Next Article in Journal
A Mechanistic Study of the Osteogenic Effect of Arecoline in an Osteoporosis Model: Inhibition of Iron Overload-Induced Osteogenesis by Promoting Heme Oxygenase-1 Expression
Previous Article in Journal
Chitosan-Loaded Lagenaria siceraria and Thymus vulgaris Potentiate Antibacterial, Antioxidant, and Immunomodulatory Activities against Extensive Drug-Resistant Pseudomonas aeruginosa and Vancomycin-Resistant Staphylococcus aureus: In Vitro and In Vivo Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dietary Polyphenols: Review on Chemistry/Sources, Bioavailability/Metabolism, Antioxidant Effects, and Their Role in Disease Management

1
Department of Pharmaceutical Sciences, School of Biotechnology and Pharmaceutical Sciences, Vignan’s Foundation for Science, Technology & Research (Deemed to be University), Guntur 522213, India
2
Department of Pharmaceutical Sciences & Technology, Birla Institute of Technology, Ranchi 835215, India
3
School of Biotechnology and Bioinformatics, D Y Patil Deemed to be University, Navi Mumbai 400614, India
4
Department of Medical Laboratory Sciences, College of Applied Medical Sciences, Majmaah University, Al Majmaah 11952, Saudi Arabia
*
Author to whom correspondence should be addressed.
Antioxidants 2024, 13(4), 429; https://doi.org/10.3390/antiox13040429
Submission received: 7 February 2024 / Revised: 22 March 2024 / Accepted: 26 March 2024 / Published: 30 March 2024

Abstract

:
Polyphenols, as secondary metabolites ubiquitous in plant sources, have emerged as pivotal bioactive compounds with far-reaching implications for human health. Plant polyphenols exhibit direct or indirect associations with biomolecules capable of modulating diverse physiological pathways. Due to their inherent abundance and structural diversity, polyphenols have garnered substantial attention from both the scientific and clinical communities. The review begins by providing an in-depth analysis of the chemical intricacies of polyphenols, shedding light on their structural diversity and the implications of such diversity on their biological activities. Subsequently, an exploration of the dietary origins of polyphenols elucidates the natural plant-based sources that contribute to their global availability. The discussion extends to the bioavailability and metabolism of polyphenols within the human body, unraveling the complex journey from ingestion to systemic effects. A central focus of the review is dedicated to unravelling the antioxidant effects of polyphenols, highlighting their role in combating oxidative stress and associated health conditions. The comprehensive analysis encompasses their impact on diverse health concerns such as hypertension, allergies, aging, and chronic diseases like heart stroke and diabetes. Insights into the global beneficial effects of polyphenols further underscore their potential as preventive and therapeutic agents. This review article critically examines the multifaceted aspects of dietary polyphenols, encompassing their chemistry, dietary origins, bioavailability/metabolism dynamics, and profound antioxidant effects. The synthesis of information presented herein aims to provide a valuable resource for researchers, clinicians, and health enthusiasts, fostering a deeper understanding of the intricate relationship between polyphenols and human health.

1. Introduction

The overall lifestyle of an individual is compromised due to professional competition, irregular food intake, and social-behavioral habits like drinking and smoking [1]. Their significant impact on the biological processes has led to disorders like hypertension, diabetics, aging, neurodegenerative diseases, heart diseases, etc. [1,2]. Such irregularities trigger the production of free radicals causing oxidative stress, a lead mechanism resulting in the imbalance between prooxidants and antioxidants [2]. Many clinicians and scientists identified polyphenols to be the key component with anti-oxidant effects and play a major role in overcoming these disorders [1]. Being the secondary metabolite of plants, they are abundant in nature with a wide range of structural differences, which has led to their categorization [3].
Antioxidants attenuate the effects of the activities of pro-oxidants or free radicals by quenching their oxidative activity. These free radicals are species capable of non-dependent existence with one or more unpaired electrons [4]. Oxidative stress is characterized by the generation of reactive oxygen species [5] that may be involved in cellular reactions with DNA, lipids, and proteins, among other macromolecules [6]. Such cellular reactions have been attributed to the cause of most diseases, such as inflammation, cancer, brain dysfunction, cardiovascular disease, organ failures, and the general decline in the human immune system [7]. Even though various synthetic antioxidant agents have been available over the years, their cytotoxicity has numerous side effects [8]. Due to this, there is a need for various kinds of phyto-antioxidants that are fruit, vegetable, and herbal plant-based, have hardly any side effects, and are cheaper than synthetic agents to be used as a source of alternative medicine [9]. These phyto-antioxidants effectively inhibit oxidative species involved in various diseases, offering potential therapeutic benefits against conditions like cardiovascular disease, neurodegenerative disorders, and cancer. Additionally, their anti-inflammatory properties make them promising candidates for managing inflammatory conditions such as obesity and type 2 diabetes mellitus [8].
Polyphenols stand out as the leading stars in the realm of pharmacological therapeutics, showcasing their prowess as primary components in our daily dietary intake [10]. Aside from being an essential antioxidant agent, this group of compounds has been widely reported to possess anti-allergic, antihypertensive, anti-inflammatory, anticancer, antiviral, and antimicrobial potencies [11] of which most of these biological activities are consequences of the antioxidative effect of this class of organic molecules. As a group of compounds that exhibit antioxidant activities, polyphenols prevent aging [12] and degenerative diseases such as cardiovascular, cancer and neurodegenerative diseases [13]. Most fruits and medicinal plants with high polyphenolic content exhibit antioxidant potency. Even though the basic structural unit in polyphenols is the phenolic ring [14], they are composed of a wide range of complex structures with structural features that enhance their antioxidant potency [15].
We posit that the heterogeneity in molecular architecture inherent to polyphenols underpins their classification and modulates their source-dependent bioavailability, thereby dictating their physiological and pharmacokinetic profiles. This structural variation is hypothesized to be pivotal in mediating the interaction of polyphenols with cellular pathways, which is crucial for their metabolic assimilation and therapeutic efficacy in mitigating various pathologies. Furthermore, the distinct chemical signatures of polyphenols are anticipated to confer differential impacts on food quality and commercial viability, as well as on the epigenome, suggesting a potential for modulating gene expression. Consequently, this hypothesis integrates the chemical individuality of polyphenols with their biotransformation, health-promoting attributes, economic significance, and epigenetic influence, proposing a comprehensive and synergistic impact of these bioactive compounds on both clinical and commercial paradigms. This refined hypothesis encapsulates the scientific breadth of dietary polyphenols, linking their chemical diversity to a spectrum of biological and economic consequences.
The overall process of literature search strategy and study selection was done according to PRISMA (Preferred reporting items for systematic review and meta-analysis) guidelines and is presented in Figure 1. The search terms and keywords for the study selection were dietary polyphenols OR food polyphenols AND chemistry and sources AND health and disease management AND antioxidant effects AND bioavailability and metabolism. The in vitro and in vivo (animal) studies along with pharmacological activity were used as further criteria for literature search. The search was carried out for last 40 years (1985 to 2024) of publications. Three independent reviewers (G.R., S.G. and S.C.) conducted the literature search in the scientific databases and assessed/verified the eligibility of the studies based on the title and abstract. Disagreement between reviewers was sorted out through consultation with fifth and sixth reviewers (M.R. and R.P.S.) to arrive at a consensus. The inclusion criteria were (i) studies involving dietary or food polyphenols and their role in health and disease management, (ii) studies reported on bioavailability and metabolism of dietary polyphenols, (iii) studies performed in polyphenol-rich extracts and their effects in in vitro or in vivo models, (iv) studies conducted on antioxidant role and pharmacological effects of health-promoting polyphenols, and (v) studies published from 1985 to December 2024 (40 years, both years included). The exclusion criteria were (i) studies involving dietary polyphenols intake as food supplements, food supplements and their role in disease management, (ii) studies on systematic reviews, meta-analysis and case reports, (iii) studies involving clinical data and human subjects, (iv) papers published before 1985, and (v) published articles in a language different from English. A total of 3134 published records were identified from the database search (PubMed Web of Science, Embase, ScienceDirect, Scopus) and other sources. After removing 109 duplicate articles, 3025 papers were screened and 2031 were excluded based on title and/or abstract. The full-text of eligible studies (n = 994) was read, 299 articles were excluded because not meeting the inclusion criteria (n = 367) or not of interest/pertinent/relevant (n = 176). At the end of the selection process, 451 papers were finally selected and included in the study.

2. Chemistry of Polyphenols

The classification of polyphenols is highly based on the topology of the phenolic ring and the side atoms or molecules attached to it [14]. Due to the varied topology and diverse availability, polyphenols are classified as follows:

2.1. Phenolic Acids

The phenolic structure with the carbon backbone of these acids consists of CO2H and OH group on the C1 and C6 carbon atoms [16]. They can either be hydroxybenzoic acids or hydroxycinnamic acids type, with a non-phenolic nature [12,17]. Some of the most common of this class are ferulic acid, sinapic acid, caffeic acid, chlorogenic acid, cinnamic acids, vanillic acid, gallic acid, syringic acid, protocatechuic acid, and coumaric acid [12]. The shikimate pathway from L-tyrosine has been reported as the biosynthetic source of phenolic acids [18], where biosynthetic sequences involving deamination [19], hydroxylation and methylation give rise to the acids [20].

2.2. Flavonoids

Structurally, they have the general backbone C6-C3-C6. Both C6 units (A and B ring) are phenolics [21]. Consequent upon the variation in the chromane ring (ring C), flavonoids have been classified as flavanones (hesperitin, naringenin, hesperidin, eriodyctiol), flavones (apigenin, tangeritin, baicalein, profilin), chalcones (phloretin, arbutin, phlorizin, chalco-naringenin), flavonols (quercetin, myricetin, rutin, morin, kaempferol), isoflavonoids (genistein, genistin, daidzein, daidzein), and anthocyanin (cyanidin, malvidin, delphinidin, peonidin [22,23,24]. C-glucosides and O-glucosides, integral subclasses of flavonoids, are characterized by a glucose molecule linked to a flavonoid backbone via a glycosidic bond, distinguishing them within the broader category of polyphenolic compounds found abundantly in plants. These glucosides are prevalent in various plant sources, including fruits, vegetables, grains, and herbs, contributing to their color, flavor, and nutritional profile. Exhibiting antioxidant, anti-inflammatory, and anticancer properties, C and O glucosides play crucial roles in human health by scavenging free radicals, modulating signaling pathways, and inhibiting enzymes involved in inflammation and carcinogenesis. Notable examples such as vitexin, isovitexin, quercetin-3-O-glucoside (isoquercetin), and kaempferol-3-O-glucoside (astragalin) exemplify the diversity and significance of these compounds in dietary sources, emphasizing their potential implications for human health and underscoring their importance in flavonoid chemistry [25,26].

2.3. Stilbenes

Stilbenes which are synthesized via the phenylpropanoid pathway is also a class of naturally occurring organic molecules distributed in fruits and other plant sources [27]. Structurally, they are made up of two aromatic rings (A and B) linked by a methylene bridge [28]. Some known stilbenes include resveratrol, piceatannol, pinosylvin, and rhapontigenin, among others [29]. Resveratrol is one of the most widely studied stilbene due to its wide range of pharmacological activities [28].

2.4. Tannins

These compounds have a molecular weight range between 500 to 30,000 Da and are distributed in various plant foods and beverages [30]. They can be either water soluble or insoluble, depending on their molecular weight. Based on their chemical structure, they are classified as condensed tannins, hydrolyzable tannins, phloro-tannins and complex tannins. The shikimate pathway is reported to be the biosynthetic route for the generation of tannins. Some of the known tannins include gallotannins, ellagitannins, proanthocyanidins, etc. [31].

2.5. Lignans

Lignans are formed by the union of two cinnamic acid molecules. Their structure consists of two benzylbutane, and are therefore called diphenolic compounds [32]. Lignans are low molecular weight phenolic compounds present in the cell walls of plants. They are plant components present in the glycosidic form. They provide rigidity to plant cells, facilitate mineral transport through vascular bundles, and protect plant cells from external injury. They are synthesized by the shikimic acid pathway in the plant cells [33].

2.6. Coumarins

Coumarins are a class of naturally occurring lactones composed of a benzopyrone framework [34]. The major classes of coumarins include simple coumarins, furocoumarins, dihydrofurocoumarins, pyranocoumarins (linear and angular), phenylcoumarins, and biscoumarins [35]. Mostly found coumarins are aesculin, aesculetin, umbelliferone, coumestrol, bergapten, psoralen etc. The shikimic acid pathway is the biosynthetic route for coumarins [36].
The sub-classification of each of the polyphenols and structures has been mentioned in Table 1 and Figure 2, and few of them are discussed in the dietary sources, majorly because of their availability.

3. Dietary Sources

According to recent research on their health advantages, it is essential to include polyphenols in the everyday diet [49]. The amount of each polyphenol is determined by the extraction process and in terms of approximate dry and fresh weight [50]. These compounds can be found in a variety of foods, such as berries, apples, citrus fruits, spinach, kale, broccoli, oats, whole grains, tea, coffee, and red wine, offering individuals numerous opportunities to incorporate polyphenols into their diets. Various sources of dietary polyphenols are displayed in Figure 3. The quantity of each polyphenol in these dietary sources can vary significantly, influenced by factors like the extraction process and the source material’s state, whether it is dry or fresh [50]. For example, the concentration of polyphenols may differ between fresh fruits and their dried counterparts, highlighting the importance of considering food processing methods when assessing polyphenol content. Factors like the specific variety and ripeness of fruits and vegetables can also impact polyphenol levels, further underscoring the complexity of estimating polyphenol intake from dietary sources, as shown in Table 2. Understanding the sources and approximate concentrations of polyphenols in foods is crucial for individuals seeking to optimize their dietary intake for health benefits. By incorporating a diverse array of polyphenol-rich foods into their diet, individuals can not only enhance flavor and culinary enjoyment but also naturally obtain these bioactive compounds, known for their antioxidant, anti-inflammatory, and potentially disease-preventive properties. Embracing a diet rich in polyphenol-containing foods can thus be a proactive step toward supporting overall health and well-being. The examples of polyphenols extracted as a dietary source as mentioned in Table 2 and are discussed in the section below.

3.1. Phenolic Acids

They are abundant in nature as herbs [51] and dietary [52] constituents, possess anti-oxidative potencies among others [53], and play a vital role in the control of diseases [54]. Mushrooms have been reported to be one of the major sources of phenolic acids, which substantially account for their therapeutic activities such as antioxidant, antimicrobial, and antitumor potencies [55]. Fruits and cereal grains are the abundant source of caffeic and ferulic acids [56].

3.2. Flavonoids

This is one of the classes of polyphenols that are widely distributed in vegetables, flowers, fruits, grains, wine, and tea [23]. They have variable polyphenolic structures and constitute the major components of polyphenolic compounds [57].

3.3. Stilbenes

Stilbenes are mainly found in Vitis vinifera, Arachis hypogaea, Sorghum bicolor, Polygonum cuspidatum, Rhodomyrtus tomentosa, Rheum undulatum, Melaleuca leucadendron, and Euphorbia lagascae [58]. They are among the class of compounds that give protection to plants against viral and other microbial-related attack [59,60]. Pinosylvin is present in high amounts in the heartwood of pine trees [61].

3.4. Tannins

Berries and their derivatives such as jams, jellies, and juices, are important sources of tannins [62]. They are responsible for imparting astringent and bitter taste to food products [31]. Due to their astringent property, they can alter protein structure, disrupt the activity of digestive enzymes, and form chelates with calcium and ferric ions, thereby reducing their bioavailability [31]. Epigallocatechin and epigallocatechin-3-gallate are the major tannins present in green tea [63].

3.5. Lignans

Dietary sources of lignans include rapeseed, flax seeds, sesame seeds, legumes, rye, barley, vegetables, fruits, fish, meat, etc. [64]. The common lignans present in almonds, cashew nuts, pecans, and other nuts are lariciresinol, matairesinol, secoisolariresinol, cyclolariciresinol, and 7-hydroxymatairesinol [65]. The lignans exert antioxidant, antiviral, and anticancer effects.

