; Ryu, C.H.; et al. Anthocyanins Derived from Vitis coignetiae Pulliat Contributes Anti-Cancer Effects by Suppressing NF-κB Pathways in Hep3B Human Hepatocellular Carcinoma Cells and In Vivo. Molecules 2020, 25, 5445. [Google Scholar] [CrossRef]
  65. Fenn, M.A.; Giovannoni, J.J. Phytohormones in fruit development and maturation. Plant J. 2021, 105, 446–458. [Google Scholar] [CrossRef]
  66. Corso, M.; Perreau, F.; Mouille, G.; Lepiniec, L. Specialized phenolic compounds in seeds: Structures, functions, and regulations. Plant Sci. 2020, 296, 110471. [Google Scholar] [CrossRef] [PubMed]
  67. Li, K.; Xing, C.; Yao, Z.; Huang, X. Pbr MYB 21, a novel MYB protein of Pyrus betulaefolia, functions in drought tolerance and modulates polyamine levels by regulating arginine decarboxylase gene. Plant Biotechnol. J. 2017, 15, 1186–1203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Qin, W.; Xie, L.; Li, Y.; Liu, H.; Fu, X.; Chen, T.; Hassani, D.; Li, L.; Sun, X.; Tang, K. An R2R3-MYB transcription factor positively regulates the glandular secretory trichome initiation in Artemisia annua L. Front. Plant Sci. 2021, 12, 657156. [Google Scholar] [CrossRef] [PubMed]
  69. Anwar, M.; Yu, W.; Yao, H.; Zhou, P.; Allan, A.C.; Zeng, L. NtMYB3, an R2R3-MYB from narcissus, regulates flavonoid biosynthesis. Int. J. Mol. Sci. 2019, 20, 5456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Xu, F.; Ning, Y.; Zhang, W.; Liao, Y.; Li, L.; Cheng, H.; Cheng, S. An R2R3-MYB transcription factor as a negative regulator of the flavonoid biosynthesis pathway in Ginkgo biloba. Funct. Integr. Genom. 2014, 14, 177–189. [Google Scholar] [CrossRef] [PubMed]
  71. Yan, J.; Wang, B.; Zhong, Y.; Yao, L.; Cheng, L.; Wu, T. The soybean R2R3 MYB transcription factor GmMYB100 negatively regulates plant flavonoid biosynthesis. Plant Mol. Biol. 2015, 89, 35–48. [Google Scholar] [CrossRef]
  72. Premathilake, A.T.; Ni, J.; Bai, S.; Tao, R.; Teng, Y. R2R3-MYB transcription factor PpMYB17 positively regulates flavonoid biosynthesis in pear fruit. Planta 2020, 252, 1–16. [Google Scholar] [CrossRef]
  73. Sun, Z.; Linghu, B.; Hou, S.; Liu, R.; Wang, L.; Hao, Y.; Han, Y.; Zhou, M.; Liu, L.; Li, H. Tartary buckwheat FtMYB31 gene encoding an R2R3-MYB transcription factor enhances flavonoid accumulation in Tobacco. J. Plant Growth Regul. 2020, 39, 564–574. [Google Scholar] [CrossRef]
  74. Cultrone, A.; Cotroneo, P.S.; Reforgiato Recupero, G. Cloning and molecular characterization of R2R3-MYB and bHLH-MYC transcription factors from Citrus sinensis. Tree Genet. Genomes 2010, 6, 101–112. [Google Scholar] [CrossRef]
  75. Xie, X.B.; Li, S.; Zhang, R.F.; Zhao, J.; Chen, Y.C.; Zhao, Q.; Yao, Y.X.; You, C.X.; Zhang, X.S.; Hao, Y.J. The bHLH transcription factor MdbHLH3 promotes anthocyanin accumulation and fruit colouration in response to low temperature in apples. Plant Cell Environ. 2012, 35, 1884–1897. [Google Scholar] [CrossRef]
  76. Qi, Y.; Zhou, L.; Han, L.; Zou, H.; Wang, Y. PsbHLH1, a novel transcription factor involved in regulating anthocyanin biosynthesis in tree peony (Paeonia suffruticosa). Plant Physiol. Biochem. 2020, 154, 396–408. [Google Scholar] [CrossRef]
  77. Mano, H.; Ogasawara, F.; Sato, K.; Higo, H.; Minobe, Y. Isolation of a regulatory gene of anthocyanin biosynthesis in tuberous roots of purple-fleshed sweet potato. Plant Physiol. 2007, 143, 1252–1268. [Google Scholar] [CrossRef] [Green Version]
  78. Peniche-Pavía, H.A.; Guzmán, T.J.; Magaña-Cerino, J.M.; Gurrola-Díaz, C.M.; Tiessen, A. Maize Flavonoid Biosynthesis, Regulation, and Human Health Relevance: A Review. Molecules 2022, 27, 5166. [Google Scholar] [CrossRef]
  79. Qian, Y.; Zhang, T.; Yu, Y.; Gou, L.; Pi, E. Regulatory mechanisms of bHLH transcription factors in plant adaptive responses to various abiotic stresses. Front. Plant Sci. 2021, 12, 1143. [Google Scholar] [CrossRef]
  80. Qiu, Z.; Wang, H.; Li, D.; Byu, B.; Cao, B. Identification of Candidate HY5-Dependent and-Independent Regulators of Anthocyanin Biosynthesis in Tomato. Plant Cell Physiol. 2019, 60, 643–656. [Google Scholar] [CrossRef]
  81. Malacarne, G.; Coller, E.; Czemmel, S.; Vrhovsek, U.; Engelen, K.; Goremykin, V.; Bogs, J.; Moser, C. The grapevine VvibZIPC22 transcription factor is involved in the regulation of flavonoid biosynthesis. J. Exp. Bot. 2016, 67, 3509–3522. [Google Scholar] [CrossRef] [Green Version]
  82. Singh, D.; Singh, H.; Singh, N.; Dwivedi, S.; Trivedi, P.K. Tobacco HY5, NtHY5, positively regulates flavonoid biosynthesis and enhances salt stress tolerance. bioRxiv 2022. [Google Scholar] [CrossRef]
  83. Morishita, T.; Kojima, Y.; Maruta, T.; Nishizawa-Yokoi, A.; Yabuta, Y.; Shigeoka, S. Arabidopsis NAC transcription factor, ANAC078, regulates flavonoid biosynthesis under high-light. Plant Cell Physiol. 2009, 50, 2210–2222. [Google Scholar] [CrossRef] [Green Version]
  84. An, X.H.; Tian, Y.; Chen, K.Q.; Liu, X.J.; Liu, D.D.; Xie, X.B.; Cheng, C.G.; Cong, P.H.; Hao, Y.J. MdMYB9 and MdMYB11 are involved in the regulation of the JA-induced biosynthesis of anthocyanin and proanthocyanidin in apples. Plant Cell Physiol. 2015, 56, 650–662. [Google Scholar] [CrossRef] [Green Version]
  85. Skirycz, A.; Jozefczuk, S.; Stobiecki, M.; Muth, D.; Zanor, M.I.; Witt, I.; Mueller-Roeber, B. Transcription factor AtDOF4; 2 affects phenylpropanoid metabolism in Arabidopsis thaliana. New Phytol. 2007, 175, 425–438. [Google Scholar] [CrossRef]