3.6. Coumarins

They are present in a variety of plant species belonging to the families of Apiaceae (Umbelliferae), Rutaceae, Asteraceae (Compositae), Fabaceae (Leguminosae), Oleaceae, Moraceae, and Thymelaeaceae [34]. Amongst them, the Apiaceae family is an important source of coumarins. The plants consisting of coumarins are used widely for culinary purposes and as flavoring agents in the food industry [66].
Table 2. Dietary sources of some polyphenols with their concentration in dry and fresh weight.
Table 2. Dietary sources of some polyphenols with their concentration in dry and fresh weight.
PolyphenolDietary SourceConcentration
(~Dry/Fresh Weight)
References
Gallic acidBlack currant7.67 to 39.70 mg/100 mg[56,67]
Protocatechuic acidRaspberry215.28 mg/100 g[68,69]
p-Hydroxybenzoic acidStrawberry2–8 mg/100 g[68,70]
Caffeic acidKiwi0.009–0.04 µg/g[68,71]
Chlorogenic acidCherry1–4 mg/100 g[68,72,73]
Coumaric acidCarrots6.8 mg/100 g[68,72,74]
Sinapic acidApple, Pear, Coffee etc.0.140 mg/g[68,72]
AnthocyaninsAubergine12.08 mg/100 g[75,76]
ApigeninParsley, Chamomile45 mg/g[68]
QuercetinCurly kale8 mg/100 g[77,78]
KaempferolLeek32.5 mg/100 g[77,78]
MyricetinBroccoli, Red wine, Blueberry, Beans, Tomato, Black tea, etc.3.8 to 22.6 mg/L[79,80]
EpigeninCelery19.1 mg/100 g[79,81]
GenisteinSoybeans0.03–0.2 mg/100 g[79]
HesperidinGrapefruit juice0.93 mg/mL[82,83]
NaringeninLemon juice1.77 mg/100 mL[82,83]
CatechinBeans8 to 12 mg/100 g[84,85]
EpicatechinApricot, Cherry, Peach, Blackberry, Apple, Green tea, etc.8.3 mg/100 g[84,86]
MorinGuava-[82]

4. Bioavailability and Metabolism

Polyphenols post-consumption is digested in the stomach. In their native forms, polyphenols are present as conjugates with, majorly, a glycosidic side chain. These conjugations render the absorption of polyphenols by the small intestine rather difficult. Extracellular enzymes, such as glucosidases, help deconjugate the polyphenol, thus aiding in its absorption. Unabsorbed polyphenols are transferred to the large intestine where they undergo microbial modification to facilitate absorption. The deconjugated polyphenols are transferred to the liver via the hepato-portal vein, where most polyphenol metabolism occurs. Polyphenols are reconjugated in the liver with the help of various enzymes and make their way to the target sites. The liver also helps in excretion of excess polyphenols. Polyphenols are excreted in two major ways: urinary excretion by kidneys and biliary excretion, which are further processed and excreted finally as feces. The illustration is depicted in Figure 4 [87].

4.1. Phenolic Acids

The wide distribution of phenolic acids among plant-derived foods has been reported [88]. Coffee is one of the sources of phenolic acids that have been used to monitor the ADME (adsorption, distribution, metabolism and excretion) properties of phenolic acids [89]. The metabolic processes have been revealed to involve; methylation [90], glucuronidation [91], and sulfation [92], which takes place in the small intestine and subsequently in the liver [93], producing conjugates of these metabolic processes [94]. Such metabolites (glucuronidated, methylated, and sulfated phenolic acids) have been found in the biological fluids of patients who ingested coffee [95]. The presence of glucuronide and sulfate conjugates of hydroxycinnamic acids has been reported from the consumption of a particular wine drink [96]. Caffeic, ferulic, and chlorogenic acids are rapidly absorbed from the stomach and intestine in their free form; further, they are conjugated through detoxification enzymes [97]. Metabolism detoxifies the compounds by transforming them into an easily excretable form. However, some metabolites are therapeutically active in themselves [98]. The mode and site of metabolism depend on the phenolic acid’s chemical structure [99]. For instance, the colon is the only site of metabolism for chlorogenic acid [100]. Free phenolic acids are more easily absorbed than the corresponding esters [101].

4.2. Flavonoids

Due to the polyphenolic nature of flavonoid, it is commonly known for its antioxidant activity. Through flavonoid reaction with free radicals, reactive oxygen species are stabilized [102]. They prevent lipid peroxidation and the formation of superoxide via various pathways [103]. With the exception of flavan-3-ols, most flavonoids exist as glycosides [104]. Lower molecular weight proanthocyanidins have been reported to be better superoxide and hydroxyl radical scavengers, and xanthine oxidase inhibitors [105]. Among the class of flavonoids, it has been shown that isoflavones are the most absorbed while flavanol glycosides, flavanones, and flavonols have intermediate absorption. Anthocyanins, proanthocyanidins, and flavanol gallate were reported to have reduced absorption and bioavailability [106]. Kidney bean has recently been shown to have a lower bioavailability than two other (soya bean and faba bean) dietary legumes [107]. It has been revealed that the bioavailability of flavonoids is generally low. However, the bioavailability varies according to the class and molecular size of the flavonoids involved. Low bioavailability is found among high molecular weight flavonoids [108]. Conjugation is one of the metabolic processes found in flavonoids. Flavonoids undergo glucuronidation, sulfation, and methylation to form the corresponding esters, usually without aglycone in the plasma [109]. The small intestine is the first site (phase I metabolism) of flavonoid conjugation, and then the liver, where phase II metabolism occurs. Unabsorbed metabolites undergo further modifications in the microflora colon; the gut microbiota plays a vital role in metabolizing flavonoids, breaking them down into smaller, more bioactive compounds. Through processes like deglycosylation and ring cleavage, gut bacteria transform flavonoids into metabolites with enhanced health benefits. This microbial metabolism influences various physiological processes, such as antioxidant activity and inflammation modulation. The composition and activity of the gut microbiota vary between individuals, affecting the extent and pattern of flavonoid metabolism. Understanding this interaction is essential for harnessing the full therapeutic potential of flavonoids through dietary or microbiome-targeted interventions [110]. Glycosylation has also been noted as a major factor affecting bioavailability; quercetin glycosides were absorbed ten times faster than the corresponding rutinosides [111,112] in humans. This flavonoid possesses poor intestinal absorption and is only absorbed by the enterohepatic system when hydrolyzed into its constituents, quercetin and sugar moiety, by the microflora of the lower bowel. Results from an animal study using rats identified 3-methoxy-4-hydroxy phenylacetic acid, 3,4-dihydroxy phenylacetic acid, and m-hydroxy phenyl acetic acid as metabolites after oral administration of quercetin, while 3-O-methylquercetin has also been identified as of its metabolites from bile [113]. The bioavailability of quercetin has been reported to be low. This is the quantity of flavonoid that is present and unchanged in the systemic circulation [114]. The low bioavailability has been attributed to its high metabolism rate [115,116]. There are some reported flavonoids that undergo metabolic conversion in the enterocytes and liver [117], and such flavonoids would only have the metabolites found in the plasma [118]. Microemulsion, nano- delivery systems, microencapsulation and enzymatic methylation are some of the methods employed to combat the challenge posed by flavonoid bioavailability [119,120,121].

4.3. Stilbenes

Stilbenes also have a record of low bioavailability; however, they are distributed to various tissues in the form of glucuronides and sulfate conjugates in the tissues and plasma [122]. Among these, glucuronide has been found to be predominant as a metabolite, and its formation is more rapid [123]. The oral absorption of resveratrol in humans is found to be 75%; however, its bioavailability was found to be only 1% [124]. About 90% of resveratrol is subjected to fermentation by gut microbes. After absorption, the resveratrol conjugates enter the systemic circulation and penetrate the target tissues to execute their physiological action. In clinical trials it was found that a high rate of metabolic breakdown of resveratrol into resveratrol-3-O-sulfate, resveratrol-4′-O-glucuronide, and resveratrol-3-O-glucuronide limits its use in pharmacological applications [125,126]. Several research studies have been carried out to improve stilbenes’ bioavailability, solubility, and stability. The bioavailability of rhaponticin, a methylated derivative of resveratrol, was improved by modifying liposomes with polyethylene glycol. It was observed that its distribution was lowered in the gastrointestinal tract thereby enhancing the plasma concentration by 4.5 times [127]. Piceatannol, a monomer derivative of resveratrol, exhibited a higher absorption rate (maximum serum concentration was 2.6 times higher) and better metabolic stability (higher area under the curve) than resveratrol [128]. It was demonstrated that oxy-resveratrol was less bioavailable in comparison to resveratrol and underwent extensive hepatic metabolism to be excreted in bile and urine [129]. Hence, the bioavailability of the stilbenes largely depends on their structures.

4.4. Tannins

There are few reports on the bioavailability of tannins; however, it has been reported that procyanidins from apple juice gave a recovery of about 90% after consumption [130]. The bioavailability of tannins largely depends on factors such as chemical and biological degradation, gut and liver metabolism, membrane permeability and many more [131]. The degree of polymerization and solubility of tannins affects their rate of absorption, as highly polymerized tannins cannot be absorbed. Low solubility, formation of insoluble complexes and irreversible binding with DNA and proteins limit the absorption of ellagic acid [132]. In the small intestine, high molecular weight proanthocyanidins form complexes, thus interfering with their digestion. On the other hand, absorption of gallic acid from the small intestine takes place either by rapid permeation or in the form of conjugates [133]. The colonic bacteria help in the metabolism of hydrolyzable tannins. It is reported that ellagitannin is hydrolyzed by microbiota in the large intestine to ellagic acid and urolithin B [134]. Therefore, it is observed that metabolism and absorption of tannins take place at different parts of the gastrointestinal tract. Tannins have anti-nutritional properties and can retard the absorption of vitamins and minerals. They have also shown significant antioxidant activity [135]. Some studies report that tannins can exert pharmacological effects either as absorbable metabolite or as an un-absorbable complex [136].

4.5. Lignans

Lignans are metabolized into mammalian lignans by intestinal bacteria. Secoisolariciresinol diglucoside, a lignan present in flax seeds, is converted in the human colon to enterodiol, and entero-lactone exhibits anticancer effects [137]. The conversion of lignans to enterolignans largely depends on the activity of gut bacteria. The amount of precursor intake, gut microbial activity, and degree of conjugation determines enterolignan exposure. The lignans are conjugated as glucuronides or sulfates in the intestinal epithelium and liver; later, they are excreted in the urine and bile. The re-excreted compounds are then deconjugated by bacterial β-glucuronidase to undergo enterohepatic recycling [138]. The enterolignans take approximately 8–10 h to appear in the circulation after ingestion, while plant lignans take only 2 h [139].

4.6. Coumarins

These compounds are attributed to many pharmacological properties such as antioxidant, antimicrobial, anticoagulant, anti-inflammatory, anti-hypertensive, neuroprotective, etc. [140]. The pharmacokinetic profile of aesculin and aesculetin were studied in rats. After oral administration of aesculin (120 mg/kg), the Cmax and AUC values were 340.3 ng/mL and 377.3 h ng/mL, respectively; on the other hand, for aesculetin it was found to be 316.5 ng/mL and 1276.5 h ng/mL, respectively. Further, the oral bioavailability of aesculin was found to be 0.62%, possibly due to poor oral absorption and first-pass metabolism [141]. In a non-compartmental analysis, the mean oral bioavailability of aesculetin was reported to be 19% [142]. The coumarin bergapten obtained an oral bioavailability of 69.5 ± 44.2% at a 15 mg/kg concentration and had good absorption from the gastrointestinal tract. However, its presence in the rat urine was detected only after 72 h [143].

5. Antioxidant Effects of Dietary Polyphenols

Epidemiological studies have shown that consumption of dietary polyphenols can reduce the risk of chronic diseases [10]. Phyto-antioxidants are the main contributors to the total antioxidant activity of fruits and vegetables [144]. They protect the cells from oxidative damage and increase the plasma antioxidant capacity [145].
The intricate relationship between flavonoids and the gut microbiota underscores the pivotal role of microbial metabolism in shaping the bioavailability and health effects of these polyphenolic compounds. Within the large intestine, flavonoids undergo extensive biotransformation by the diverse microbial community, yielding metabolites with potent physiological activities. Processes such as deglycosylation, ring cleavage, and demethylation lead to the generation of bioactive compounds that exert antioxidant, anti-inflammatory, and metabolic modulatory effects. This microbial metabolism enhances the bioavailability of flavonoids and influences their distribution and efficacy in various tissues. Understanding the dynamic interplay between flavonoids and gut microbiota holds promise for unlocking novel therapeutic strategies aimed at optimizing human health through dietary interventions and microbiome-targeted approaches [110,145].
Moreover, human health is greatly affected by the gut microbiota, and a case study was performed to comprehend the impact of polyphenols on it. According to the Ma et al., 2020 case study, the organisms provided with controlled administration of polyphenols were observed to have an abundance of Lactobacillus and Bifidobacterium and reduced levels of pathogenic Clostridium [49]. More studies related to anti-cancer properties have highlighted the positive effects of polyphenols. The antioxidant potential of polyphenols has been reported in various in vitro and in vivo models; a few are mentioned in Table 3 [146].