  86. Wang, N.; Liu, W.; Zhang, T.; Jiang, S.; Xu, H.; Wang, Y.; Zhang, Z.; Wang, C.; Chen, X. Transcriptomic analysis of red-fleshed apples reveals the novel role of MdWRKY11 in flavonoid and anthocyanin biosynthesis. J. Agr. Food Chem. 2018, 66, 7076–7086. [Google Scholar] [CrossRef]
  87. Liu, S.; Wang, L.; Cao, M.; Pang, S.; Wang, L. Identification and characterization of long non-coding RNAs regulating flavonoid biosynthesis in Ginkgo biloba leaves. Ind. Crop. Prod. 2020, 158, 112980. [Google Scholar] [CrossRef]
  88. Ma, X.; Zhang, X.; Traore, S.M.; Xin, Z.; Yin, D. Genome-wide identification and analysis of long noncoding RNAs (lncRNAs) during seed development in peanut (Arachis hypogaea L.). BMC Plant Biol. 2020, 20, 1–14. [Google Scholar] [CrossRef]
  89. Gupta, O.P.; Karkute, S.G.; Banerjee, S.; Meena, N.L.; Dahuja, A. Contemporary understanding of miRNA-based regulation of secondary metabolites biosynthesis in plants. Front. Plant Sci. 2017, 8, 374. [Google Scholar] [CrossRef] [Green Version]
  90. Sharma, D.; Tiwari, M.; Pandey, A.; Bhatia, C.; Sharma, A.; Trivedi, P.K. MicroRNA858 is a potential regulator of phenylpropanoid pathway and plant development. Plant Physiol. 2016, 171, 944–959. [Google Scholar] [CrossRef] [Green Version]
  91. Glevitzky, I.; Dumitrel, G.A.; Glevitzky, M.; Manuela, B.P. Statistical analysis of the relationship between antioxidant activity and the structure of flavonoid compounds. Rev. Chim. 2019, 70, 3103–3107. [Google Scholar] [CrossRef]
  92. Zuo, A.R.; Dong, H.H.; Yu, Y.Y.; Shu, Q.L.; Zheng, L.X.; Yu, X.Y.; Cao, S.W. The antityrosinase and antioxidant activities of flavonoids dominated by the number and location of phenolic hydroxyl groups. Chin. Med. 2018, 13, 1–12. [Google Scholar] [CrossRef] [Green Version]
  93. Kirschweng, B.; Tátraaljai, D.; Földes, E.; Pukánszky, B. Natural antioxidants as stabilizers for polymers. Polym. Degrad. Stabil. 2017, 145, 25–40. [Google Scholar] [CrossRef] [Green Version]
  94. Engwa, G.A. Free Radicals and the Role of Plant Phytochemicals as Antioxidants Against Oxidative Stress-Related Diseases. In Phytochemicals-Source of Antioxidants and Role in Disease Prevention; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef] [Green Version]
  95. Lee-Hilz, Y.Y.; Boerboom, A.M.; Westphal, A.H.; Berkel, W.J.; Aarts, J.M.; Rietjens, I.M. Pro-oxidant activity of flavonoids induces EpRE-mediated gene expression. Chem. Res. Toxicol. 2006, 19, 1499–1505. [Google Scholar] [CrossRef]
  96. Zeng, N.; Zhang, G.; Hu, X.; Pan, J.; Gong, D. Mechanism of fisetin suppressing superoxide anion and xanthine oxidase activity. J. Funct. Foods 2019, 58, 1–10. [Google Scholar] [CrossRef]
  97. Md Idris, M.H.; Mohd Amin, S.N.; Mohd Amin, S.N.; Nyokat, N.; Khong, H.Y.; Selvaraj, M.; Zakaria, Z.A.; Shaameri, Z.; Hamzah, A.S.; Teh, L.K.; et al. Flavonoids as dual inhibitors of cyclooxygenase-2 (COX-2) and 5-lipoxygenase (5-LOX): Molecular docking and in vitro studies. Beni Suef Univ. J. Basic. Appl. Sci. 2022, 11, 117. [Google Scholar] [CrossRef]
  98. Samsonowicz, M.; Regulska, E.; Kalinowska, M. Hydroxyflavone metal complexes-molecular structure, antioxidant activity and biological effects. Chem. Biol. Interact. 2017, 273, 245–256. [Google Scholar] [CrossRef] [PubMed]
  99. Kejík, Z.; Kaplánek, R.; Masařík, M.; Babula, P.; Matkowski, A.; Filipenský, P.; Veselá, K.; Gburek, J.; Sýkora, D.; Martásek, P.; et al. Iron Complexes of Flavonoids-Antioxidant Capacity and Beyond. Int. J. Mol. Sci. 2021, 22, 646. [Google Scholar] [CrossRef] [PubMed]
  100. Lesjak, M.; Beara, I.; Simin, N.; Pintać, D.; Majkić, T.; Bekvalac, K.; Orčić, D.; Mimica-Dukić, N. Antioxidant and anti-inflammatory activities of quercetin and its derivatives. J. Funct. Foods 2018, 40, 68–75. [Google Scholar] [CrossRef]
  101. Simunkova, M.; Barbierikova, Z.; Jomova, K.; Hudecova, L.; Valko, M. Antioxidant vs. Prooxidant Properties of the Flavonoid, Kaempferol, in the Presence of Cu(II) Ions: A ROS-Scavenging Activity, Fenton Reaction and DNA Damage Study. Int. J. Mol. Sci. 2021, 22, 1619. [Google Scholar] [CrossRef]
  102. Zhang, K.; Dai, Z.; Zhang, W.; Gao, Q.; Dai, Y.; Xia, F.; Zhang, X. EDTA-based adsorbents for the removal of metal ions in wastewater. Coord. Chem. Rev. 2021, 434, 213809. [Google Scholar] [CrossRef]
  103. Grin, P.M.; Dwivedi, D.J.; Chathely, K.M.; Trigatti, B.L.; Prat, A.; Seidah, N.G.; Liaw, P.C.; Fox-Robichaud, A.E. Low-density lipoprotein (LDL)-dependent uptake of Gram-positive lipoteichoic acid and Gram-negative lipopolysaccharide occurs through LDL receptor. Sci. Rep. 2018, 8, 10496. [Google Scholar] [CrossRef] [Green Version]
  104. Torello, C.O.; Alvarez, M.C.; Olalla Saad, S.T. Polyphenolic Flavonoid Compound Quercetin Effects in the Treatment of Acute Myeloid Leukemia and Myelodysplastic Syndromes. Molecules 2021, 26, 5781. [Google Scholar] [CrossRef]
  105. Papi, S.; Ahmadizar, F.; Hasanvand, A. The role of nitric oxide in inflammation and oxidative stress. Immunopathol. Persa 2019, 5, e08. [Google Scholar] [CrossRef]
  106. Gutiérrez-Venegas, G.; Ventura-Arroyo, J.A.; Arreguín-Cano, J.A.; Ostoa-Pérez, M.F. Flavonoids inhibit iNOS production via mitogen activated proteins in lipoteichoic acid stimulated cardiomyoblasts. Int. Immunopharmacol. 2014, 21, 320–327. [Google Scholar] [CrossRef]
  107. Terao, J. Factors modulating bioavailability of quercetin-related flavonoids and the consequences of their vascular function. Biochem. Pharmacol. 2017, 139, 15–23. [Google Scholar] [CrossRef] [PubMed]
  108. Jiang, L.; Li, Z.; Xie, Y.; Liu, L.; Cao, Y. Cyanidin chloride modestly protects Caco-2 cells from ZnO nanoparticle exposure probably through the induction of autophagy. Food Chem. Toxicol. 2019, 127, 251–259. [Google Scholar] [CrossRef] [PubMed]
  109. Mirza, S.; Sharma, G.; Parshad, R.; Gupta, S.D.; Pandya, P.; Ralhan, R. Expression of DNA methyltransferases in breast cancer patients and to analyze the effect of natural compounds on DNA methyltransferases and associated proteins. J. Breast Cancer 2013, 16, 23–31. [Google Scholar] [CrossRef] [Green Version]
  110. Salehi, B.; Quispe, C.; Chamkhi, I.; Omari, N.E.; Balahbib, A.; Sharifi-Rad, J.; Bouyahya, A.; Akram, M.; Iqbal, M.; Docea, A.O.; et al. Pharmacological Properties of Chalcones: A Review of Preclinical Including Molecular Mechanisms and Clinical Evidence. Front. Pharm. 2021, 11, 2068. [Google Scholar] [CrossRef]
  111. Sun, T.L.; Li, W.Q.; Tong, X.L.; Liu, X.Y.; Zhou, W.H. Xanthohumol attenuates isoprenaline-induced cardiac hypertrophy and fibrosis through regulating PTEN/AKT/mTOR pathway. Eur. J. Pharmacol. 2021, 891, 173690. [Google Scholar] [CrossRef] [PubMed]
  112. Wu, S.C.; Yang, Z.Q.; Liu, F.; Peng, W.J.; Qu, S.Q.; Song, X.B.; Zhu, K.; Shen, J.Z. Antibacterial effect and mode of action of flavonoids from licorice against methicil-lin-resistant Staphylococcus aureus. Front. Microbiol. 2019, 10, 2489. [Google Scholar] [CrossRef] [Green Version]
  113. Yang, L.; Xie, X.; Tu, Z.; Fu, J.; Xu, D.; Zhou, Y. The signal pathways and treatment of cytokine storm in COVID-19. Sig. Transduct. Target. Ther. 2021, 6, 255. [Google Scholar] [CrossRef]
  114. Efenberger-Szmechtyk, M.; Nowak, A.; Czyzowska, A. Plant extracts rich in polyphenols: Antibacterial agents and natural preservatives for meat and meat products. Crit. Rev. Food Sci. Nutr. 2021, 61, 149–178. [Google Scholar] [CrossRef]
  115. Tang, S.M.; Deng, X.T.; Zhou, J.; Li, Q.P.; Ge, X.X.; Miao, L. Pharmacological basis and new insights of quercetin action in respect to its anti-cancer effects. Biomed. Pharma. 2020, 121, 109604. [Google Scholar] [CrossRef]
  116. Barreca, D.; Gattuso, G.; Bellocco, E.; Calderaro, A.; Trombetta, D.; Smeriglio, A.; Lagana, G.; Daglia, M.; Menehini, S.; Nabavi, S.M. Flavanones: Citrus phytochemical with health-promoting properties. Biofactors 2017, 43, 495–506. [Google Scholar] [CrossRef]
  117. Chang, X.; Zhang, T.; Wang, J.; Liu, Y.; Yan, P.; Meng, Q.; Yin, Y.; Wang, S. SIRT5-related desuccinylation modification contributes to quercetin-induced protection against heart failure and high-glucose-prompted cardiomyocytes injured through regulation of mitochondrial quality surveillance. Oxid. Med. Cell. Longev. 2021, 2021, 5876841. [Google Scholar] [CrossRef] [PubMed]
  118. Testai, L.; Piragine, E.; Piano, I.; Flori, L.; Da Pozzo, E.; Miragliotta, V.; Pirone, A.; Citi, V.; Mannelli, L.D.C.; Brogi, S.; et al. The citrus flavonoid naringenin protects the myocardium from ageing-dependent dysfunction: Potential role of SIRT1. Oxid. Med. Cell. Longev. 2020, 2020, 4650207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Zhang, L.; Xu, X.; Jiang, T.; Wu, K.; Ding, C.; Liu, Z.; Zhang, X.; Yu, T.; Song, C. Citrus aurantium Naringenin Prevents Osteosarcoma Progression and Recurrence in the Patients Who Underwent Osteosarcoma Surgery by Improving Antioxidant Capability. Oxid. Med. Cell. Longev. 2018, 7, 8713263. [Google Scholar] [CrossRef] [Green Version]
  120. Guardia, T.; Rotelli, A.E.; Juarez, A.O.; Pelzer, L.E. Anti-inflammatory properties of plant flavonoids. Effects of rutin, quercetin and hesperidin on adjuvant arthritis in rat. Il Farm. 2001, 56, 683–687. [Google Scholar] [CrossRef]
  121. Pandey, P.; Khan, F. A mechanistic review of the anticancer potential of hesperidin, a natural flavonoid from citrus fruits. Nutr. Res. 2021, 92, 21–31. [Google Scholar] [CrossRef]
  122. Yang, J.; Liu, L.; Li, M.; Huang, X.; Yang, H.; Li, K. Naringenin inhibits pro-inflammatory cytokine production in macrophages through inducing MT1G to suppress the activation of NF-κB. Mol. Immunol. 2021, 137, 155–162. [Google Scholar] [CrossRef]
  123. Zhong, L.; Lin, Y.; Wang, C.; Niu, B.; Xu, Y.; Zhao, G.; Zhao, J. Chemical Profile, Antimicrobial and Antioxidant Activity Assessment of the Crude Extract and Its Main Flavonoids from Tartary Buckwheat Sprouts. Molecules 2022, 27, 374. [Google Scholar] [CrossRef]
  124. Wu, C.; Chen, R.L.; Wang, Y.; Wu, W.Y.; Li, G. Acacetin alleviates myocardial ischemia/reperfusion injury by inhibiting oxidative stress and apoptosis via the Nrf-2/HO-1 pathway. Pharm. Biol. 2022, 60, 553–561. [Google Scholar] [CrossRef]
  125. Wu, Q.; Kroon, P.A.; Shao, H.; Needs, P. Differential effects of quercetin and two of its derivatives, isorhamnetin and isorhamnetin-3-glucuronide, in inhibiting the proliferation of human breast-cancer MCF-7 cells. J. Agr. Food Chem. 2018, 66, 7181–7189. [Google Scholar] [CrossRef]
  126. Lang, S.J.; Schmiech, M.; Hafner, S.; Paetz, C.; Simmet, T. Chrysosplenol d, a flavonol from Artemisia annua, induces ERK1/2-mediated apoptosis in triple negative human breast cancer cells. Int. J. Mol. Sci. 2020, 21, 4090. [Google Scholar] [CrossRef]
  127. Chen, X.; Wu, Y.; Gu, J.; Liang, P.; Qin, J. Anti-invasive effect and pharmacological mechanism of genistein against colorectal cancer. Biofactors 2020, 46, 620–628. [Google Scholar] [CrossRef]