5.1. Phenolic Acids

In in vivo and in vitro studies, the antioxidant activity of chlorogenic and caffeic acid has been reported, where caffeic acid exhibited more antioxidant effects than chlorogenic acid. It was found that both compounds show protective effects against I/R injury, but the uptake of caffeic acid by Cao-2 cells was much more than that of chlorogenic acid [155]. The chlorogenic acid isomers, neochlorogenic acid and crypto chlorogenic acid present in Prunus domestica L. have been reported to exhibit antioxidant activities. Both the isomers showed superoxide radical scavenging activity and inhibited oxidation of methyl linoleate [156]. This scavenging effect of crypto chlorogenic acid, which is also a major component of mulberry leaf, is responsible for its protective effect on β-cells function in diabetes [157]. In another study, the three isomers of caffeoylquinic acid, namely 3-O-caffeoylquinic acid, 4-O-caffeoylquinic acid, and 5-O-caffeoylquinic acid, with three isomers of di caffeoylquinic acid namely, 3,5-dicaffeoyl-quinic acid, 3,4-dicaffeoylquinic acid, and 4,5-dicaffeoylquinic acid exhibited antioxidant potency alongside protected DNA from damage [158]. However, the dicaffeoylquinic acid isomers exhibited a better antioxidant effect as they contain more hydroxyl groups. On the contrary, the corresponding caffeoyl-quinic acid isomers exhibited similar antioxidant effects [159].
Further, chlorogenic acid, alongside its associated moieties, caffeic and quinic acids, has been reported to significantly act against rat hepatoma cell line (AH109) [160]. However, they were observed not to affect the proliferation of the hepatoma. A structure-activity relationship observation depicts that caffeic and quinic acid confers an additive inhibitory potency on chlorogenic acid following the suppression percentage recorded for each [161]. The observations also suggested that the 3,4-dihydroxy group in caffeic acid is partly responsible for its anti-invasive activity [162]. Additionally, in the carrageenan-induced [163] inflammation model, oral administration of chlorogenic acid resulted in activity that was comparable to indomethacin, the reference drug [164]. El-khadragy and colleagues revealed in recent research that chlorogenic acid, through its antioxidant properties, improved the testicular damage induced by arsenic [165]. Also, chlorogenic acid has been revealed to possess anti-obesity [166], antidiabetic [167], antimicrobial [168], and antihypertensive [169] effects. The antidiabetic activity of this compound has been linked to its involvement with glucose metabolism [170].
Furthermore, YES-10® is a combination therapy of Erigeron annus (L.). PERS. and Clematis mandshurica RUPR. (CMR) leaves which are rich in chlorogenic acid and scutellarin [171] have been reported to increase superoxide dismutase (SODs) in CA1 pyramidal neurons [172]. It also protected the CA1 from TI injury. This particular feature has proposed YES-10® as a candidate for the development of drug to guard the brain against ischemic damage [173].
Sinapic acid and the corresponding alkyl esters are among the phenolic acids that have exhibited considerable antioxidant effects. Experimental reports from DPPH and FRAP assay suggest that sinapic acid has a higher activity than the corresponding alkyl esters [174]. This polyphenolic acid is present in coffee [175], vegetables [176] and cereal grains. An isolated sinapic ester derivative, 6-O-sinapoyl sucrose was characterized on isolation and revealed antioxidant activity using DPPH free radical scavenging test [177]. A range of sinapic acid derivatives found in rapeseed have been shown to possess antioxidant potencies [178]. The anti-inflammatory, anticancer, neuroprotective, antimutagenic, and antiglycemic activities of the sinapic acid derivatives have been attributed to their antioxidant potencies [179]. Sinapic acid has been described to attenuate high blood pressure, vascular dysfunction, cardiac fibrosis and myocardial by preventing oxidative stress through its antioxidative potency [180]. SH-SY5Y is one of the human-derived cells employed in neurodegenerative-associated research [181]. More recently, it has been shown to defend SH-SY5Y human neuroblastoma cells against 6-hydroxy dopamine-induced neurotoxicity. It was achieved through a significant oxidative blockage and reactive oxygen species (ROS) attenuation overproduction [182].
Ferulic acid, over the years, has been known for its antioxidant activities and stands as a major feature of why it has been a choice as cosmetic and food additives and is well found and indispensable in other industrial applications [183]. This class of phenolics has been described as the most effective among other phenolic acids whose scavenging potency was tested by Kikuzaki and colleagues [184]. As a result of the antioxidant potencies of phenolic acids, this class of compounds has also exhibited antioxidant-related activities [185]. Ferulic acid has been found to reduce memory dysfunction while exerting protective effects against oxidative stress as well as apoptosis resulting from IRI injury [186] through the inhibition of the TLR4/MYD83 signaling pathway [187]. This activity results from the activation of the P38 MAPK-mediated signal cascade [188]. The promotion of functional recovery from ischemic injury induced by middle cerebral artery occlusion in rats by an extract from rice is attributed to its ferulic acid component [189]. It also inhibits nitric oxide synthase proteins [190]. Solid-state fermentation has been used to release ferulic acid in wheat bran which exhibited a better scavenging effect than the free ferulic acid [191]. Kelainy et al. investigated the activity of ferulic acid against lead-induced oxidative stress [192]. In this study, animals were orally administered with lead acetate for 10 days at a dose of 20 mg/kg. Oxidative stress was observed in the form of excess production of LPO, ROS, MDA, and DNA damage alongside reduced levels of sex hormones [193]. Animals treated with ferulic acid exhibited no DNA damage such as DNA apoptotic fragmentation, compared to the control group. Generally, the observation recorded for ferulic acid-treated animals revealed that the compound possesses a protective activity against lead acetate-induced oxidative stress [194]. The result of the study conducted by Wang et al. revealed the antioxidative potency of ferulic acid and its ability to induce lipid metabolism in weaned piglets through the usage of dietary ferulic acid supplementation [195]. Oral administration of ferulic acid or ethyl ferulate has been proved under experimental investigation to cause retinal protective effect in a sodium iodate-induced model of retinal degeneration [196] in human retinal pigment epithelial cell lines. It has also reduced sodium iodate-induced retinal degeneration’s morphological and functional features [197]. An in vitro study using a set of biophysical methods has revealed the protective function of ferulic acid on amyloid formation of bovine β-lactoglobulin, which is associated with neurodegenerative disorders among other diseases such as systemic amyloidosis [198]. Ferulic acid and ferulic derivatives synthesized by Lambruschini and colleagues were reported to possess a protective potency of human endothelial cord vein cells against the oxidative damage of hydrogen peroxide [199]. The use of supplements containing this polyphenol in oxidized fish oil-induced oxidative stress, has been reported to cause antioxidative effects in tilapia [200]. It has also been shown to have a neuroprotective effect on amyloid-beta (Aβ) neurotoxicity-induced Alzheimer’s disease (AD) [201].
The cinnamic acid present in Barringtonia asiatica stem-bark has also been reported to have an antioxidant effect and antifungal activity against Aspergillus niger, Fusarium oxysporium, Candida tropicalis, and Aspergillus flavin [202]. Among the 23 cinnamic ester derivatives [203] screened against C. albicans strains, only methyl caffeate and methyl 2-nitro cinnamate exhibited the largest antifungal activity [204]. Various cinnamic acid derivatives have recently been reported to possess anticancer, anti-TB, and antimalarial activities [205]. Protocatechuic acid, due to its antioxidant properties, has been found to be beneficial in cancer, ulcer, fibrosis [206], inflammation, [207] atherosclerosis, and viral infections [208]. It also protects the cardiac and kidneys from oxidative stress [209].
Gallic acid and its derivatives have been reported for their antioxidant properties. They scavenged hypochlorous acid. Gallic acid lauryl ester in ethanol reduced the peroxidation of ox brain phospholipids. However, gallic acid, its methyl ester and propyl gallate have exhibited a prooxidative effect by enhancing sugar deoxyribose damage in the presence of hydrogen peroxide and ferric-EDTA [210,211]. The reduction of ethanol-induced gastric ulcers by gallic acid in rats has been associated with the Nrf2/HO-1 antioxidative signaling pathway and anti-apoptosis function [212]. Its antioxidant defense mechanism has exhibited a protective effect against cadmium chloride-induced alterations in Wistar rats [213].

5.2. Flavonoids

Many flavonoids have been reported for their antioxidant activity [214]. Pekkarinen and coworkers have reported the antioxidant activity of quercetin, myricetin, kaempferol, and (+)-catechin in methyl linoleate [215]. All the flavonoids were observed to have inhibited the formation of hydrogen peroxides in methyl linoleate, although kaempferol alongside rutin exhibited a relatively weak antioxidative response, while quercetin and myricetin showed a better inhibition in comparison to α-tocopherol [215]. The exceptional antioxidant potency of quercetin has been attributed to the presence and structural nature of the C-ring. The double bond between carbon 2 and 3 in the C-ring and the presence of the keto group in C4 are reportedly responsible for the antioxidant potencies of quercetin [216]. The absence of a hydroxyl group at the C3 which is present in quercetin and (+)-catechin accounts for the relatively low antioxidative effects of rutin. It has also been proven in this study that the antioxidant activity of flavonoids is independent of their solubility, as the more hydrophobic quercetin had a better hydroperoxides formation inhibition than the hydrophilic glycoside rutin. Hence, their antioxidant activity was attributed to both the presence of the C3 hydroxyl group, and the metal chelation potency of the complex formed by the C4 and C3 keto group and hydroxyl group, respectively [217]. In another structural activity relationship report, it was revealed that 31, 41-dihydroxy catechin in the β-ring, 2,3 unsaturated alongside the oxo functional group at the position four in the C-ring of the flavonoids, hyperoxides, and myricetin have been attributed to the antioxidant potency of the flavonoid [218]. It has been noted that the free radical scavenging effect of flavonoids is a result of the formation of less reactive phenoxyl radicals through electron or hydrogen donation, hence inhibiting the hydrogen peroxide-driven Fenton reaction [219]. Their chelating ability has been reported as a consequence of the 4-carbonyl, 5-hydroxyl, and catechol groups on the β-ring [220].
Among six quercetin derivatives that were collectively studied for their antioxidant activity using the in vitro method by DPPH and FRAP assay, quercetin, tamarixetin, isorhamnetin, and quercetin-3-O-glucuronide exhibited significant antioxidant potency [221]. Quercetin acts mechanistically in various ways towards preventing oxidative stress. It inhibits the activation of NF-κB and MAPK signaling pathways alongside the expression of apoptosis-related proteins brought about by LPS/d-GUIN [222], enhances Nrf2/GSH antioxidant signal transduction pathways [223], regulation of phosphoinositide-3-kinasepathway [224]. Other mechanisms include the regulation of TLR2 signaling pathways [225], and the protection of BEAS-2B cells from Cr (VI) induction through targeting of miR-21-PDCD4 signaling [226]. Quercetin also enhances Nrf2-ARE signaling [227] and controls the transcription activities of NF-κB and AP-1 [228]. In order to improve the bioavailability of quercetin, modification to the parent structure has been performed [229] through derivatization and recombination with other active groups [230]. Quercetin metallic complexes have also been shown to possess better antioxidant activity than quercetin, the parent compound [231]. Three quercetin cyclodextrin complexes, β-cyclodextrin, sulfobutyl ether-β-cyclodextrin, and hydroxy propyl-β-cyclodextrin, showed good reactivity with DPPH [232]. Copper complexed quercetin has been evaluated using DPPH. The complex exhibited better activity than the free flavonoid [233]. Wu and colleagues have demonstrated in their research that quercetin-loaded nanoparticles prepared using the nanoprecipitation method exhibit more antioxidant activity than the pure drug [234]. Some metallic complexes of quercetin have also shown good antioxidant potency. The DPPH scavenging potency of the quercetin copper complex is more than that of free flavonoids [235]. However, another study revealed that the scavenging activity of free quercetin decreased after being chelated with tin (II) ion [236]. A study on the determination of quercetin in the plasma of healthy individuals (volunteers) after intake of flavonoid-containing meals revealed that 31 position methylated, sulfate, and glucuronic acid conjugates as the major metabolites of quercetin [237]. Quercetin-cadmium complex has also shown a lesser antioxidant activity than the free quercetin [238]. Quercetin-DNA complex has exhibited a stronger free radical scavenging and antioxidant potency than free quercetin both experimentally and computationally [239]. In order to improve the cellular penetration and antioxidant potency of quercetin, nanoparticles of titanium dioxide in a study have been conjugated with quercetin. The result revealed that the quercetin–nanoparticle conjugate enhances a high quercetin bioavailability and stability while exhibiting antioxidant potency against reactive oxygen species (ROS) without observable toxicity [240]. The antihyperglycemic effect of quercetin in fructose-streptozotocin-induced diabetic rats has been attributed to the flavonoid’s ability to improve the pancreatic antioxidant status [241]. Copper (II) complexes of quercetin and curcumin have been synthesized, characterized, and investigated for their biological activity. Both complexes exhibited significant scavenging activity with no observable toxicity in the eukaryotic experimental model of Saccharomyces cerevisiae [242]. Quercetin and naringenin use antioxidant mechanisms to protect stored boar semen. In research where this was investigated, it was revealed that while naringenin attenuated ROS concentration, quercetin was potent in quenching superoxide, and both molecules have been demonstrated to be potential semen enhancement supplements [243]. The human serum albumin nano-complex of quercetin attenuates hydrogen peroxide-induced neuron mortality through an observable increase in the action of SOD and CAT. The formation of the complex also improved the bioavailability of quercetin [244].
Kaempferol is another naturally occurring flavonoid with known antioxidant potency. It has been reported to act against rotenone-induced Parkinson’s disease (PD) by increasing antioxidant makers in the PD model and SH-SY5Y cells [245]. Aside from the catalytic and anticancer activity of gold nanoparticles synthesized using kaempferol glucoside, the nano complex also exhibited a higher radical scavenging activity when compared to the flavonoid extract [246]. This flavonoid stimulates NF-κB signaling, down-regulates interleukin-6 (IL-6). It also increases the mRNA expression of IL-β, TNF-α, and NF-κB1 [247]. Kaempferol derivatives isolated from Bryophyllum pinnatum exhibited antioxidant and antimicrobial activity. Among the derivatives, the α-rhamnoisorobin derivative demonstrated the highest activity [248]. Kaempferol has been reported to exhibit NO-release T-cell proliferation compared to its glycosides, thereby possessing a better antioxidant activity [249].
Hesperidin, a flavonoid in sweet orange and lemon, is also known for its various biological effects [250]. It has been shown to possess a DPPH scavenging effect [251]. However, aglycone hesperetin demonstrated better antioxidant and neuroprotective activity than the parent compound and thus has been suggested to be a promising drug potential in the treatment and prevention of neurodegenerative disorders [252]. Superoxide dismutase-like activity and antiproliferative activities of hesperidin were improved when complexed with vanadium metal [253]. Hesperidin decreased level of marker enzymes and improved the antioxidant status in nicotine-induced lung toxicity in rats [254]. Mechanistically, hesperidin exhibits antioxidant activities through the inhibition of nitric oxide radicals, lipid peroxidation, attenuation of TBARS, and hydroxy radical inhibition [255]. It also ameliorates spinal cord injury-induced motor dysfunction and neuro-pathological degeneration in rat models through mechanistic effects via the Nrf-2/HO-1 pathway [256].
The research conducted by Wang’s research group on the antioxidant effect of myricetin on oxidative stress–induced cell damage has suggested that myricetin has the potency to inhibit the reactive oxygen species and oxidation-induced adverse effect, thereby protecting the cellular environment from free radical damage while inducing the cellular antioxidant enzyme defense system and repairs the damage of DNA and lipid [257]. A computational study has suggested that myricetin derivatives, 3,4-di-O-alpha-L-rhamnopyranoside are potential antioxidants [258]. Myricetin has been reported to act against other cell oxidation-related disease such as inflammation, diabetes, epilepsy, cardiac disease, cancer and Alzheimer’s disease [259]. Recently, myricetin has been reported to have protected the lymphocytes of healthy individuals and pre-cancerous MGUS patients against hydrogen peroxide induced reactive oxygen species related oxidative damage [260].
Isorhamnetin isolated from the sea buck thorn marc has shown to possess a scavenging activity on DPPH radical, reducing Fe3+ to Fe2+ and chelates iron [261]. The antioxidant potency of this flavonoid has been attributed to its experimentally verified ability of enhancing cortico-hippo campal learning and memory capability in scopolamine induced amnesia in rats by increasing glutathione level, catalase and superoxide activities [262]. Isorhamnetin activates Nrf2-ARE pathway through phosphorylation of ERK1/2, PKCo and AMPK in hepatocytes. It also inhibits ROS production [263]. It has been used to attenuate and treat neurodegenerative disorders [264]. Due to its ability to inhibit the stability of Aβ aggregates thereby prevents cells against Aβ-induced cytotoxicity among other related activities. Isorhamnetin has been described as having the potential of preventing the initiation of Alzheimer’s diseases [265]. Among other highlighted mode of actions, the flavonoid has been reported to activate PBK/AKT signal transduction pathway, activates Nrf2/ARE, Nrf2/HO-1 and ERK pathways and thus inhibits apoptosis, protect cells from peroxide damage [266]. It also induces cell death through ERK/MAPK and ROS pathways in OSCC cells [267]. A comparative study revealed that naringenin has a higher antioxidant capacity than the corresponding glycoside [268].
Naringenin is one of the flavonoids widely distributed in fruits, tomatoes, and citrus fruits [269] with several reported antioxidant, anticancer, anti-inflammatory, neuroprotective, and cardioprotective activities [270]. Oxidative stress and toxicity induced by Lambda-cyhalothrin, a synthetic insecticide in the liver of male rats has been shown to be inhibited by naringenin [271]. This was achieved via modulation of oxidative-nitrosative stress, MMP-9, and cytokine levels. It has been shown to ameliorate diabetic neuropathic pain [272]. It possesses a protective ability of Aβ-induced neuronal cytotoxicity (in vitro) [273]. This flavonoid and its synthetic derivatives have demonstrated the ability to affect antioxidant enzyme activities of erythrocytes and liver in higher cholesterol-fed rats [274]. Baki and colleagues synthesized, characterized and assayed naringenin oxime for its antioxidant activity. The result of the study showed that the derivative exhibited an antioxidant potency higher than the parent compound [275]. Naringenin, curcumin and quercetin were reported to have improved memory retention, learning acquisition and also prevented memory extinction. Through the enhancement of antioxidant concentration and activities, naringenin and quercetin prevent the alteration of the brain’s antioxidant defense system [276]. Y(III) and EU(IV) complexes of naringenin-2-hydroxy benzoyl hydrazone were observed to have an active scavenging effect on OH [277]. It potentiates endogenous antioxidant status during hyperglycemia and attenuates the over-production of nitric oxide-induced inflammation [278]. In regulating doxorubicin–induced liver dysfunction, naringenin has been reported to inhibit ROS-induced lipid peroxidation, ROS production, prevent the reduction of the antioxidant armory; catalase, glutathione, reductase, superoxide dismutase (SOD) glutathione and peroxide (GPx) [279]. This flavonoid has also attenuated neuronal apoptosis in MG-treated NSC34 cells, hence improving antioxidant defense [280].
Apigenin is a naturally occurring flavonoid in vegetables, fruits and medicinal plants. It has a wide-reported range of antioxidant activities [281]. Apigenin has been revealed to act against the oxidative stress induced by carcinogens through attenuation of lipid peroxidation in N-nitroso-diethylamine-induced and phenobarbital-enhanced hepatocellular carcinogenesis [282]. Through modulation of pro-inflammatory cytokines and antioxidant activity, apigenin induced observable change in the level of pituitary-ovarian axis hormones in polycystic ovary syndrome in rat models [283]. In methotrexate–induced hepatotoxicity, apigenin caused alteration in antioxidant, inflammation and lipid peroxidation factors [284]. In a cellular oxidant defense, the flavonoid has inhibited streptozotocin-induced pancreatic β-cell damage [285]. Attenuation of circulating oxidants (LPO, OH) and improvement in the levels SOD, CAT, and GPx, as well as non-enzymic antioxidants have been attributed to luteolin in an azoxymethane-induced colon carcinogenesis [286].
Various studies that have shown Fisetin has antioxidant activity [287]. This antioxidant potential is also attributed to its neuroprotective, anti-inflammatory, and anticarcinogenic effects [288]. Quantum chemical-based calculations have shown that the 3-OH moiety of fisetin possesses the lowest bond dissociation energy followed by the 3,4-OH groups which account for its ability to donate hydrogen to potential free radicals. The B and C rings have been suggested to be more efficient in their antioxidative potencies than the A ring [289]. Bidya and colleagues demonstrated fisetin’s ameliorative effect on cisplatin-induced nephrotoxicity in rats through antioxidant defense and modulation of nuclear factor-factor kappa β (NF-κB) activation [290]. Reportedly, fisetin improves rotenone–induced motor impairments significantly. It has been hypothetically stated that fisetin could enhance mitochondrial enzyme activity, hence attenuating oxidative stress [291]. Fisetin has been observed to inhibit the potentiation of cell death by iron and copper in a therapeutic attempt to lower glutathione levels [292]. Oxidative stress and neuroinflammation induced by D-galactose have reportedly been ameliorated by fisetin in mice brains through the regulation of endogenous antioxidant mechanisms. It suppresses the activated P-JNK/NF-κB pathway [293] alongside the downstream targets and controls endogenous antioxidant mechanisms, hence inhibiting the accumulation of reactive oxygen species (ROS) [294]. This flavonoid enhances the expression of the antioxidant PON 2 through the activation of PPARY, attenuating vascular smooth muscle cell migration and proliferation and ameliorating neointimal hyperplasia after intimal injury [295]. Recently, the anticancer activity of fisetin alongside kaempferol has been attributed to their antioxidant activities [296]. It has been proven to inhibit proinflammatory makers such as NO, iNOS, IL1-β and TNF-α [297]. Through the activation of the reperfusion injury salvage kinase signaling pathway, fisetin has reportedly reduced myocardial ischemia [298]. Fibromyalgia induced by reserpine has been attenuated by fisetin through ROS and serotonergic pathway modulation [299]. The oral bioavailability of fisetin has been reported to enhance its neuroprotective activity in rat models of rotenone-induced Parkinson’s disease [300].
Three of the most common anthocyanidins, cyanidin, delphinidin, and malvidin have been predicted to have a radical scavenging ability using an ab initio computational study [301]. Malvidin has attenuated lipopolysaccharide-induced mitogen-activated protein kinase activation, and mitochondrial depolarization, among other oxidative factors [302]. It has been suggested that malvidin is vital in oxidative and inflammatory stress induced by the THP-1 cell line [303]. Delphinidin has attenuated the oxidative stress induced by hydrogen peroxide in HePG2 cells [304].