  128. Hirchaud, F.; Hermetet, F.; Ablise, M.; Fauconnet, S.; Vuitton, D.A.; Prétet, J.L.; Mougin, C. Isoliquiritigenin induces caspase-dependent apoptosis via downregulation of HPV16 E6 expression in cervical cancer Ca Ski cells. Planta Med. 2013, 79, 1628–1635. [Google Scholar] [CrossRef] [PubMed]
  129. Yao, Z.; Xu, X.; Huang, Y. Daidzin inhibits growth and induces apoptosis through the JAK2/STAT3 in human cervical cancer HeLa cells. Saudi J. Biol. Sci. 2021, 28, 7077–7081. [Google Scholar] [CrossRef] [PubMed]
  130. Michaelis, M.; Sithisarn, P.; Cinatl, J. Effects of flavonoid-induced oxidative stress on anti-H5N1 influenza a virus activity exerted by baicalein and biochanin A. BMC Res. Notes. 2014, 7, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Guntaka, R.V. New Insights to Prevent Liver Fibrosis by Targeting YB-1 and Collagen Genes. OBM Hepatol. Gastroenterol. 2019, 3, 1. [Google Scholar] [CrossRef]
  132. Wu, J.; Zhou, T.; Wang, Y.; Jiang, Y.; Wang, Y. Mechanisms and Advances in Anti-Ovarian Cancer with Natural Plants Component. Molecules 2021, 26, 5949. [Google Scholar] [CrossRef] [PubMed]
  133. Figueiredo-Rinhel, A.S.; Santos, E.O.; Kabeya, L.M.; Azzolini, A.E.; Simões-Ambrosio, L.M.; Lucisano-Valim, Y.M. The flavonols quercetin, myricetin, kaempferol, and galangin inhibit the net oxygen consumption by immune complex-stimulated human and rabbit neutrophils. Z. Nat. C J. Biosci. 2014, 69, 346–356. [Google Scholar] [CrossRef]
  134. Farkhondeh, T.S.; Samarghandian Bafandeh, F. The Cardiovascular Protective Effects of Chrysin: A Narrative Review on Experimental Researches. Cardiovasc. Hematol. Agents Med. Chem. 2019, 17, 17–27. [Google Scholar] [CrossRef]
  135. Khazeei Tabari, M.A.; Iranpanah, A.; Bahramsoltani, R.; Rahimi, R. Flavonoids as promising antiviral agents against SARS-CoV-2 infection: A mechanistic review. Molecules 2021, 26, 3900. [Google Scholar] [CrossRef]
  136. Sugawara, T.; Sakamoto, K. Quercetin enhances motility in aged and heat-stressed Caenorhabditis elegans nematodes by modulating both HSF-1 activity, and insulin-like and p38-MAPK signalling. PLoS ONE 2020, 15, e0238528. [Google Scholar] [CrossRef]
  137. Sanches-Silva, A.; Testai, L.; Nabavi, S.F.; Battino, M.; Devi, K.P.; Tejada, S.; Sureda, A.; Xu, S.W.; Yousefi, B.; Majidinia, M.; et al. Therapeutic potential of polyphenols in cardiovascular diseases: Regulation of mTOR signaling pathway. Pharma. Res. 2020, 152, 104626. [Google Scholar] [CrossRef] [PubMed]
  138. Kiani, K.; Dhuli, K.; Anpilogov, K.; Bressan, S.; Dautaj, A.; Dundar, M.; Beccari, T.; Ergoren, M.; Bertelli, M. Natural compounds as inhibitors of SARS-CoV-2 endocytosis: A promising approach against COVID-19. Acta Med. 2020, 91, 13. [Google Scholar] [CrossRef]
  139. Kandhari, K.; Mishra, J.P.N.; Singh, R.P. Acacetin inhibits cell proliferation, survival, and migration in human breast cancer cells. Int. J. Pharma. Biol. Sci. 2019, 9, 443–452. [Google Scholar] [CrossRef]
  140. Zhou, B.; Fang, L.; Dong, Y.; Yang, J.; Chen, X.; Zhang, N.; Zhu, Y.; Huang, T. Mitochondrial quality control protects photoreceptors against oxidative stress in the H2O2-induced models of retinal degeneration diseases. Cell Death Dis. 2021, 12, 413. [Google Scholar] [CrossRef] [PubMed]
  141. Wang, M.; Liu, Y.; Pan, R.L.; Wang, R.Y.; Ding, S.L.; Dong, W.R.; Sun, X.B. Protective effects of Myrica rubra flavonoids against hypoxia/reoxygenation-induced cardiomyocyte injury via the regulation of the PI3K/Akt/GSK3β pathway. Int. J. Mol. Med. 2019, 43, 2133–2143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Wang, X.; Yang, Y.; An, Y.; Fang, G. The mechanism of anticancer action and potential clinical use of kaempferol in the treatment of breast cancer. Biomed. Pharm. 2019, 117, 109086. [Google Scholar] [CrossRef]
  143. Chan, K.K.L.; Siu, M.K.Y.; Jiang, Y.X.; Wang, J.J.; Leung, T.H.Y.; Ngan, H.Y.S. Estrogen receptor modulators genistein, daidzein and ERB-041 inhibit cell migration, invasion, proliferation and sphere formation via modulation of FAK and PI3K/AKT signaling in ovarian cancer. Cancer Cell Int. 2018, 18, 65. [Google Scholar] [CrossRef] [Green Version]
  144. Zhang, H.J.; Deng, A.J.; Zhang, Z.H.; Yu, Z.H.; Liu, Y.; Peng, S.Y.; Wu, L.Q.; Qin, H.L.; Wang, W.J. The protective effect of epicatechin on experimental ulcerative colitis in mice is mediated by increasing antioxidation and by the inhibition of NF-κB pathway. Pharm. Rep. 2016, 68, 514–520. [Google Scholar] [CrossRef]
  145. Prince, P.S.M. A biochemical, electrocardiographic, electrophoretic, histopathological and in vitro study on the protective effects of (−) epicatechin in isoproterenol-induced myocardial infarcted rats. Eur. J. Pharm. 2011, 671, 95–101. [Google Scholar] [CrossRef]
  146. Wang, B.; Wu, L.; Chen, J.; Dong, L.; Chen, C.; Wen, Z.; Hu, J.; Fleming, I.; Wang, D.W. Metabolism pathways of arachidonic acids: Mechanisms and potential therapeutic targets. Sig. Transduct. Target. Ther. 2021, 6, 94. [Google Scholar] [CrossRef]
  147. Ribeiro, D.; Freitas, M.; Tomé, S.M.; Silva, A.M.S.; Laufer, S.; Lima, J.L.F.C.; Fernandes, E. Flavonoids Inhibit COX-1 and COX-2 Enzymes and Cytokine/Chemokine Production in Human Whole Blood. Inflammation 2014, 38, 858–870. [Google Scholar] [CrossRef] [PubMed]
  148. Al-Khayri, J.M.; Sahana, G.R.; Nagella, P.; Joseph, V.; Alessa, F.M.; Al-Mssallem, M.Q. Flavonoids as Potential Anti-Inflammatory Molecules: A Review. Molecules 2022, 27, 2901. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, Q.; Yue, S.J. Editorial: Flavonoids and Cardiovascular Metabolism. Front. Nutr. 2022, 9, 939798. [Google Scholar] [CrossRef] [PubMed]
  150. Liu, X.M.; Liu, Y.J.; Huang, Y.; Yu, H.J.; Yuan, S.; Tang, B.W.; Wang, P.G.; He, Q.Q. Dietary total flavonoids intake and risk of mortality from all causes and cardiovascular disease in the general population: A systematic review and meta-analysis of cohort studies. Mol. Nutr. Food Res. 2017, 61, 1601003. [Google Scholar] [CrossRef]
  151. Sanchez, M.; Romero, M.; Gomez-Guzman, M.; Tamargo, J.; Perez-Vizcaino, F.; Duarte, J. Cardiovascular effects of flavonoids. Curr. Med. Chem. 2019, 26, 6991–7034. [Google Scholar] [CrossRef]
  152. Mulvihill, E.E.; Burke, A.C.; Huff, M.W. Citrus Flavonoids as Regulators of Lipoprotein Metabolism and Atherosclerosis. Annu. Rev. Nutr. 2016, 36, 275–299. [Google Scholar] [CrossRef]
  153. Chacko, B.K.; Chandler, R.T.; D’Alessandro, T.L.; Mundhekar, A.; Khoo, N.K.H.; Botting, N.; Barnes, S.; Patel, R.P. Anti-inflammatory effects of isoflavones are dependent on flow and human endothelial cell PPARgamma. J. Nutr. 2007, 137, 351–356. [Google Scholar] [CrossRef] [Green Version]
  154. Wallace, T.C.; Slavin, M.; Frankenfeld, C.L. Systematic Review of Anthocyanins and Markers of Cardiovascular Disease. Nutrients 2016, 8, 32. [Google Scholar] [CrossRef] [Green Version]
  155. Chu, Q.; Zhang, S.; Chen, M.; Ha, W.; Ji, R.; Chen, W.; Zheng, X. Cherry anthocyanins regulate NAFLD by promoting autophagy pathway. Oxid. Med. Cell. Longev. 2019, 2019, 4825949. [Google Scholar] [CrossRef] [Green Version]
  156. Guo, Y.; Zhang, B.Y.; Peng, Y.F.; Chang, L.C.; Li, Z.Q.; Zhang, X.X.; Zhang, D.J. Mechanism of action of flavonoids of oxytropis falcata on the alleviation of myocardial ischemia–reperfusion injury. Molecules 2022, 27, 1706. [Google Scholar] [CrossRef]
  157. Sebghatollahi, Z.; Ghanadian, M.; Agarwal, P.; Ghaheh, H.S.; Mahato, N.; Yogesh, R. Citrus flavonoids: Biological activities, implementation in skin health, and topical applications: A Review. ACS Food Sci. Technol. 2022, 2, 1417–1432. [Google Scholar] [CrossRef]
  158. Chambers, C.S.; Viktorová, J.; Rěhorŏvá, K.; Biedermann, D.; Turková, L.; Macek, T.; Křen, V.; Valentová, K. Defying multidrug resistance! Modulation of related transporters by flavonoids and flavonolignans. J. Agric. Food Chem. 2019, 68, 1763–1779. [Google Scholar] [CrossRef] [PubMed]
  159. Zhang, H.; Wang, X.; He, D.; Zou, D.; Zhao, R.; Wang, H.; Li, S.; Xu, Y.; Abudureheman, B. Optimization of flavonoid extraction from Xanthoceras sorbifolia Bunge flowers, and the antioxidant and antibacterial capacity of the extract. Molecules 2021, 27, 113. [Google Scholar] [CrossRef]
  160. Yuan, G.; Guan, Y.; Yi, H.; Lai, S.; Sun, Y.; Cao, S. Antibacterial activity and mechanism of plant flavonoids to gram-positive bacteria predicted from their lipophilicities. Sci. Rep. 2021, 11, 10471. [Google Scholar] [CrossRef] [PubMed]
  161. Al Aboody, M.S.; Mickymaray, S. Anti-fungal efficacy and mechanisms of flavonoids. Antibiotics 2020, 9, 45. [Google Scholar] [CrossRef] [Green Version]
  162. Cushnie, T.P.T.; Lamb, A.J. Antimicrobial activity of flavonoids. Int. J. Antimicrob. Ag. 2005, 26, 343–356. [Google Scholar] [CrossRef] [PubMed]
  163. Pasupuleti, V.R.; Arigela, C.S.; Gan, S.H.; Salam, S.K.N.; Krishnan, K.T.; Rahman, N.A.; Jeffree, M.S. A review on oxidative stress, diabetic complications, and the roles of honey polyphenols. Oxid. Med. Cell. Longev. 2020, 2020, 8878172. [Google Scholar] [CrossRef]
  164. Morkunas, I.; Ratajczak, L. The role of sugar signaling in plant defense responses against fungal pathogens. Acta Physiol. Plant. 2014, 36, 1607–1619. [Google Scholar] [CrossRef] [Green Version]
  165. Traganos, F.; Kimmel, M.; Bueti, C.; Darzynkiewicz, Z. Effects of inhibition of RNA or protein synthesis on CHO cell cycle progression. J. Cell. Physiol. 1987, 133, 277–287. [Google Scholar] [CrossRef]
  166. Rajendran, P.; Rengarajan, T.; Nandakumar, N.; Palaniswami, R.; Nishigaki, Y.; Nishigaki, I. Kaempferol, a potential cytostatic and cure for inflammatory disorders. Eur. J. Med. Chem. 2014, 86, 103–112. [Google Scholar] [CrossRef]
  167. Li, A.; Zhao, Z.; Zhang, S.; Zhang, Z.; Shi, Y. Fungicidal activity and mechanism of action of glabridin from Glycyrrhiza glabra L. Int. J. Mol. Sci. 2021, 22, 10966. [Google Scholar] [CrossRef]
  168. Lalani, S.; Poh, C.L. Flavonoids as antiviral agents for Enterovirus A71 (EV-A71). Viruses 2020, 12, 184. [Google Scholar] [CrossRef] [Green Version]
  169. Roschek, B.; Fink, R.C.; McMichael, M.D.; Li, D.; Alberte, R.S.J.P. Elderberry flavonoids bind to and prevent H1N1 infection in vitro. Phytochemistry 2009, 70, 1255–1261. [Google Scholar] [CrossRef]
  170. Sithisarn, P.; Michaelis, M.; Schubert-Zsilavecz, M.; Cinatl, J., Jr. Differential antiviral and anti-inflammatory mechanisms of the flavonoids biochanin A and baicalein in H5N1 influenza A virus-infected cells. Antivir. Res. 2013, 97, 41–48. [Google Scholar] [CrossRef]
  171. Badshah, S.L.; Faisal, S.; Muhammad, A.; Poulson, B.G.; Emwas, A.H.; Jaremko, M. Antiviral activities of flavonoids. Biomed. Pharmacother. 2021, 140, 111596. [Google Scholar] [CrossRef] [PubMed]
  172. Ezzati, M.; Yousefi, B.; Velaei, K.; Safa, A. A review on anti-cancer properties of Quercetin in breast cancer. Life Sci. 2020, 248, 117463. [Google Scholar] [CrossRef]