5.3. Stilbenes

The antioxidant activity of four stilbenes was evaluated by ORAC assay, where resveratrol exhibited the highest activity, followed by oxyresveratrol, pterostilbene, and pinosylvin. While in the ABTS assay, highest activity was shown by oxyresveratrol, followed by resveratrol, pinosylvin and pterostilbene [305].
Resveratrol has shown oxidative properties through lipid oxidation [306], prevention of DNA damage [307], inhibition of NF-κB activation, increasing the SOD activity, decreasing urinary8-hydroxy-2′-deoxy guanosine, activation of Nrf2 [308]. Resveratrol reduces the risk of cardiovascular disease by preventing lipid peroxidation [309]. It further prevents the oxidation of polyunsaturated fatty acids and is found to have stronger antioxidant activity than α-tocopherol [310]. It up regulates the expression of manganese superoxide dismutase (Mn-SOD) in myoblast cells through a mechanism dependent upon the nuclear factor (erythroid-derived 2)-like 2 (NRF2) thereby reducing the risk of hypertension [308]. Another study demonstrated that administration of resveratrol increases the activity of SOD and catalase [311]. Resveratrol exerts neuroprotective functions by up-regulating the endogenous antioxidant enzymes, inhibiting modulated NF-κB and peroxisome proliferator-activated receptors alpha (PPARα) [312,313]. In addition, due to its antioxidant properties, it counteracts the Aβ toxicity in AD patients [314]. The antidiabetic and anti-inflammatory activity of resveratrol is mainly attributed to its ability to reduce oxidative damage, reduce pro-inflammatory cytokines, and inhibit apoptosis [315]. It also prevents the development and progression of cancer by modulating the cellular pathways involved in inflammation, apoptosis, cell proliferation, metastasis etc. [316]. Similarly, oxyresveratrol showed inhibitory activity against DPPH, hydroxyl, nitric oxide, and hydrogen peroxide radicals [317]. It exerts neuroprotective activity by modulating MAPK and NF-κB signaling pathways [318]. When rat cortical neurons treated with amyloid β protein, were exposed to oxyresveratrol, it reduced ROS production, and glutamate release and suppressed the calcium levels in the cytoplasm, indicating a potential candidate in the treatment of AD [319]. However, this compound showed weak cytotoxicity against various cancer cell lines [320].
In a study, the compound piceatannol (200 mg/kg) showed better antioxidant activity than BHT (200 mg/kg), a synthetic antioxidant used in the preservation of food products. The higher antioxidant potential of piceatannol might be due to the presence of more hydroxyl groups than BHT [321]. Later, the protective effects of piceatannol on lipopolysaccharide (LPS) insult in mouse brain endothelial cell line (bEnd.3) was examined. The results of the study showed that piceatannol upregulated the expression of adhesion molecules (ICAM-1 and VCAM-1) and iNOS in LPS-treated bEnd.3 cells, which suggests its anti-inflammatory and antioxidant activity [322]. In addition, through its antioxidant mechanism piceatannol protected rat cardiomyocyte (H9c2) cells, normal human lung fibroblasts [152] cells, mouse lymphocytic leukemia (L1210) cells, human leukemia (K562) cells, and human promyelocytic leukemia (HL-60) cells from hydrogen peroxide-mediated cytotoxicity [323]. This stilbene provides protection against oxidative stress and inflammation to human retinal pigment epithelial (ARPE-19) cells through induction of heme oxygenase-1 (HO-1) enzyme [324].
Pinosylvin reduced the neutrophil count and decreased the concentration of oxidants for adjuvant arthritis both in the in vitro and in vivo rat models [325]. Another study demonstrated that pinosylvin given to adjuvant arthritic rats at a per oral dose of 50mg/kg improved NF-κB activity in the liver and lung, HO-1 expression and LOX activity in the lung, MCP-1 and F2-isoprostanes levels in the plasma [326].
Rhaponticin and its aglycone form, rhapontigenin inhibits the modulated NF-κB. The presence of oxygen-containing functional groups such as –OH and –OCH3 in the benzene ring might be responsible for their activity [327]. Further rhapontigenin up regulated the expression of SIRT1 in the THP-1 human monocytic cell line [328] which in turn inhibited the activity of NF-κB [329].

5.4. Tannins

The antioxidant potency of green tea has been attributed to one of the major phytoconstituents, catechin, which is widely distributed in green leaves [330]. The antioxidant of this flavonoid has been attributed to various molecular structural features. 1,4-pyrone moiety, alongside the 3-OH group has been described to affect the antioxidant potency of luteolin and catechin [331]. In comparing the metabolites of catechin and epicatechin with the parent compound, it was observed that the cleavage of the C-ring alongside α and β-oxidation increases their antioxidant activity. Among the metabolites, 1-(3′,4′-dihydroxyphenyl)-3-(2″,4″,6″-trihydroxy phenyl) propan-2-ol showed a high antioxidant activity twice as much as catechin, while the highest antioxidant activity was reported for 2-(3′,4′-dihydroxy phenyl) acetic acid whose DPPH and ABTS free scavenging activity equals that of the parent compound while its reducing ability is reported to be significantly higher than both catechin and epicatechin [332]. A density functional theory (DFT) study has suggested that the addition of N(CH3)2, an electron-donating group enhances the antioxidant potency of catechin [333]. The catechin derivative, catechin-5-O-gallate, exhibits antioxidant activity [334]. The antioxidant potency of Tinospora cordifolia has been attributed to the presence of epicatechin in the stem plant [335]. Some antioxidative attributes have been made for the bark extract of the Japanese knot weed rhizome, which contains (−)-epicatechin [336]. (−)-Epicatechin-3-gallate (EGCG), a major anticancer molecule in tea significantly scavenges OH radicals and inhibits the activation of NF-κB. The NF-κB was induced by Cr (IV) and 12-O-tetradecoryl-phorbol-13-acetate (TPA) [337]. In order to improve the lipid membrane permeability of EGCG, a novel lipophilic EGCG derivative, monoalkylated EGCG has been recently synthesized. However, its antioxidant activity decreased against DPPH free radicals, and cellular experiments suggest that the lipid moiety improves the antioxidant capacity of the EGCG derivative [338].
Corilagen, an ellagitannin, induces protective effects in cerebral ischemic injury by inhibiting oxidative stress and promoting angiogenesis by activating the Nrf2 signaling pathway [339]. Such a protective effect has also been recorded for this polyphenol for renal calcium oxalate crystal-induced oxidative stress apoptosis and inflammatory effect through the P13K/Akt and PPAR-γ pathways [340]. In N9 murine microglia cells, it has attenuated oxidative effects in tert-butyl hydroperoxide-induced oxidative stress [341]. The hepatoprotective and antioxidative properties of Terminalia catappa L. have been attributed to the corilagen, one of its bioactive constituents [342]. It reduces sleep deprivation-induced memory impairments through the regulation of NOX2 and Nrf2 activating factors [343]. It also affects MAPK and NF-κB signaling pathways and thus has been reported to ameliorate acetaminophen-induced hepatotoxicity in mouse models [344].

5.5. Lignans

In anin vitrosystem, secoisolariciresinol diglycoside and its mammalian lignan metabolites, enterodiol and enterolactone, inhibited the linoleic acid peroxidation at a concentration of 10 µM [345]. Further, enterolactone exhibited a significant cytotoxic effect in acute myeloid leukemia cells by enhancing DNA fragmentation and the intrinsic apoptotic pathway [346]. Another study demonstrated that dihydroguayaretic acid, guayacasin and isopregomisin are powerful antioxidants with activities similar to that of propyl gallate [347]. The antioxidant activity of schisandrene was evaluated by 2′,7′-dichlorodihydrofluorescein diacetate (DCFH-DA) cellular-based assay. A study on the structure-activity relationship suggested that the presence of exocyclic methylene functionality was responsible for antioxidant activity [348]. The flaxseed lignin, secoisolariciresinol diglucoside, enterolactone, and enterodiol exert antioxidant activity against DNA damage and lipid peroxidation and, therefore, were found to be beneficial in cancer, hypercholesterolemia, hyperglycemia, atherosclerosis, and lupus nephritis. In preclinical anticancer models, lignans reduced growth, progression, and angiogenesis [349]. Further, neolignan, isolariciresinol, and isolariciresinol isolated from German Riesling wine were reported to show antioxidant activity [350]. The sesame lignans, sesamol, sesamin and sesamolin showed lower antioxidant activity than tocols (α- and γ-tocopherols and α-tocotrienol) and butylated hydroxytoluene (BHT) in rat liver microsomes and cumene hydroperoxide (CumOOH)/Fe2+-ADP-NADPH (enzymatic) system [351].

5.6. Coumarins

Umbelliferone shows antioxidant activity in both in vitro and in vivo models. In the DPPH radical scavenging assay, it exhibited 59.6% inhibition in comparison to the standard drug which showed 96% inhibition [352]. In a site-specific deoxyribose degradation assay, it was reported that umbelliferone inhibited membrane reactive free hydroxyl radical by 63.6% [353]. Further, in ABTS scavenging and ferric reducing assay it had significant reducing capacity and had higher antiradical power than ascorbic acid [354]. Later, in irradiated lymphocytes, this compound not only reduced intracellular reactive oxygen species levels but also restored mitochondrial membrane potential and prevented DNA damage [355]. Umbelliferone also shows strong antinociceptive and anti-inflammatory activities [356]. In an in vitro study model, umbelliferone present in the ethanol extract of banana flower inhibited α-glucosidase, the polyol pathway, and protein glycation, and activated the peroxisome proliferator-activated receptors (PPARγ and PPARβ) thereby exhibiting anti-hyperglycemic activity [357,358]. Further, in streptozotocin-induced diabetic rats, umbelliferone at a dose of 30 mg/kg body weight showed significant glycemic control. Additionally, the antioxidant activity of umbelliferone prevented the liver cells from oxidative damage in streptozotocin-induced diabetic rats [359]. The coumarin inhibited reactive oxygen species generation, induced apoptosis, initiated cell cycle arrest and DNA fragmentation thereby exhibiting antitumor activity against liver hepatocellular cell lines [360]).
Aesculetin is reported to show radical scavenging activity in a dose-dependent manner [361]. Another study demonstrated that it exhibited the strongest scavenging activity among the ten isolated coumarins from A. dahurica [362]. It improved the cell viability of hydrogen peroxide-treated Caco-2 cells by enhancing mRNA expression of Nrf2 and increasing the activity of glutathione peroxidase [363].
Coumestrol exerts antioxidant activity against hydrogen peroxide-induced oxidative stress in HepG2 cells. It also prevents lipid peroxidation, ROS production and reduces SOD activity [364]. It shows an anti-inflammatory effect in LPS-activated microglia by inhibiting the production of nitric oxide, inducible nitric oxide, monocyte chemoattractant protein-1 (MCP-1), and IL-6 [365]. Yuk and co-workers studied the anti-inflammatory effects of coumestrol in lipopolysaccharide (LPS)-induced RAW264.7 macrophages and in acute lung injured mice model. Treatment with the compound reduced the production of nitric oxide (NO), tumor necrosis factor-α (TNF-α) and interleukin-6 (IL-6) in the macrophages; it further suppressed activated nuclear factor-kappa β. Furthermore, in bronchoalveolar lavage fluid of acute lung-injured mice, it reduced the level of ROS, TNF-α, MCP-1, and IL-6, inhibited NO release and suppressed the expression of NF-κB and SOD3 [366]. Moreover, its antioxidant and anti-apoptotic properties can be beneficial in treating diabetes and neurological disorders [367,368].
Bergapten, the furanocoumarin displayed moderate antioxidant activity than α-tocopherol [369,370]. In a study, the precognitive effects of bergapten attenuated the oxidative stress markers in scopolamine-induced memory impairment. It also restored the acetylcholine levels to normalcy [371]. It further provides neuroprotection in chronic constriction damage by inhibiting the overexpression of COX-2, TNF-α, and NF-ĸB [372]. In the animal model of acetic acid-induced colitis, bergapten reduced inflammation and inflammatory cell infiltration [373].
Psoralen is used in the treatment of skin diseases as it inhibits the oxidation of unsaturated lipids and prevents the impairment of barrier functions of biomembranes [374]. It even exerts strong anti-osteoporotic effects via the regulation of multiple molecular pathways such as the wnt/β-catenin, apoptosis signaling kinase 1 (ASK1)/c-jun N-terminal kinase (JNK) and the protein kinase B (AKT), and the expression of miR-488, peroxisome proliferators-activated receptor-gamma (PPARγ), and matrix metalloproteinases (MMPs) [375]. In human hepatoma cell line SMMC7721, it activates the ER signaling pathway, thereby inducing apoptosis [376]. Another study suggested that the presence of psoralen and isopsoralen in the extract of P. corylifolia L. induces apoptosis in four cancer cell lines (KB, KBv200, K562 and K562/ADM), contributing to its anticancer activity [377].
Eriodyctiol is a flavonoid with a wide range of biological activity. Specifically, it exhibits its antioxidant potency in various ways [378]. It modulates ROS in human keratinocyte cells [379], Nrf2, and downstream shielding phase-II enzyme activation [380], affects vanilloid receptors [379], and nitric oxide concentration, attenuates lipid peroxidation, and regulates the expression of Nrf2/HO-1, and β-glutamyl cysteine synthase pathways, among others [381]. Eriodyctiol and its prenylated derivative sigmoidin have been reported to demonstrate a comparable radical scavenging activity though the latter exhibited a higher cytotoxicity against cancer cells [382]. Through Nrf2/ARE signaling pathway activation, Eriodyctiol isolated from Dracocephalum rupestre PC12 cell death in a hydrogen peroxide-induced neurotoxicity [383]. By inhibiting the production of ROS, TNF-α, IL-1β, creatinine, MDA, blood urea nitrogen, TBARS among other regulating activities, Eriodyctiol has shown protective effects on cisplatin-induced kidney injury [384].
DPPH, FRAP, and ABTS antioxidant screening methods have been used to evaluate the scavenging potency of the luteolin complex, vanadium (IV) oxide sulfate monohydrate (VOSO4.H2O). The antioxidant of this flavonoid was found to increase after the formation of the complex [385]. The phospholipid complex of this flavonoid was also revealed to have a significant scavenging effect against DPPH radicals [386]. Luteolin causes apoptotic cell death in human activity by activating the HT-29 cells pathway which is mediated by mitochondria [387].
Morin protects against altered sperm parameters and testicular oxidative stress induced by bicalutamide (BCT) [388]. Morin hydrate has earlier been reported as an effective hepatic protector in both in vivo and in vitro studies [389]. It has been reported as a potential antioxidant in attenuating free radical-mediated damage to cardiovascular cells [390]. Complex formation enhances the scavenging effect of morin [391,392].
Genistein, a soya bean isoflavone, regulates the expression of antioxidant genes, modulates longevity-associated gene expression, and reduces peroxides through increased levels of MnSOD mRNA expression [393]. It has been reported to protect the kidney against IRI by attenuating both oxidative stress and inflammation [394]. As regards L-NAME-induced cardiac remodeling and dysfunction in rats, genistein has been found to be cardioprotective [395]. Genistein and daidzein have been revealed to have neuroprotective effects, enhancing choline metabolism and mitigating the PC12 cell damage induced by chlorpyrifos [396]. This study observed a better antioxidant effect for a combined therapy. Cong and colleagues reported the amelioration effect of genistein in cognitive deficits induced by chronic sleep deprivation, Nrf2 alongside downstream targets in the cortex and hippocampus of CSD-treated mice was activated. The isoflavone inhibited NF-κB, iNOS, and COX-2 activation alongside cytokines such as TNF-α, IL-6, and IL-1β [397]. Genistein improves kidney damage induced by morphine [398]. Empirical report have suggested a similar antioxidant potency for genistein and daidzein toward DNA oxidative insult [399]. As a supplement, daidzein improves embryo growth and development in early pregnancy, attributed to the isoflavone’s ability to improve the immune and antioxidant status of amniotic fluid [400]. Through the alteration of MAPK’s pathway and regulation of the inflammatory pathway in cisplatin-induced nephrotoxicity, daidzein suppresses oxidative stress and apoptosis [401]. It has also demonstrated an improved growth performance alongside antioxidant properties in weaned and growing pigs [402]. Nevertheless, daidzein has been reported to exhibit a prooxidant effect rather than being antioxidative, specifically in the brains of rats, as it decreases glutathione concentration, which weakens the body’s antioxidative defense system. It has also been suggested that the C4 keto moiety and C2 and C3 double bond in flavonoids may not be unconnected with this activity [403].
Rutin has been described as having an antioxidant activity in DPPH. It is also an inhibitor of lipid peroxidation [404]. As regards bioavailability, it has been found to elevate plasma flavonoid levels significantly [405]. In streptozotocin-induced diabetic rats, it has shown antihyperglycemic and antioxidant potency [406]. Alongside caffeic acid, it has been reported to exhibit antiaging and antioxidant potentials [407,408,409,410].

6. Polyphenols in Disease Management

Polyphenols, ubiquitous in plant-derived dietary sources, exhibit profound implications in the realm of cardiovascular health, obesity, type 2 diabetes mellitus (T2DM), inflammation, cancer, and neurodegenerative diseases. Through their potent antioxidant properties, polyphenols effectively mitigate oxidative stress, thus attenuating the progression of atherosclerosis and reducing the risk of cardiovascular events. Moreover, polyphenols demonstrate promising anti-obesity effects by modulating adipocyte metabolism, promoting adipogenesis, and regulating lipid homeostasis, thereby presenting a potential therapeutic avenue in combating obesity and its associated cardiovascular complications. In the context of T2DM, polyphenols exhibit multifaceted actions, including improving insulin sensitivity, enhancing pancreatic β-cell function, and ameliorating glucose uptake, collectively contributing to glycemic control and reducing the risk of diabetic complications. Furthermore, polyphenols exert anti-inflammatory effects by suppressing pro-inflammatory cytokines and signaling pathways, thus attenuating chronic low-grade inflammation implicated in the pathogenesis of cardiovascular disease, obesity, T2DM, and various other chronic diseases, as shown in Figure 5 [411,412,413,414,415,416,417,418,419,420,421,422,423,424,425,426,427]. Additionally, polyphenols demonstrate chemo-preventive properties by inhibiting carcinogenesis, modulating cell cycle progression, inducing apoptosis, and suppressing tumor angiogenesis, highlighting their potential in cancer prevention and adjuvant therapy. Moreover, polyphenols hold promise in neuroprotection by mitigating neuroinflammation, preserving neuronal function, and inhibiting protein misfolding and aggregation, thereby potentially delaying the onset and progression of neurodegenerative diseases such as Alzheimer’s and Parkinson’s diseases [252,264]. Collectively, the multifaceted bioactivity of polyphenols underscores their therapeutic potential in mitigating the burden of chronic diseases across diverse pathological contexts [411,412,413,414,415,416,417,418,419,420,421,422,423,424,425,426,427].

6.1. Diabetes

Polyphenols exert their effects on diabetes through intricate molecular pathways, influencing insulin sensitivity, glucose metabolism, and related complications. Flavonoids, such as quercetin, have been implicated in enhancing insulin signaling pathways, potentially improving cellular responsiveness to insulin through the PI3K/AKT pathway [411]. Additionally, these compounds may modulate carbohydrate digestion enzymes like α-amylase and α-glucosidase, impacting postprandial glucose levels via the AMPK pathway [412].
The anti-inflammatory actions of polyphenols involve the NF-κB pathway, potentially reducing chronic inflammation associated with insulin resistance. Their antioxidant properties, regulated by pathways like Nrf2, counteract oxidative stress, preserving pancreatic beta-cell function crucial for insulin secretion as shown in Figure 6 [413].
Polyphenols may also impact glucose homeostasis through interactions with the gut microbiota, influencing the gut-brain axis via the serotonin signaling pathway. Furthermore, in addressing diabetes complications, polyphenols could mitigate oxidative stress-related damage through the MAPK pathway, providing protective effects against diabetic nephropathy and retinopathy [413].
However, individual responses to polyphenols may vary due to factors like genetic differences and metabolism, necessitating personalized approaches. Ongoing research aims to elucidate the specific interactions between polyphenols and these pathways, offering potential insights for optimizing their integration into diabetes management strategies.

6.2. Respiratory Health

Polyphenols, prevalent in plant-derived foods, have emerged as potential contributors to respiratory health by targeting key pathways associated with inflammation and oxidative stress, as shown in Figure 7. In particular, flavonoids and phenolic acids, major polyphenols, exhibit anti-inflammatory effects by inhibiting the NF-κB pathway, suggesting a potential role in mitigating airway inflammation in conditions like asthma and COPD [414]. Additionally, their antioxidant properties, mediated through the Nrf2 pathway, could counteract oxidative stress, offering protection against respiratory damage. Certain polyphenols, such as quercetin, may influence bronchodilation by modulating calcium channels and phosphodiesterase, providing a novel approach to address airway constriction. Moreover, polyphenols may modulate immune responses and impact mucin production, suggesting potential immunomodulatory and mucolytic effects in respiratory disorders. While ongoing research seeks to elucidate the precise mechanisms, integrating polyphenol-rich foods into the diet holds promise for supporting respiratory well-being, with personalized approaches to optimize their effectiveness [415].

6.3. Pregnancy and Maternal Health

Polyphenolic compounds, widely distributed in a variety of plant-derived sources, have garnered significant interest due to their potential influence on maternal health and pregnancy outcomes. The intricate molecular structure of polyphenols, encompassing flavonoids, phenolic acids, and extracts rich in polyphenolic content, is currently under scrutiny for its antioxidative properties, which are crucial in alleviating oxidative stress-an established contributor to complications during pregnancy. The modulation of cellular signaling pathways, such as the nuclear factor erythroid 2-related factor 2 (Nrf2) pathway [416], is central to understanding the intricate interactions through which polyphenols exert their protective effects. Moreover, these bioactive compounds exhibit anti-inflammatory properties by inhibiting pro-inflammatory mediators and activating pathways like nuclear factor-kappa B (NF-κB), potentially influencing the inflammatory milieu during pregnancy. Additionally, polyphenols may impact vascular function by promoting endothelial health, involving the regulation of nitric oxide (NO) production and endothelial nitric oxide synthase (eNOS) activity [417]. Hormonal balance, particularly in estrogen metabolism, emerges as another facet through which polyphenols may contribute to maternal well-being. Despite encouraging findings, there remains a need for comprehensive investigations into optimal dosages, bioavailability, and specific subclasses of polyphenols pertinent to pregnancy, offering a more nuanced understanding of their therapeutic potential during this critical life stage [418].

6.4. Polyphenols and Microbiome-Brain Axis

In the intricate relationship known as the Microbiome-Brain Axis, polyphenols, a diverse group of bioactive compounds prevalent in various plant-based foods, have emerged as pivotal modulators. Recent research highlights the ability of polyphenols to influence microbial composition and diversity in the gut, leading to the production of bioactive metabolites, such as short-chain fatty acids (SCFAs). These metabolites, derived from microbial fermentation of polyphenols, can traverse the blood-brain barrier and impact neuronal function, potentially exerting neuroprotective effects. Polyphenols have been shown to interact with key signaling pathways, including the gut-brain axis’s intricate communication networks, involving the vagus nerve and the enteric nervous system [419]. Moreover, polyphenols demonstrate anti-inflammatory and antioxidant properties, influencing gut microbiota homeostasis and subsequently contributing to neural health. As the understanding of the Polyphenols and Microbiome-Brain Axis deepens, elucidating the specific mechanisms and pathways involved in this complex crosstalk holds significant promise for developing targeted interventions to promote brain health and potentially mitigate neurological disorders [420,421].

6.5. Polyphenols and Bone Health

Polyphenols, a diverse class of naturally occurring compounds found in various plant-based foods, have emerged as potential contributors to bone health, presenting a novel avenue of exploration in the field of skeletal physiology. Recent research suggests that polyphenols may influence bone metabolism through intricate mechanisms involving modulation of osteoblast and osteoclast activity. Flavonoids, a subgroup of polyphenols, have been shown to stimulate osteoblastic differentiation and mineralization, key processes in bone formation, potentially mediated through pathways such as Wnt/β-catenin signaling. Additionally, polyphenols exhibit antioxidant and anti-inflammatory properties, mitigating oxidative stress and inflammation, factors implicated in bone resorption [422]. Polyphenol-induced activation of nuclear factor erythroid 2-related factor 2 (Nrf2) pathway may also contribute to cellular defense against oxidative damage in bone cells. Furthermore, the potential impact of polyphenols on gut microbiota modulation, with implications for calcium absorption and bone mineral density, adds another layer of complexity to their role in bone health. While promising, further elucidation of the specific polyphenolic subclasses, optimal dosage, and underlying molecular pathways involved is essential for advancing our understanding and leveraging polyphenols as potential adjuncts in promoting skeletal well-being [423].

6.6. Polyphenols and Autoimmune Disorders

Within the realm of autoimmune disorders, the intricate interplay between polyphenols, a diverse group of bioactive compounds abundant in various plant-derived foods, has garnered increasing attention, offering a novel avenue for therapeutic exploration. Emerging research indicates that polyphenols may exert immunomodulatory effects through intricate mechanisms involving the regulation of immune cell function. Flavonoids, a prominent subclass of polyphenols, have been implicated in the suppression of pro-inflammatory cytokines, such as tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6), potentially mediated through the inhibition of nuclear factor-kappa B (NF-κB) signaling pathways. Moreover, polyphenols exhibit antioxidant properties, mitigating oxidative stress, a known contributor to autoimmune pathogenesis [424]. Polyphenol-induced modulation of T-helper cell balance, regulatory T-cell activity, and the suppression of autoantibody production underscore their multifaceted impact on immune regulation. The potential interplay between polyphenols and gut microbiota further adds complexity, as alterations in microbial composition may influence immune responses in autoimmune disorders. While promising, comprehensive investigations into the specific polyphenolic subclasses, dosage, and molecular pathways involved are imperative for advancing our understanding and harnessing the therapeutic potential of polyphenols in the context of autoimmune disorders [425].

6.7. Polyphenols and Metabolic Syndrome

Polyphenols, a diverse group of bioactive compounds ubiquitous in various plant-derived sources, have garnered considerable interest for their potential to mitigate metabolic syndrome, representing a novel avenue in the realm of metabolic health. Recent research suggests that polyphenols may exert beneficial effects through intricate pathways involved in metabolic regulation. Flavonoids, a subclass of polyphenols, have demonstrated potential in ameliorating insulin resistance, a central feature of metabolic syndrome, by modulating signaling pathways such as the phosphoinositide 3-kinase/protein kinase B (PI3K/Akt) pathway. Polyphenols also exhibit anti-inflammatory properties, targeting key mediators like nuclear factor-kappa B (NF-κB) and interleukin-6 (IL-6) [426], thereby attenuating chronic low-grade inflammation, often associated with metabolic syndrome. Furthermore, polyphenols may influence lipid metabolism, regulating pathways such as peroxisome proliferator-activated receptor gamma (PPARγ) and adenosine monophosphate-activated protein kinase (AMPK), contributing to improved lipid profiles and adipose tissue function. The potential impact of polyphenols on gut microbiota composition, affecting metabolic processes and inflammation, adds an additional layer of complexity to their role in metabolic health. While promising, comprehensive investigations into optimal dosages, specific polyphenolic subclasses, and underlying molecular mechanisms are essential for advancing our understanding and harnessing the therapeutic potential of polyphenols [427,428,429,430,431,432,433,434,435,436].

6.8. Polyphenols in Neurodegenerative Diseases

In the realm of neurological health, the spotlight increasingly falls on polyphenols, natural compounds abundant in a variety of dietary sources, including fruits, vegetables, and beverages like tea and wine. Due to their multifaceted properties, these bioactive molecules have emerged as promising candidates for mitigating the onset and progression of neurodegenerative diseases. With antioxidant, anti-inflammatory, and neuroprotective effects, polyphenols hold the potential to preserve neuronal function and combat the underlying mechanisms of conditions such as Alzheimer’s and Parkinson’s diseases [181,198]. Mechanistically, polyphenols exert their neuroprotective effects through various molecular pathways. They modulate oxidative stress by scavenging free radicals and upregulating antioxidant enzymes such as superoxide dismutase and catalase. Additionally, polyphenols inhibit the activation of inflammatory pathways mediated by nuclear factor-kappa B (NF-κB) and mitogen-activated protein kinases (MAPKs), thereby reducing neuroinflammation. Moreover, polyphenols interact with key proteins involved in neurodegenerative processes, such as beta-amyloid and alpha-synuclein, preventing their aggregation and fibril formation. Furthermore, polyphenols may enhance neuroplasticity and synaptic function by activating signaling pathways such as the brain-derived neurotrophic factor (BDNF) pathway. These molecular insights underscore the significance of polyphenols in the quest for effective treatments against these debilitating disorders [252,264].

7. Polyphenols and Epigenetics

Polyphenols, a diverse group of naturally occurring compounds in plant-derived foods, have garnered substantial attention for their potential health-promoting effects [437,438,439,440,441]. A burgeoning area of research involves the intricate interplay between polyphenols and epigenetic mechanisms, which regulate gene expression without altering the underlying DNA sequence [442,443,444,445]. The multifaceted structures of polyphenols, such as flavonoids, phenolic acids, lignans, and stilbenes, enable them to interact with enzymes involved in epigenetic modifications, such as DNA methyltransferases and histone acetyltransferases [446].
The epigenetic modulation by polyphenols encompasses several key aspects. Polyphenols may influence DNA methylation, a process crucial for gene silencing or activation. Studies have suggested that certain polyphenols possess the capacity to inhibit DNA methyltransferases, thereby affecting the methylation status of specific genes associated with various physiological processes [447].
Secondly, polyphenols can impact histone modifications, playing a pivotal role in chromatin structure and gene accessibility. Polyphenols may serve as histone deacetylase (HDAC) inhibitors, influencing the acetylation status of histones and consequently regulating gene expression. This modulation can have implications for diverse cellular functions, including cell cycle control, apoptosis, and inflammation [448].
Furthermore, polyphenols may interact with non-coding RNAs, such as microRNAs (miRNAs), which play a crucial role in post-transcriptional gene regulation. By affecting the expression levels of specific miRNAs, polyphenols can indirectly influence the expression of their target genes and, consequently, various cellular processes [449].
The impact of polyphenols on epigenetic processes holds promise for therapeutic applications. Understanding the epigenetic mechanisms influenced by polyphenols may offer insights into their potential roles in preventing or managing diseases with an underlying epigenetic component, such as cancer, cardiovascular diseases, and neurodegenerative disorders [450].
Despite the promising avenues of research, it is important to acknowledge the complexity of these interactions. The bioavailability, metabolism, and specific molecular targets of polyphenols can vary, influencing their efficacy in epigenetic modulation. Thus, further elucidation of the specific polyphenolic compounds, dosage, and duration required for optimal epigenetic effects is imperative for harnessing their full therapeutic potential. Exploring the dynamic interplay between polyphenols and epigenetics remains a captivating frontier in nutritional and medical sciences, holding promise for personalized interventions and developing novel strategies for health promotion and disease prevention [451].