  173. Veeramuthu, D.; Raja, W.R.T.; Al-Dhabi, N.A.; Savarimuthu, I. Flavonoids: Anticancer properties. In Flavonoids: From Biosynthesis to Human Health; IntechOpen: London, UK, 2017; p. 287. [Google Scholar]
  174. Coelho, P.L.C.; Oliveira, M.N.; da Silva, A.B.; Pitanga, B.P.; Silva, V.D.; Faria, G.P.; Sampaio, G.P.; Costa, M.D.D.; Braga-de-Souza, S.; Costa, S.L. The flavonoid apigenin from Croton betulaster Mull inhibits proliferation, induces differentiation and regulates the inflammatory profile of glioma cells. Anticancer Drug. 2016, 27, 960–969. [Google Scholar] [CrossRef] [PubMed]
  175. Lewis, K.A.; Jordan, H.R.; Tollefsbol, T.O. Effffects of SAHA and EGCG on Growth Potentiation of Triple-Negative Breast Cancer Cells. Cancers 2018, 11, 23. [Google Scholar] [CrossRef] [Green Version]
  176. Adinew, G.M.; Taka, E.; Mendonca, P.; Messeha, S.S.; Soliman, K.F. The anticancer effects of flavonoids through miRNAs modulations in triple-negative breast cancer. Nutrients 2021, 13, 1212. [Google Scholar] [CrossRef] [PubMed]
  177. Lecumberri, E.; Dupertuis, Y.M.; Miralbell, R.; Pichard, C. Green tea polyphenol epigallocatechin-3-gallate (EGCG) as adjuvant in cancer therapy. Clin. Nutr. 2013, 32, 894–903. [Google Scholar] [CrossRef] [Green Version]
  178. Gao, Y.; Tollefsbol, T.O. Combinational Proanthocyanidins and Resveratrol Synergistically Inhibit Human Breast Cancer Cells and Impact Epigenetic-Mediating Machinery. Int. J. Mol. Sci. 2018, 19, 2204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Selvakumar, P.; Badgeley, A.; Murphy, P.; Anwar, H.; Sharma, U.; Lawrence, K.; Lakshmikuttyamma, A. Flavonoids and Other Polyphenols Act as Epigenetic Modifiers in Breast Cancer. Nutrients 2020, 12, 761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Chen, J.; Liu, Y.; Zhao, Z.; Qiu, J. Oxidative stress in the skin: Impact and related protection. Int. J. Cosmet. Sci. 2021, 43, 495–509. [Google Scholar] [CrossRef] [PubMed]
  181. Sharifi-Rad, M.; Anil Kumar, N.V.; Zucca, P.; Varoni, E.M.; Dini, L.; Panzarini, E.; Rajkovic, J.; Fokou, P.V.T.; Azzini, E.; Peluso, I.; et al. Lifestyle, oxidative stress, and antioxidants: Back and forth in the pathophysiology of chronic diseases. Front. Physiol. 2020, 11, 694. [Google Scholar] [CrossRef] [PubMed]
  182. Hoang, H.T.; Moon, J.Y.; Lee, Y.C. Natural Antioxidants from Plant Extracts in Skincare Cosmetics: Recent Applications, Challenges and Perspectives. Cosmetics 2021, 8, 106. [Google Scholar] [CrossRef]
  183. Saewan, N.; Jimtaisong, A. Natural products as photoprotection. J. Cosmet. Dermatol. 2015, 14, 47–63. [Google Scholar] [CrossRef]
  184. Saewan, N.; Jimtaisong, A. Photoprotection of natural flavonoids. J. Appl. Pharma Sci. 2013, 3, 129–141. [Google Scholar] [CrossRef] [Green Version]
  185. Mottaghipisheh, J.; Taghrir, H.; Boveiri Dehsheikh, A.; Zomorodian, K.; Irajie, C.; Mahmoodi Sourestani, M.; Iraji, A. Linarin, a Glycosylated Flavonoid, with Potential Therapeutic Attributes: A Comprehensive Review. Pharmaceuticals 2021, 14, 1104. [Google Scholar] [CrossRef]
  186. Atif, A.; Naveed, A.; Muhammad, S.K.; Fatima, R.; Furqan, M.I.; Muhammad, T.; Minhaj, U.; Ehsan, E. Moisturizing effect of cream containing Moringa oleifera (Sohajana) leaf extract by biophysical techniques: In vivo evaluation. J. Med. Plants Res. 2013, 7, 386–391. [Google Scholar] [CrossRef]
  187. Cordeiro, M.L.S.; Martins, V.G.Q.A.; Silva, A.P.; Rocha, H.A.O.; Rachetti, V.P.S.; Scortecci, K.C. Phenolic Acids as Antidepressant Agents. Nutrients 2022, 14, 4309. [Google Scholar] [CrossRef]
  188. Aboonabi, A.; Singh, I. Chemopreventive role of anthocyanins in atherosclerosis via activation of Nrf2–ARE as an indicator and modulator of redox. Biomed. Pharmacother. 2015, 72, 30–36. [Google Scholar] [CrossRef]
  189. Tran, P.L.P.; Le, H.P. Plant Flavonoids as Potential Natural Antioxidants in Phytocosmetics. J. Tech. Educ. Sci. 2022, 70B, 86–93. [Google Scholar] [CrossRef]
  190. Bjorklund, G.; Shanaida, M.; Lysiuk, R.; Butnariu, M.; Peana, M.; Sarac, I.; Strus, O.; Smetanina, K.; Chirumbolo, S. Natural Compounds and Products from an Anti-Aging Perspective. Molecules 2022, 27, 7084. [Google Scholar] [CrossRef]
  191. Micek, I.; Nawrot, J.; Seraszek-Jaros, A.; Jenerowicz, D.; Schroeder, G.; Spiżewski, T.; Suchan, A.; Pawlaczyk, M.; Gornowicz-Porowska, J. Taxifolin as a Promising Ingredient of Cosmetics for Adult Skin. Antioxidants 2021, 10, 1625. [Google Scholar] [CrossRef] [PubMed]
  192. Hwang, Y.S.; Chang, B.Y.; Kim, D.S.; Cho, H.K.; Kiml, S.Y. Effects of the Syzygium aromaticum L. extract on antioxidation and inhibition of matrix metalloproteinase in human dermal fibroblast. Asian Pac. J. Trop. Bio. 2019, 9, 53. [Google Scholar] [CrossRef]
  193. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The immunomodulatory and anti-inflammatory role of polyphenols. Nutrients 2018, 10, 1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Phung, H.M.; Lee, S.; Hong, S.; Lee, S.; Jung, K.; Kang, K.S. Protective Effect of Polymethoxyflavones Isolated from Kaempferia parviflora against TNF-α-Induced Human Dermal Fibroblast Damage. Antioxidants 2021, 10, 1609. [Google Scholar] [CrossRef]
  195. Lee, S.; Jang, T.; Kim, K.H.; Kang, K.S. Improvement of Damage in Human Dermal Fibroblasts by 3, 5, 7-Trimethoxyflavone from Black Ginger (Kaempferia parviflora). Antioxidants 2022, 11, 425. [Google Scholar] [CrossRef]
  196. Chaiprasongsuk, A.; Panich, U. Role of phytochemicals in skin photoprotection via regulation of Nrf2. Front. Pharma. 2022, 13, 823881. [Google Scholar] [CrossRef]
  197. Awuchi, C.G.; Twinomuhwezi, H.; Victory, I.S.; Ikechukwu, A.O. Food Additives and Food Preservatives for Domestic and Industrial Food Applications. J. Anim. Health 2020, 2, 1–16. [Google Scholar]
  198. Ruiz-Cruz, S.; Chaparro-Hernández, S.; Hernández-Ruiz, K.L.; Cira-Chávez, L.A.; Estrada-Alvarado, M.I.; Gassos-Ortega, L.E.; Ornelas-Paz, J.; Lopez Mata, M.A. Flavonoids: Important biocompounds in food. In Flavonoids: From Biosynthesis to Human. Health; IntechOpen: London, UK, 2017; pp. 353–369. [Google Scholar] [CrossRef] [Green Version]
  199. Prakash, D.; Kumar, N. Cost Effective Natural Antioxidants. Nutrients, Dietary Supplements, and Nutriceuticals; Humana Press: Totowa, NJ, USA, 2011; pp. 163–187. [Google Scholar] [CrossRef]
  200. Babuskin, S.; Babu, P.A.S.; Sasikala, M.; Sabina, K.; Archana, G.; Sivarajan, M.; Sukumar, M. Antimicrobial and antioxidant effects of spice extracts on the shelf life extension of raw chicken meat. Int. J. Food Microbiol. 2014, 171, 32–40. [Google Scholar] [CrossRef]
  201. Ji, H.F.; Li, X.J.; Zhang, H.Y. Natural products and drug discovery: Can thousands of years of ancient medical knowledge lead us to new and powerful drug combinations in the fight against cancer and dementia? EMBO Rep. 2009, 10, 194–200. [Google Scholar] [CrossRef] [Green Version]
  202. Gutiérrez-del-Río, I.; López-Ibáñez, S.; Magadán-Corpas, P.; Fernández-Calleja, L.; Pérez-Valero, A.; Tuñón-Granda, M.; Miguelez, E.; Villar, C.; Lombo, F. Terpenoids and polyphenols as natural antioxidant agents in food preservation. Antioxidants 2021, 10, 1264. [Google Scholar] [CrossRef]
  203. Rathod, N.B.; Ranveer, R.C.; Benjakul, S.; Kim, S.K.; Pagarkar, A.U.; Patange, S.; Ozogul, F. Recent developments of natural antimicrobials and antioxidants on fish and fishery food products. Compre. Rev. Food Sci. Food Saf. 2021, 20, 4182–4210. [Google Scholar] [CrossRef] [PubMed]
  204. Manzoni, C.; Kia, D.A.; Vandrovcova, J.; Hardy, J.; Wood, N.W.; Lewis, P.A.; Ferrari, R. Genome, transcriptome and proteome: The rise of omics data and their integration in biomedical sciences. Brief. Bioinform. 2018, 19, 286–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Li, P.; Ruan, Z.; Fei, Z.; Yan, J.; Tang, G. Integrated transcriptome and metabolome analysis revealed that flavonoid biosynthesis may dominate the resistance of Zanthoxylum bungeanum against stem canker. J. Agric. Food Chem. 2021, 69, 6360–6378. [Google Scholar] [CrossRef]
  206. Yuan, Y.; Zuo, J.; Zhang, H.; Zu, M.; Yu, M.; Liu, S. Transcriptome and metabolome profiling unveil the accumulation of flavonoids in Dendrobium officinale. Genomics 2022, 114, 110324. [Google Scholar] [CrossRef]
  207. Zhang, Z.; Tian, C.; Zhang, Y.; Li, C.; Li, X.; Yu, Q.; Wang, S.; Wang, X.; Chen, X.; Feng, S. Transcriptomic and metabolomic analysis provides insights into anthocyanin and procyanidin accumulation in pear. BMC Plant Biol. 2020, 20, 1–14. [Google Scholar] [CrossRef] [Green Version]
  208. Zhang, Q.; Wang, L.; Liu, Z.; Zhao, Z.; Zhao, J.; Wang, Z.; Zhou, G.; Liu, P.; Liu, M. Transcriptome and metabolome profiling unveil the mechanisms of Ziziphus jujuba Mill. peel coloration. Food Chem. 2020, 312, 125903. [Google Scholar] [CrossRef]
  209. Peng, X.Q.; Ai, Y.J.; Pu, Y.T.; Wang, X.J.; Li, Y.H.; Wang, Z.; Zhuang, W.B.; Yu, B.J.; Zhu, Z.Q. Transcriptome and metabolome analyses reveal molecular mechanisms of anthocyanin-related leaf color variation in poplar (Populus deltoides) cultivars. Front. Plant Sci. 2023, 14, 554. [Google Scholar] [CrossRef]
  210. Teng, H.; Zheng, Y.; Cao, H.; Huang, Q.; Xiao, J.; Chen, L. Enhancement of bioavailability and bioactivity of diet-derived flavonoids by application of nanotechnology: A review. Crit. Rev. Food Sci. Nutr. 2023, 63, 378–393. [Google Scholar] [CrossRef]
  211. Jannat, K.; Paul, A.K.; Bondhon, T.A.; Hasan, A.; Nawaz, M.; Jahan, R.; Mahboob, T.; Nissapatorn, V.; Wilairatana, P.; Pereira, M.d.L.; et al. Nanotechnology Applications of Flavonoids for Viral Diseases. Pharmaceutics 2021, 13, 1895. [Google Scholar] [CrossRef] [PubMed]
  212. Dobrzynska, M.; Napierala, M.; Florek, E. Flavonoid nanoparticles: A promising approach for cancer therapy. Biomolecules 2020, 10, 1268. [Google Scholar] [CrossRef]
  213. Siddiqui, I.A.; Adhami, V.M.; Bharali, D.J.; Hafeez, B.B.; Asim, M.; Khwaja, S.I.; Ahmad, N.; Cui, H.D.; Mousa, S.A.; Mukhtar, H. Introducing nanochemoprevention as a novel approach for cancer control: Proof of principle with green tea polyphenol epigallocatechin-3-gallate. Cancer Res. 2009, 69, 1712–1716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Tan, B.J.; Liu, Y.; Chang, K.L.; Lim, B.K.; Chiu, G.N. Perorally active nanomicellar formulation of quercetin in the treatment of lung cancer. Int. J. Nanomed. 2012, 7, 651–661. [Google Scholar] [CrossRef] [Green Version]
  215. Sindhu, R.K.; Verma, R.; Salgotra, T.; Rahman, M.H.; Shah, M.; Akter, R.; Murad, W.; Mubin, S.; Bibi, P.; Qusti, S.; et al. Impacting the remedial potential of nano delivery-based flavonoids for breast cancer treatment. Molecules 2021, 26, 5163. [Google Scholar] [CrossRef]
Figure 1. The basic molecular structure of flavonoids, classification, and distribution in various plants.