8. Polyphenols, Food and Commercial Importance

8.1. Polyphenols in Food Processing

The effects of food processing techniques on polyphenols, vital bioactive compounds in plant-derived foods, are pivotal considerations in shaping the content, structure, and bioavailability of these compounds and, consequently, their potential health benefits. Heat treatment, including cooking and pasteurization, can lead to the thermal breakdown of certain polyphenols, but moderate heat can enhance extractability. Extraction methods, such as solvent extraction or supercritical fluid extraction, profoundly influence the types and concentrations of polyphenols in the final product [428]. Fermentation processes generate both metabolites with potential health benefits and may cause the breakdown of specific polyphenolic structures. Storage conditions, exposing foods to light, air, and elevated temperatures, can contribute to polyphenol degradation. Moreover, these processing techniques influence the bioavailability of polyphenols, affecting their absorption, distribution, metabolism, and excretion within the human body [429].
While certain methods may result in polyphenol loss, others enhance bioavailability or generate bioactive metabolites. A comprehensive understanding of the intricate interactions between food processing and polyphenols is crucial for optimizing the health benefits of polyphenol-rich foods and informing dietary recommendations for promoting processed foods that retain or enhance their polyphenolic content [430].

8.2. Utilizing Polyphenols from Food Waste

Deriving polyphenols from waste food stands as a pioneering strategy, capitalizing on resource efficiency and environmental conservation as shown in Figure 8. Food waste, replete with bioactive compounds, emerges as a reservoir rich in potentially valuable polyphenols [431,432,433]. Examples of extraction avenues from diverse food waste streams include:
  • Fruit and vegetable peelings: The outer layers of commonly discarded fruits and vegetables house a substantial concentration of polyphenols. Peelings, especially those sourced from apples, citrus fruits, carrots, and potatoes, can be extracted through diverse extraction methodologies [434,435];
  • Wine residues: Post-wine production, pomace comprising grape skins, seeds, and stems retains polyphenols. Extracting these compounds from wine residue diminishes waste and yields polyphenolic extracts with applications across various industries [436,437];
  • Coffee grounds: Residual polyphenols persist in used coffee grounds after coffee brewing. Extracting polyphenols from spent coffee grounds presents a sustainable avenue for waste utilization [438,439];
  • Tea residues: Spent tea leaves or residue post-brewing represent an additional source of polyphenols. Employing suitable extraction techniques enables the recovery of polyphenols from these discarded tea remnants [440,441];
  • Waste from fruit and vegetable processing: By-products generated during the processing of fruits and vegetables, encompassing peels, seeds, and cores, often become discarded waste. These by-products can be investigated as potential sources of polyphenols through adept extraction processes [442];
  • Brewing and distillation of by-products: Residues ensuing from brewing beer or distilling spirits, such as spent grains, may contain polyphenols. Employing innovative extraction methods facilitates the recovery of polyphenols from these by-products [443,444];
Figure 8. Extracting polyphenols from food waste represents an innovative approach.
Figure 8. Extracting polyphenols from food waste represents an innovative approach.
Antioxidants 13 00429 g008
This polyphenol extraction paradigm from food waste is congruent with the principles of a circular economy, emphasizing waste minimization and maximal resource utilization. It is imperative to ensure that extraction methods align with sustainability principles and that resultant polyphenolic extracts adhere to stringent safety and quality standards for prospective applications in industries such as food, pharmaceuticals, or cosmetics. Concurrently, focused endeavors should be directed toward the development of efficient and eco-friendly extraction processes to scale this approach feasibly.

9. Conclusions

Consumption of plant-based products, depending on their phyto-antioxidant content, reduces the risk of development and progression of many chronic diseases and improves human health. In conclusion, our comprehensive review has elucidated the substantial health advantages associated with the consumption of plant-based products enriched in phyto-antioxidants. This dietary modality has demonstrated a discernible reduction in the susceptibility to and progression of diverse chronic diseases, thereby substantiating its potential as a cornerstone in augmenting human health.
Although the predominant paradigm attributes the salutary effects to the antioxidant activity inherent in polyphenols, it is imperative to acknowledge the intricate and incompletely understood mechanisms of action, particularly concerning specific phenolic compounds. Mitigating this lacuna in knowledge necessitates a nuanced exploration of polyphenol bioavailability, with the anticipation that such insights will conduce to a more sophisticated comprehension of their health-propagating attributes.
Expanding the purview of our investigation, we have probed the multifaceted roles of polyphenols in disease management. Specifically, we have scrutinized their potential impact on conditions such as diabetes, respiratory health, pregnancy, maternal health, autoimmune disorders, and metabolic syndrome. This nuanced examination of polyphenols in diverse pathophysiological contexts enhances our grasp of their versatile applications and underscores their potential as therapeutic modalities in varied health paradigms.
Furthermore, our inquiry extends to the realm of polyphenols in food processing, wherein we have scrutinized their influence on the nutritional profile and bioactivity of processed foodstuffs. A foray into the burgeoning field of polyphenols in epigenetics has augmented our understanding of their putative influence on gene expression and cellular homeostasis.
Recognizing the imperatives of sustainability, we have addressed the deployment of polyphenols derived from residual food sources. This facet not only mitigates food wastage but also unravels novel avenues for harnessing the health benefits of polyphenols from unconventional reservoirs.
Notwithstanding the strides made in polyphenol research, challenges persist, notably the impact of structural variations on optimal bioavailability, metabolic kinetics, and modes of biological activity. Accordingly, ongoing investigations are dedicated to structural modifications of polyphenols, with the overarching aim of enhancing their bioavailability and therapeutic efficacy. This strategic pursuit portends to unlock the full therapeutic potential of polyphenols, offering precision interventions for disease prevention and management, thereby cementing the pivotal role of plant-based products in fortifying holistic human well-being. The comprehensive exploration of these diverse facets underscores the exigency for sustained research and accentuates the potential of polyphenols as invaluable contributors to human health across a spectrum of conditions.
The hypothesis underlying the importance of the comprehensive review posits that a thorough exploration and synthesis of existing literature on polyphenols in plant-based products enriched with phyto-antioxidants is pivotal for advancing our understanding of their profound impact on human health. This hypothesis is rooted in the premise that by meticulously examining the diverse roles and mechanisms of action of polyphenols across various chronic diseases and physiological processes, the review serves as a cornerstone for elucidating the intricate interplay between dietary factors and health outcomes. Furthermore, the hypothesis suggests that by critically analyzing the current state of knowledge, identifying gaps, and proposing future research directions, the comprehensive review not only contributes to scientific knowledge but also informs clinical practice, dietary guidelines, and public health initiatives aimed at promoting optimal health and preventing chronic diseases. Ultimately, the hypothesis underscores the indispensable role of comprehensive reviews in synthesizing evidence, fostering scientific discourse, and driving advancements in research and healthcare practices.

Author Contributions

Conceptualization, M.R.; methodology, M.R.; formal analysis, S.G.; investigation, G.R. and R.P.S.; resources, M.R.; data curation, S.G., G.R. and R.P.S.; writing—original draft preparation, M.R. and S.G.; writing—review and editing, G.R., R.P.S. and S.C.; visualization, J.K.; supervision, M.R.; project administration, M.R.; funding acquisition, J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data is included in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sharifi-Rad, M.; Anil Kumar, N.V.; Zucca, P.; Varoni, E.M.; Dini, L.; Panzarini, E.; Rajkovic, J.; Tsouh Fokou, P.V.; 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]
  2. Ghazizadeh, H.; Saberi-Karimian, M.; Aghasizadeh, M.; Sahebi, R.; Ghazavi, H.; Khedmatgozar, H.; Timar, A.; Rohban, M.; Javandoost, A.; Ghayour-Mobarhan, M. Pro-oxidant–antioxidant balance (PAB) as a prognostic index in assessing the cardiovascular risk factors: A narrative review. Obes. Med. 2020, 19, 100272. [Google Scholar] [CrossRef]
  3. Pandey, K.B.; Rizvi, S.I. Plant polyphenols as dietary antioxidants in human health and disease. Oxid. Med. Cell. Longev. 2009, 2, 270–278. [Google Scholar] [CrossRef] [PubMed]
  4. Jakubczyk, K.; Dec, K.; Kałduńska, J.; Kawczuga, D.; Kochman, J.; Janda, K. Reactive oxygen species-sources, functions, oxidative damage. Pol. Merkur. Lek. 2020, 48, 124–127. [Google Scholar]
  5. Antonucci, S.; Di Sante, M.; Tonolo, F.; Pontarollo, L.; Scalcon, V.; Alanova, P.; Menabò, R.; Carpi, A.; Bindoli, A.; Rigobello, M.P.; et al. The determining role of mitochondrial reactive oxygen species generation and monoamine oxidase activity in doxorubicin-induced cardiotoxicity. Antioxid. Redox Signal. 2021, 34, 531–550. [Google Scholar] [CrossRef] [PubMed]
  6. Bergendi, L.; Beneš, L.; Ďuračková, Z.; Ferenčik, M. Chemistry, physiology and pathology of free radicals. Life Sci. 1999, 65, 1865–1874. [Google Scholar] [CrossRef]
  7. Mishra, T.; Khullar, M.; Bhatia, A. Anticancer potential of aqueous ethanol seed extract of Ziziphus mauritiana against cancer cell lines and Ehrlich ascites carcinoma. Evid.-Based Complement. Altern. Med. 2011, 2011, 765029. [Google Scholar] [CrossRef] [PubMed]
  8. Attia, M.S.; Sharaf, M.H.; Hashem, A.H.; Mahfouz, A.Y.; Daigham, G.E.; Al-Askar, A.A.; Abdelgawad, H.; Omar, M.S.; Thabet, A.E.; Abdalmohsen, M.M.; et al. Application of Rhizopus microsporus and Aspergillus oryzae to enhance the defense capacity of eggplant seedlings against Meloidogyne incognita. Not. Bot. Horti Agrobot. Cluj-Napoca 2023, 51, 13300. [Google Scholar] [CrossRef]
  9. Varijakzhan, D.; Chong, C.-M.; Abushelaibi, A.; Lai, K.-S.; Lim, S.-H.E. Middle Eastern plant extracts: An alternative to modern medicine problems. Molecules 2020, 25, 1126. [Google Scholar] [CrossRef]
  10. Scalbert, A.; Johnson, I.T.; Saltmarsh, M. Polyphenols: Antioxidants and beyond. Am. J. Clin. Nutr. 2005, 81, 215S–217S. [Google Scholar] [CrossRef]
  11. Daglia, M. Polyphenols as antimicrobial agents. Curr. Opin. Biotechnol. 2012, 23, 174–181. [Google Scholar] [CrossRef] [PubMed]
  12. Han, X.; Shen, T.; Lou, H. Dietary polyphenols and their biological significance. Int. J. Mol. Sci. 2007, 8, 950–988. [Google Scholar] [CrossRef]
  13. Tsao, R. Chemistry and Biochemistry of Dietary Polyphenols. Nutrients 2010, 2, 1231–1246. [Google Scholar] [CrossRef] [PubMed]
  14. Green, P.S.; Gordon, K.; Simpkins, J.W. Phenolic A ring requirement for the neuroprotective effects of steroids. J. Steroid Biochem. Mol. Biol. 1997, 63, 229–235. [Google Scholar] [CrossRef] [PubMed]
  15. Abbas, M.; Saeed, F.; Anjum, F.M.; Afzaal, M.; Tufail, T.; Bashir, M.S.; Ishtiaq, A.; Hussain, S.; Suleria, H.A.R. Natural polyphenols: An overview. Int. J. Food Prop. 2017, 20, 1689–1699. [Google Scholar] [CrossRef]
  16. Kim, K.-H.; Tsao, R.; Yang, R.; Cui, S.W. Phenolic acid profiles and antioxidant activities of wheat bran extracts and the effect of hydrolysis conditions. Food Chem. 2006, 95, 466–473. [Google Scholar] [CrossRef]
  17. Rudrapal, M.; Khairnar, S.J.; Khan, J.; Dukhyil, A.B.; Ansari, M.A.; Palai, S.; Devi, R. Dietary polyphenols and their role in oxidative stress-induced human diseases: Insights into protective effects, antioxidant potentials and mechanism (s) of action. Front. Pharmacol. 2022, 13, 806470. [Google Scholar] [CrossRef] [PubMed]
  18. Ahmed, A.; Tariq, A.; Habib, S. Interactive biology of auxins and phenolics in plant environment. Plant Phenolics Sustain. Agric. 2020, 1, 117–133. [Google Scholar]
  19. Lin, W.; Li, Y.; Lu, Q.; Lu, H.; Li, J. Combined analysis of the metabolome and transcriptome identified candidate genes involved in phenolic acid biosynthesis in the leaves of Cyclocarya paliurus. Int. J. Mol. Sci. 2020, 21, 1337. [Google Scholar] [CrossRef]
  20. Marchiosi, R.; Dos Santos, W.D.; Constantin, R.P.; De Lima, R.B.; Soares, A.R.; Finger-Teixeira, A.; Mota, T.R.; de Oliveira, D.M.; Foletto-Felipe, M.D.P.; Abrahão, J.; et al. Biosynthesis and metabolic actions of simple phenolic acids in plants. Phytochem. Rev. 2020, 19, 865–906. [Google Scholar] [CrossRef]
  21. Spigoni, V.; Mena, P.; Fantuzzi, F.; Tassotti, M.; Brighenti, F.; Bonadonna, R.C.; Del Rio, D.; Cas, A.D. Bioavailability of bergamot (Citrus bergamia) flavanones and biological activity of their circulating metabolites in human pro-angiogenic cells. Nutrients 2017, 9, 1328. [Google Scholar] [CrossRef] [PubMed]
  22. Tzanova, M.; Atanasov, V.; Yaneva, Z.; Ivanova, D.; Dinev, T. Selectivity of current extraction techniques for flavonoids from plant materials. Processes 2020, 8, 1222. [Google Scholar] [CrossRef]
  23. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [PubMed]
  24. Ahmad, R.; Fatma, T. Flavonoids as Food Supplement with Special Reference to Algae and Cyanobacteria. In Algae and Sustainable Technologies; CRC Press: Boca Raton, FL, USA, 2020; pp. 231–244. [Google Scholar]
  25. Yu, X.; Meenu, M.; Xu, B.; Yu, H. Impact of processing technologies on isoflavones, phenolic acids, and antioxidant capacities of soymilk prepared from 15 soybean varieties. Food Chem. 2020, 345, 128612. [Google Scholar] [CrossRef]
  26. Anggraini, T.; Syukri, D.; Manasikan, T.; Nakano, K. Anthocyanin profile of Syzygium oleana young leaves and fruits using triple quadrupole mass spectrometer: Identification of a new peonidin. Biodiversitas 2020, 21, 5893–5900. [Google Scholar] [CrossRef]
  27. Deng, Y.; Lu, S. Biosynthesis and Regulation of Phenylpropanoids in Plants. Crit. Rev. Plant Sci. 2017, 36, 257–290. [Google Scholar] [CrossRef]
  28. King, R.E.; Bomser, J.A.; Min, D.B. Bioactivity of resveratrol. Compr. Rev. Food Sci. Food Saf. 2006, 5, 65–70. [Google Scholar] [CrossRef]
  29. Oh, W.Y.; Gao, Y.; Shahidi, F. Stilbenoids: Chemistry, occurrence, bioavailability and health effects—A review. J. Food Bioact. 2021, 13, 20–31. [Google Scholar] [CrossRef]
  30. Serrano, J.; Puupponen-Pimiä, R.; Dauer, A.; Aura, A.; Saura-Calixto, F. Tannins: Current knowledge of food sources, intake, bioavailability and biological effects. Mol. Nutr. Food Res. 2009, 53, S310–S329. [Google Scholar] [CrossRef]
  31. Samtiya, M.; Aluko, R.E.; Dhewa, T. Plant food anti-nutritional factors and their reduction strategies: An overview. Food Prod. Process. Nutr. 2020, 2, 6. [Google Scholar] [CrossRef]
  32. Adlercreutz, H.; Mazur, W. Phyto-Oestrogens and Western Diseases. Ann. Med. 1997, 29, 95–120. [Google Scholar] [CrossRef] [PubMed]
  33. Imai, T.; Nomura, M.; Fukushima, K. Evidence for involvement of the phenylpropanoid pathway in the biosynthesis of the norlignan agatharesinol. J. Plant Physiol. 2006, 163, 483–487. [Google Scholar] [CrossRef] [PubMed]
  34. Ribeiro, C.V.C.; Kaplan, M.A.C. Tendênciasevolutivas de famíliasprodutoras de cumarinasem Angiospermae. Química Nova 2002, 25, 533–538. [Google Scholar] [CrossRef]
  35. Venugopala, K.N.; Rashmi, V.; Odhav, B. Review on natural coumarin lead compounds for their pharmacological activity. BioMed Res. Int. 2013, 2013, 963248. [Google Scholar] [CrossRef] [PubMed]
  36. Dewick, P.M. Medicinal Natural Products: A Biosynthetic Approach, 3rd ed.; John Wiley & Sons: Nottingham, UK, 2002; pp. 499–500. [Google Scholar]
  37. Lin, Y.; Luo, T.; Weng, A.; Huang, X.; Yao, Y.; Fu, Z.; Li, Y.; Liu, A.; Li, X.; Chen, D.; et al. Gallic acid alleviates gouty arthritis by inhibiting NLRP3 inflammasome activation and pyroptosis through enhancing Nrf2 signaling. Front. Immunol. 2020, 11, 580593. [Google Scholar] [CrossRef] [PubMed]
  38. Boo, Y.C. p-Coumaric acid as an active ingredient in cosmetics: A review focusing on its antimelanogenic effects. Antioxidants 2019, 8, 275. [Google Scholar] [CrossRef] [PubMed]
  39. Espíndola, K.M.M.; Ferreira, R.G.; Narvaez, L.E.M.; Silva Rosario, A.C.R.; Da Silva, A.H.M.; Silva, A.G.B.; Monteiro, M.C. Chemical and pharmaco-logical aspects of caffeic acid and its activity in hepatocarcinoma. Front. Oncol. 2019, 9, 541. [Google Scholar] [CrossRef] [PubMed]
  40. Alam, M.A. Anti-hypertensive effect of cereal antioxidant ferulic acid and its mechanism of action. Front. Nutr. 2019, 6, 121. [Google Scholar] [CrossRef]
  41. Chen, C. Sinapic acid and its derivatives as medicine in oxidative stress-induced diseases and aging. J. Oxid. Med. Cell. Longev. 2016, 2016, 3571614. [Google Scholar] [CrossRef]
  42. Zouaoui, S.; Farman, M.; Semmar, N. Review on structural trends and chemotaxonomical aspects of pharmacologically evaluated flavonoids. Curr. Top. Med. Chem. 2021, 21, 628–648. [Google Scholar] [CrossRef]
  43. Gharbavi, M.; Mojaddami, A. Flavonoids in the Treatment of Glioblastoma Using Niosomal Nanocarrier. Jundishapur J. Nat. Pharm. Prod. 2023, 18, e135027. [Google Scholar] [CrossRef]
  44. Isemura, M. Catechin in human health and disease. Molecules 2019, 24, 528. [Google Scholar] [CrossRef] [PubMed]
  45. Kalt, W.; Cassidy, A.; Howard, L.R.; Krikorian, R.; Stull, A.J.; Tremblay, F.; Zamora-Ros, R. Recent research on the health benefits of blueberries and their anthocyanins. Adv. Nutr. Int. Rev. J. 2020, 11, 224–236. [Google Scholar] [CrossRef]
  46. Den Hartogh, D.J.; Tsiani, E. Health benefits of resveratrol in kidney disease: Evidence from in vitro and in vivo studies. Nutrients 2019, 11, 1624. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, S.; Wang, J.; Fu, Z.; Familiari, G.; Relucenti, M.; Aschner, M.; Li, X.; Chen, H.; Chen, R. Matairesinol nanoparticles restore chemosensitivity and suppress colorectal cancer progression in preclinical models: Role of lipid metabolism reprogramming. Nano Lett. 2023, 23, 1970–1980. [Google Scholar] [CrossRef] [PubMed]
  48. Gadgoli, C.; Bhandekar, S.; Bhere, H.; Naik, A. Bioavailability and Pharmacokinetic Parameters of a Formulation Containing Secoisolariciresinol Diglucoside–Rich Extract. Rev. Farmacogn. 2023, 33, 394–401. [Google Scholar] [CrossRef]
  49. Ma, G.; Chen, Y. Polyphenol supplementation benefits human health via gut microbiota: A systematic review via meta-analysis. J. Funct. Foods 2020, 66, 103829. [Google Scholar] [CrossRef]
  50. Brglez Mojzer, E.; Knez Hrnčič, M.; Škerget, M.; Knez, Ž.; Bren, U. Polyphenols: Extraction methods, antioxidative action, bioavailability and anticarcinogenic effects. Molecules 2016, 21, 901. [Google Scholar] [CrossRef]
  51. Haddad, M.A.; Dmour, H.; Al-Khazaleh, J.M.; Obeidat, M.; Al-Abbadi, A.; Al-Shadaideh, A.N.; Al-Mazra’Awi, M.S.; Shatnawi, M.A.; Iommi, C. Herbs and medicinal plants in Jordan. J. AOAC Int. 2020, 103, 925–929. [Google Scholar] [CrossRef]
  52. Malenčić, Đ.; Kiprovski, B.; Bursić, V.; Vuković, G.; Ćupina, B.; Mikić, A. Dietary phenolics and antioxidant capacity of selected legumes seeds from the central Balkans. Acta Aliment. 2018, 47, 340–349. [Google Scholar] [CrossRef]
  53. Les, F.; Venditti, A.; Cásedas, G.; Frezza, C.; Guiso, M.; Sciubba, F.; López, V. Everlasting flower (Helichrysum stoechas Moench) as a potential source of bioactive molecules with antiproliferative, antioxidant, antidiabetic and neuroprotective properties. Ind. Crops Prod. 2017, 108, 295–302. [Google Scholar] [CrossRef]
  54. Ali, G.; Neda, G. Flavonoids and phenolic acids: Role and biochemical activity in plants and human. J. Med. Plants Res. 2011, 5, 6697–6703. [Google Scholar]
  55. Abdelshafy, A.M.; Belwal, T.; Liang, Z.; Wang, L.; Li, D.; Luo, Z.; Li, L. A comprehensive review on phenolic compounds from edible mushrooms: Occurrence, biological activity, application and future prospective. Crit. Rev. Food Sci. Nutr. 2021, 62, 6204–6224. [Google Scholar] [CrossRef] [PubMed]
  56. Shahidi, F.; Naczk, M. Phenolics in Food and Nutraceuticals; CRC Press: Boca Raton, FL, USA, 2003. [Google Scholar]
  57. Hazafa, A.; Rehman, K.-U.; Jahan, N.; Jabeen, Z. The role of polyphenol (flavonoids) compounds in the treatment of cancer cells. Nutr. Cancer 2020, 72, 386–397. [Google Scholar] [CrossRef] [PubMed]
  58. Rivière, C.; Pawlus, A.D.; Mérillon, J.-M. Natural stilbenoids: Distribution in the plant kingdom and chemotaxonomic interest in Vitaceae. Nat. Prod. Rep. 2012, 29, 1317–1333. [Google Scholar] [CrossRef] [PubMed]
  59. Lim, C.G.; AG Koffas, M. Bioavailability and recent advances in the bioactivity of flavonoid and stilbene compounds. Curr. Org. Chem. 2010, 14, 1727–1751. [Google Scholar] [CrossRef]
  60. Fontana, A.R.; Antoniolli, A.; Bottini, R. Grape pomace as a sustainable source of bioactive compounds: Extraction, characterization, and biotechnological applications of phenolics. J. Agric. Food Chem. 2013, 61, 8987–9003. [Google Scholar] [CrossRef] [PubMed]
  61. Preisig-Müller, R.; Schwekendiek, A.; Brehm, I.; Reif, H.J.; Kindl, H. Characterization of a pine multigene family containing elicitor-responsive stilbene synthase genes. Plant Mol. Biol. 1999, 39, 221–229. [Google Scholar] [CrossRef]
  62. Bernjak, B.; Kristl, J. A review of Tannins in berries. Agricultura 2020, 17, 27–36. [Google Scholar] [CrossRef]
  63. Henning, S.M.; Fajardo-Lira, C.; Lee, H.W.; Youssefian, A.A.; Go, V.L.W.; Heber, D. Catechin content of 18 teas and a green tea extract supplement correlates with the antioxidant capacity. Nutr. Cancer 2003, 45, 226–235. [Google Scholar] [CrossRef]
  64. Yeung, A.W.K.; Tzvetkov, N.T.; Balacheva, A.A.; Georgieva, M.G.; Gan, R.-Y.; Jozwik, A.; Pyzel, B.; Horbańczuk, J.O.; Novellino, E.; Durazzo, A.; et al. Lignans: Quantitative analysis of the research literature. Front. Pharmacol. 2020, 11, 37. [Google Scholar] [CrossRef] [PubMed]
  65. Rodríguez-García, C.; Sánchez-Quesada, C.; Toledo, E.; Delgado-Rodríguez, M.; Gaforio, J.J. Naturally lignan-rich foods: A dietary tool for health promotion? Molecules 2019, 24, 917. [Google Scholar] [CrossRef] [PubMed]
  66. Cherng, J.-M.; Chiang, W.; Chiang, L.-C. Immunomodulatory activities of common vegetables and spices of Umbelliferae and its related coumarins and flavonoids. Food Chem. 2008, 106, 944–950. [Google Scholar] [CrossRef]
  67. Kowalski, R.; de Mejia, E.G. Phenolic composition, antioxidant capacity and physical characterization of ten blackcurrant (Ribes nigrum) cultivars, their juices, and the inhibition of type 2 diabetes and inflammation biochemical markers. Food Chem. 2021, 359, 129889. [Google Scholar] [CrossRef]
  68. Clifford, M.N. Chlorogenic acids and other cinnamates–nature, occurrence and dietary burden. J. Sci. Food Agric. 1999, 79, 362–372. [Google Scholar] [CrossRef]
  69. Čanadanović-Brunet, J.; Vulić, J.; Ćebović, T.; Ćetković, G.; Čanadanović, V.; Djilas, S.; Šaponjac, V.T. Phenolic profile, antiradical and anti-tumour evaluation of raspberries pomace extract from Serbia. Iran. J. Pharm. Res. 2017, 16, 142–152. [Google Scholar] [PubMed]
  70. Trugo, L.C.; Finglas, P.M. Encyclopedia of Food Sciences and Nutrition; Trugo, L.C., Finglas, P.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2003; pp. 1498–1506. [Google Scholar]
  71. Zhang, J.; Tian, J.; Gao, N.; Gong, E.S.; Xin, G.; Liu, C.; Si, X.; Sun, X.; Li, B. Assessment of the phytochemical profile and antioxidant activities of eight kiwi berry (Actinidia arguta (Siebold & Zuccarini) Miquel) varieties in China. Food Sci. Nutr. 2021, 9, 5616–5625. [Google Scholar] [CrossRef] [PubMed]
  72. Galanakis, C.M. Polyphenols: Properties, Recovery, and Applications; Woodhead Publishing: Sawston, UK, 2018. [Google Scholar]
  73. Antognoni, F.; Potente, G.; Mandrioli, R.; Angeloni, C.; Freschi, M.; Malaguti, M.; Hrelia, S.; Lugli, S.; Gennari, F.; Muzzi, E.; et al. Fruit quality characterization of new sweet cherry cultivars as a good source of bioactive phenolic compounds with antioxidant and neuroprotective potential. Antioxidants 2020, 9, 677. [Google Scholar] [CrossRef] [PubMed]