Figure 1. The basic molecular structure of flavonoids, classification, and distribution in various plants.
Molecules 28 04982 g001
Figure 2. Flavonoid synthesis pathway. CHS (chalcone synthase) can catalyze three molecules of malonyl-CoA and one molecule of p-coumaroyl-CoA to form naringeninchalcone [56]. Malonyl-CoA is an important precursor for the synthesis of natural products, including flavonoids and polyketides [57]. CHI (chalcone isomerase) converted naringenin-chalcone into flavanones [58]. Naringenin, as an important flavonoid skeleton, is catalyzed by FNSI and FNS II (flavone synthase I and flavone synthase II) and IFS (isoflavone synthase) to form flavones and isoflavones, respectively [59]. Furthermore, flavanone-3-hydroxylase (F3H), flavonol 3′-hydroxylase (F3′H), and flavonol 3′5′-hydroxylase (F3′5′H) catalyzed naringenin to generate dihydro-myricetin, dihydro-kaempferol, and dihydro-quercetin, respectively [60]. The FLS (flavonol synthase) converted dihydroflavonols into flavonols (kaempferol, quercetin, and myricetin), which was catalyzed by the dihydroflavonol 4-reductase (DFR) to generate leucoanthocyanidins [61], which was catalyzed by leucoanthocyanidin dioxygenase (LDOX) to produce anthocyanidins [62]. Anthocyanidins and leucoanthocyanidins were further converted to proanthocyanidins catalyzed by leucoanthocyanidin reductase (LAR) and anthocyanidin reductase (ANR), respectively [63]. Modification of anthocyanins is responsible for the stabilization of vacuolar anthocyanins, including glycosylation, methylation, and acylation [64].
Figure 2. Flavonoid synthesis pathway. CHS (chalcone synthase) can catalyze three molecules of malonyl-CoA and one molecule of p-coumaroyl-CoA to form naringeninchalcone [56]. Malonyl-CoA is an important precursor for the synthesis of natural products, including flavonoids and polyketides [57]. CHI (chalcone isomerase) converted naringenin-chalcone into flavanones [58]. Naringenin, as an important flavonoid skeleton, is catalyzed by FNSI and FNS II (flavone synthase I and flavone synthase II) and IFS (isoflavone synthase) to form flavones and isoflavones, respectively [59]. Furthermore, flavanone-3-hydroxylase (F3H), flavonol 3′-hydroxylase (F3′H), and flavonol 3′5′-hydroxylase (F3′5′H) catalyzed naringenin to generate dihydro-myricetin, dihydro-kaempferol, and dihydro-quercetin, respectively [60]. The FLS (flavonol synthase) converted dihydroflavonols into flavonols (kaempferol, quercetin, and myricetin), which was catalyzed by the dihydroflavonol 4-reductase (DFR) to generate leucoanthocyanidins [61], which was catalyzed by leucoanthocyanidin dioxygenase (LDOX) to produce anthocyanidins [62]. Anthocyanidins and leucoanthocyanidins were further converted to proanthocyanidins catalyzed by leucoanthocyanidin reductase (LAR) and anthocyanidin reductase (ANR), respectively [63]. Modification of anthocyanins is responsible for the stabilization of vacuolar anthocyanins, including glycosylation, methylation, and acylation [64].
Molecules 28 04982 g002
Figure 3. Mechanisms of flavonoid antioxidant activity in vitro. (a) Direct scavenging of reactive oxygen species (ROS); (b) Activation of antioxidant enzymes; (c) Activation of metal-chelating activity; (d) Increasing α-tocopheryl radical levels; (e) Inhibiting NAPDH oxidases; (f) Mitigation of oxidative stress caused by NO; (g) Increasing uric acid levels; (h) Increasing antioxidant properties of low-molecular-weight antioxidants.
Figure 3. Mechanisms of flavonoid antioxidant activity in vitro. (a) Direct scavenging of reactive oxygen species (ROS); (b) Activation of antioxidant enzymes; (c) Activation of metal-chelating activity; (d) Increasing α-tocopheryl radical levels; (e) Inhibiting NAPDH oxidases; (f) Mitigation of oxidative stress caused by NO; (g) Increasing uric acid levels; (h) Increasing antioxidant properties of low-molecular-weight antioxidants.
Molecules 28 04982 g003
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

Chen, S.; Wang, X.; Cheng, Y.; Gao, H.; Chen, X. A Review of Classification, Biosynthesis, Biological Activities and Potential Applications of Flavonoids. Molecules 2023, 28, 4982. https://doi.org/10.3390/molecules28134982

AMA Style

Chen S, Wang X, Cheng Y, Gao H, Chen X. A Review of Classification, Biosynthesis, Biological Activities and Potential Applications of Flavonoids. Molecules. 2023; 28(13):4982. https://doi.org/10.3390/molecules28134982

Chicago/Turabian Style

Chen, Shen, Xiaojing Wang, Yu Cheng, Hongsheng Gao, and Xuehao Chen. 2023. "A Review of Classification, Biosynthesis, Biological Activities and Potential Applications of Flavonoids" Molecules 28, no. 13: 4982. https://doi.org/10.3390/molecules28134982

Article Metrics

Back to TopTop