  74. Boz, H. p-Coumaric acid in cereals: Presence, antioxidant and antimicrobial effects. Int. J. Food Sci. Technol. 2015, 50, 2323–2328. [Google Scholar] [CrossRef]
  75. Clifford, M.N. Anthocyanins–nature, occurrence and dietary burden. J. Sci. Food Agric. 2000, 80, 1063–1072. [Google Scholar] [CrossRef]
  76. Zhang, Y.; Sun, Y.; Zhang, H.; Mai, Q.; Zhang, B.; Li, H.; Deng, Z. The degradation rules of anthocyanins from eggplant peel and anti-oxidant capacity in fortified model food system during the thermal treatments. Food Biosci. 2020, 38, 100701. [Google Scholar] [CrossRef]
  77. Hollman, P.C.H.; Arts, I.C.W. Flavonols, flavones and flavanols–nature, occurrence and dietary burden. J. Sci. Food Agric. 2000, 80, 1081–1093. [Google Scholar] [CrossRef]
  78. Dabeek, W.M.; Marra, M.V. Dietary quercetin and kaempferol: Bioavailability and potential cardiovascular-related bioactivity in humans. Nutrients 2019, 11, 2288. [Google Scholar] [CrossRef]
  79. Crozier, A.; Lean, M.E.J.; McDonald, M.S.; Black, C. Quantitative Analysis of the flavonoid content of commercial tomatoes, onions, lettuce, and celery. J. Agric. Food Chem. 1997, 45, 590–595. [Google Scholar] [CrossRef]
  80. Vuorinen, H.; Määttä, K.; Törrönen, R. Content of the flavonols myricetin, quercetin, and kaempferol in Finnish berry wines. J. Agric. Food Chem. 2000, 48, 2675–2680. [Google Scholar] [CrossRef]
  81. Sung, B.; Chung, H.Y.; Kim, N.D. Role of apigenin in cancer prevention via the induction of apoptosis and autophagy. J. Cancer Prev. 2016, 21, 216–226. [Google Scholar] [CrossRef]
  82. Mushtaq, M.; Anwar, F. A Centum of Valuable Plant Bioactives; Academic Press: Cambridge, MA, USA, 2021; ISBN 9780128229231. [Google Scholar]
  83. Joven, J.; Micol, V.; Segura-Carretero, A.; Alonso-Villaverde, C.; Menéndez, J.A.; for the Bioactive Food Components Platform. Polyphenols and the modulation of gene expression pathways: Can we eat our way out of the danger of chronic disease? Crit. Rev. Food Sci. Nutr. 2014, 54, 985–1001. [Google Scholar] [CrossRef]
  84. Arts, I.C.; van de Putte, B.; Hollman, P.C. Catechin contents of foods commonly consumed in The Netherlands. 1. Fruits, vegetables, staple foods, and processed foods. J. Agric. Food Chem. 2000, 48, 1746–1751. [Google Scholar] [CrossRef]
  85. Shukla, A.S.; Jha, A.K.; Kumari, R.; Rawat, K.; Syeda, S.; Shrivastava, A. Role of catechins in chemosensitization. In Role of Nutraceuticals in Cancer Chemosensitization; Elsevier: Amsterdam, The Netherlands, 2018; pp. 169–198. [Google Scholar]
  86. Ravindranath, M.H.; Saravanan, T.S.; Monteclaro, C.C.; Presser, N.; Ye, X.; Selvan, S.R.; Brosman, S. Epicatechins purified from green tea (Camellia sinensis) differentially suppress growth of gender-dependent human cancer cell lines. Evid.-Based Complement. Altern. Med. 2006, 3, 105453. [Google Scholar] [CrossRef]
  87. Chen, L.; Cao, H.; Xiao, J. Polyphenols: Absorption, bioavailability, and metabolomics. In Polyphenols: Properties, Recovery, and Applications; Elsevier: Amsterdam, The Netherlands, 2018; pp. 45–67. [Google Scholar]
  88. Schieber, A.; Wüst, M. Volatile phenols—Important contributors to the aroma of plant-derived foods. Molecules 2020, 25, 4529. [Google Scholar] [CrossRef]
  89. Bresciani, L.; Tassotti, M.; Rosi, A.; Martini, D.; Antonini, M.; Dei Cas, A.; Mena, P. Absorption, pharmacokinetics, and urinary excretion of pyridines after consumption of coffee and cocoa-based products containing coffee in a repeated dose, crossover human intervention study. Mol. Nutr. Food Res. 2020, 64, 2000489. [Google Scholar] [CrossRef]
  90. Gómez-Juaristi, M.; Sarria, B.; Goya, L.; Bravo-Clemente, L.; Mateos, R. Experimental confounding factors affecting stability, transport and metabolism of flavanols and hydroxycinnamic acids in Caco-2 cells. Food Res. Int. 2019, 129, 108797. [Google Scholar] [CrossRef]
  91. Pan, L.L.; Yang, Y.; Hui, M.; Wang, S.; Li, C.Y.; Zhang, H.; Zhong, D.F. Sulfation predominates the pharmacokinetics, metabolism, and excretion of forsythin in humans: Major enzymes and transporters identified. Acta Pharmacol. Sin. 2021, 42, 311–322. [Google Scholar] [CrossRef]
  92. Káňová, K.; Petrásková, L.; Pelantová, H.; Rybková, Z.; Malachová, K.; Cvačka, J.; Křen, V.; Valentová, K. Sulfated metabolites of luteolin, myricetin, and ampelopsin: Chemoenzymatic preparation and biophysical properties. J. Agric. Food Chem. 2020, 68, 11197–11206. [Google Scholar] [CrossRef]
  93. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef]
  94. Jančová, P.; Šiller, M. Phase II drug metabolism. In Topics on Drug Metabolism; InTechOpen: London, UK, 2012; pp. 35–60. [Google Scholar] [CrossRef]
  95. Fumeaux, R.; Menozzi-Smarrito, C.; Stalmach, A.; Munari, C.; Kraehenbuehl, K.; Steiling, H.; Crozier, A.; Williamson, G.; Barron, D. First synthesis, characterization, and evidence for the presence of hydroxycinnamic acid sulfate and glucuronide conjugates in human biological fluids as a result of coffee consumption. Org. Biomol. Chem. 2010, 8, 5199–5211. [Google Scholar] [CrossRef]
  96. Nardini, M.; Forte, M.; Vrhovsek, U.; Mattivi, F.; Viola, R.; Scaccini, C. White wine phenolics are absorbed and extensively metabolized in humans. J. Agric. Food Chem. 2009, 57, 2711–2718. [Google Scholar] [CrossRef]
  97. Bourne, L.C.; Rice-Evans, C. Bioavailability of ferulic acid. Biochem. Biophys. Res. Commun. 1998, 253, 222–227. [Google Scholar] [CrossRef]
  98. Heleno, S.A.; Ferreira, I.C.; Ćirić, A.; Glamočlija, J.; Martins, A.; Queiroz, M.J.R.; Soković, M. Coprinopsis atramentaria extract, organic acids and synthesized methylated derivatives as antibacterial and antifungal agents. In Proceedings of the XIX Symposium of the Baltic Mycologists and Lichenologists, Skede, Latvia, 22–26 September 2014. [Google Scholar]
  99. Scalbert, A.; Williamson, G. Dietary intake and bioavailability of polyphenols. J. Nutr. 2000, 130, 2073S–2085S. [Google Scholar] [CrossRef]
  100. Plumb, G.W.; Garcia-Conesa, M.T.; Kroon, P.A.; Rhodes, M.; Ridley, S.; Williamson, G. Metabolism of chlorogenic acid by human plasma, liver, intestine and gut microflora. J. Sci. Food Agric. 1999, 79, 390–392. [Google Scholar] [CrossRef]
  101. Dama, A.; Shpati, K.; Daliu, P.; Dumur, S.; Gorica, E.; Santini, A. Targeting Metabolic Diseases: The Role of Nutraceuticals in Modulating Oxidative Stress and Inflammation. Nutrients 2024, 16, 507. [Google Scholar] [CrossRef]
  102. Kim, T.Y.; Leem, E.; Lee, J.M.; Kim, S.R. Control of reactive oxygen species for the prevention of Parkinson’s disease: The Possible Application of Flavonoids. Antioxidants 2020, 9, 583. [Google Scholar] [CrossRef]
  103. Farombi, E.O.; Akanni, O.O.; Emerole, G.O. Antioxidant and scavenging activities of flavonoid extract (Kolaviron) of Garcinia kola seeds. Pharm. Biol. 2002, 40, 107–116. [Google Scholar] [CrossRef]
  104. Peterson, J.J.; Dwyer, J.T.; Jacques, P.F.; McCullough, M.L. Associations between flavonoids and cardiovascular disease incidence or mortality in European and US populations. Nutr. Rev. 2012, 70, 491–508. [Google Scholar] [CrossRef]
  105. Arimboor, R.; Arumughan, C. Effect of polymerization on antioxidant and xanthine oxidase inhibitory potential of sea buck-thorn (H. rhamnoides) proanthocyanidins. J. Food Sci. 2012, 77, C1036–C1041. [Google Scholar] [CrossRef]
  106. Clifford, M.; Brown, J.E. Dietary flavonoids and health-broadening the perspective. In Flavonoids: Chemistry, Biochemistry and Applications; CRC Press LLC: Boca Raton, FL, USA, 2006; pp. 319–370. [Google Scholar]
  107. Zhu, L.; Li, W.; Deng, Z.; Li, H.; Zhang, B. The composition and antioxidant activity of bound phenolics in three legumes, and their metabolism and bioaccessibility of gastrointestinal tract. Foods 2020, 9, 1816. [Google Scholar] [CrossRef]
  108. Scalbert, A.; Morand, C.; Manach, C.; Rémésy, C. Absorption and metabolism of polyphenols in the gut and impact on health. Biomed. Pharmacother. 2002, 56, 276–282. [Google Scholar] [CrossRef]
  109. Landete, J.M. Updated knowledge about polyphenols: Functions, bioavailability, metabolism, and health. Crit. Rev. Food Sci. Nutr. 2012, 52, 936–948. [Google Scholar] [CrossRef]
  110. Del Rio, D.; Calani, L.; Scazzina, F.; Jechiu, L.; Cordero, C.; Brighenti, F. Bioavailability of catechins from ready-to-drink tea. Nutrition 2010, 26, 528–533. [Google Scholar] [CrossRef]
  111. Hollman, P.C.; Bijsman, M.N.; Van Gameren, Y.; Cnossen, E.P.; De Vries, J.H.; Katan, M.B. The sugar moiety is a major determinant of the absorption of dietary flavonoid glycosides in man. Free Radic. Res. 1999, 31, 569–573. [Google Scholar] [CrossRef]
  112. Khan, J.; Deb, P.K.; Priya, S.; Medina, K.D.; Devi, R.; Walode, S.G.; Rudrapal, M. Dietary flavonoids: Cardioprotective potential with antioxidant effects and their pharmacokinetic, toxicological and therapeutic concerns. Molecules 2021, 26, 4021. [Google Scholar] [CrossRef] [PubMed]
  113. Rudrapal, M.; Maji, S.; Prajapati, S.K.; Kesharwani, P.; Deb, P.K.; Khan, J.; Mohamed Ismail, R.; Kankate, R.S.; Sahoo, R.K.; Khairnar, S.J.; et al. Protective effects of diets rich in polyphenols in cigarette smoke (CS)-induced oxidative damages and associated health implications. Antioxidants 2022, 11, 1217. [Google Scholar] [CrossRef] [PubMed]
  114. Olivares-Vicente, M.; Barrajon-Catalan, E.; Herranz-Lopez, M.; Segura-Carretero, A.; Joven, J.; Encinar, J.A.; Micol, V. Plant-Derived Polyphenols in Human Health: Biological Activity, Metabolites and Putative Molecular Targets. Curr. Drug Metab. 2018, 19, 351–369. [Google Scholar] [CrossRef] [PubMed]
  115. Ohtani, H.; Iwata, R.; Imaoka, A.; Akiyoshi, T. Estimation of absolute oral bioavailability without performing an intravenous clinical study. Drug Metab. Pharmacokinet. 2021, 38, 100392. [Google Scholar] [CrossRef] [PubMed]
  116. Rich, G.T.; Buchweitz, M.; Winterbone, M.S.; Kroon, P.A.; Wilde, P.J. Towards an understanding of the low bioavailability of quercetin: A study of its interaction with intestinal lipids. Nutrients 2017, 9, 111. [Google Scholar] [CrossRef] [PubMed]
  117. Mullen, W.; Edwards, C.A.; Crozier, A. Absorption, excretion and metabolite profiling of methyl-, glucuronyl-, glucosyl- and sulpho-conjugates of quercetin in human plasma and urine after ingestion of onions. Br. J. Nutr. 2006, 96, 107–116. [Google Scholar] [CrossRef] [PubMed]
  118. Rupasinghe, H.V.; Ronalds, C.M.; Rathgeber, B.; Robinson, R.A. Absorption and tissue distribution of dietary quercetin and quercetin glycosides of apple skin in broiler chickens. J. Sci. Food Agric. 2010, 90, 1172–1178. [Google Scholar] [CrossRef] [PubMed]
  119. Thilakarathna, S.H.; Rupasinghe, H.P.V. Flavonoid bioavailability and attempts for bioavailability enhancement. Nutrients 2013, 5, 3367–3387. [Google Scholar] [CrossRef] [PubMed]
  120. Lv, L.; Gu, X.; Tang, J.; Ho, C.T. Antioxidant activity of stilbene glycoside from Polygonum multiflorum Thunb in vivo. Food Chem. 2007, 104, 1678–1681. [Google Scholar] [CrossRef]
  121. Singh, P.; Kumari, B.; Dahiya, G. Analysis of phenol and tannin in the bark extracts of medicinally important plant Moringa oleifera (Lam.) collected from different regions of Haryana. J. Pharmacogn. Phytochem. 2021, 10, 1917–1919. [Google Scholar]
  122. Yu, C.; Shin, Y.G.; Chow, A.; Li, Y.; Kosmeder, J.W.; Lee, Y.S.; Hirschelman, W.H.; Pezzuto, J.M.; Mehta, R.G.; van Breemen, R.B. Human, rat, and mouse metabolism of resveratrol. Pharm. Res. 2002, 19, 1907–1914. [Google Scholar] [CrossRef] [PubMed]
  123. Roupe, K.A.; Yáñez, J.A.; Teng, X.W.; Davies, N.M. Pharmacokinetics of selected stilbenes: Rhapontigenin, piceatannol and pinosylvin in rats. J. Pharm. Pharmacol. 2006, 58, 1443–1450. [Google Scholar] [CrossRef] [PubMed]
  124. Walle, T. Bioavailability of resveratrol. Ann. N. Y. Acad. Sci. 2011, 1215, 9–15. [Google Scholar] [CrossRef] [PubMed]
  125. Gowd, V.; Karim, N.; Shishir, M.R.I.; Xie, L.; Chen, W. Dietary polyphenols to combat the metabolic diseases via altering gut microbiota. Trends Food Sci. Technol. 2019, 93, 81–93. [Google Scholar] [CrossRef]
  126. Patel, K.R.; Brown, V.A.; Jones, D.J.; Britton, R.G.; Hemingway, D.; Miller, A.S.; West, K.P.; Booth, T.D.; Perloff, M.; Crowell, J.A.; et al. Clinical pharmacology of resveratrol and its metabolites in colorectal cancer patients. Cancer Res. 2010, 70, 7392–7399. [Google Scholar] [CrossRef] [PubMed]
  127. Sun, Y.; Zhao, Y. Enhanced pharmacokinetics and anti-tumor efficacy of PEGylated liposomal rhaponticin and plasma protein binding ability of rhaponticin. J. Nanosci. Nanotechnol. 2012, 12, 7677–7684. [Google Scholar] [CrossRef] [PubMed]
  128. Setoguchi, Y.; Oritani, Y.; Ito, R.; Inagaki, H.; Maruki-Uchida, H.; Ichiyanagi, T.; Ito, T. Absorption and metabolism of piceatannol in rats. J. Agric. Food Chem. 2014, 62, 2541–2548. [Google Scholar] [CrossRef] [PubMed]
  129. Huang, H.; Chen, G.; Lu, Z.; Zhang, J.; Guo, D. Identification of seven metabolites of oxyresveratrol in rat urine and bile using liquid chromatography/tandem mass spectrometry. Biomed. Chromatogr. 2009, 24, 426–432. [Google Scholar] [CrossRef] [PubMed]
  130. Kahle, K.; Huemmer, W.; Kempf, M.; Scheppach, W.; Erk, T.; Richling, E. Polyphenols are intensively metabolized in the human gastrointestinal tract after apple juice consumption. J. Agric. Food Chem. 2007, 55, 10605–10614. [Google Scholar] [CrossRef]
  131. Cai, Z.-Y.; Li, X.-M.; Liang, J.-P.; Xiang, L.-P.; Wang, K.-R.; Shi, Y.-L.; Yang, R.; Shi, M.; Ye, J.-H.; Lu, J.-L.; et al. Bioavailability of tea catechins and its improvement. Molecules 2018, 23, 2346. [Google Scholar] [CrossRef]
  132. Whitley, A.C.; Stoner, G.D.; Darby, M.V.; Walle, T. Intestinal epithelial cell accumulation of the cancer preventive polyphenol ellagic acid—Extensive binding to protein and DNA. Biochem. Pharmacol. 2003, 66, 907–915. [Google Scholar] [CrossRef]
  133. Konishi, Y.; Zhao, Z.; Shimizu, M. Phenolic acids are absorbed from the rat stomach with different absorption rates. J. Agric. Food Chem. 2006, 54, 7539–7543. [Google Scholar] [CrossRef] [PubMed]
  134. Cerdá, B.; Tomás-Barberán, F.A.; Espín, J.C. Metabolism of antioxidant and chemopreventive ellagitannins from strawberries, raspberries, walnuts, and oak-aged wine in humans: Identification of biomarkers and individual variability. J. Agric. Food Chem. 2004, 53, 227–235. [Google Scholar] [CrossRef] [PubMed]
  135. Dini, I.; Grumetto, L. Recent Advances in Natural Polyphenol Research. Molecules 2022, 27, 8777. [Google Scholar] [CrossRef] [PubMed]
  136. Smeriglio, A.; Barreca, D.; Bellocco, E.; Trombetta, D. Proanthocyanidins and hydrolysable tannins: Occurrence, dietary intake and pharmacological effects. Br. J. Pharmacol. 2017, 174, 1244–1262. [Google Scholar] [CrossRef] [PubMed]
  137. Touré, A.; Xueming, X. Flaxseed lignans: Source, biosynthesis, metabolism, antioxidant activity, bio-active components, and health benefits. Compr. Rev. Food Sci. Food Saf. 2010, 9, 261–269. [Google Scholar] [CrossRef] [PubMed]
  138. Rowland, I. Role of the Gut Flora in Toxicity and Cancer; Elsevier: Amsterdam, The Netherlands, 1988; ISBN 9780125999205. [Google Scholar]
  139. Peñalvo, J.L.; Heinonen, S.-M.; Aura, A.-M.; Adlercreutz, H. Dietary sesamin is converted to enterolactone in humans. J. Nutr. 2005, 135, 1056–1062. [Google Scholar] [CrossRef] [PubMed]
  140. Matos, M.J.; Santana, L.; Uriarte, E.; Abreu, O.A.; Molina, E.; Yordi, E.G. Coumarins—An important class of phytochemicals. In Phytochemicals—Isolation, Characterisation and Role in Human Health; InTechOpen: London, UK, 2015; Volume 25, pp. 533–538. [Google Scholar] [CrossRef]
  141. Rehman, S.U.; Kim, I.S.; Kang, K.S.; Yoo, H.H. HPLC determination of esculin and esculetin in rat plasma for pharmacokinetic studies. J. Chromatogr. Sci. 2015, 53, 1322–1327. [Google Scholar] [CrossRef] [PubMed]
  142. Kwak, J.-H.; Kim, Y.; Staatz, C.E.; Baek, I.-H. Oral bioavailability and pharmacokinetics of esculetin following intravenous and oral administration in rats. Xenobiotica 2021, 51, 811–817. [Google Scholar] [CrossRef]
  143. Yu, X.A.; Azietaku, J.T.; Li, J.; An, M.; He, J.; Hao, J.; Chang, Y.X. The pharmacokinetics, bioavailability and excretion of bergapten after oral and intravenous administration in rats using high performance liquid chromatography with fluorescence detection. Chem. Cent. J. 2016, 10, 62. [Google Scholar] [CrossRef]
  144. Spencer, J.P.E.; El Mohsen, M.M.A.; Minihane, A.-M.; Mathers, J.C. Biomarkers of the intake of dietary polyphenols: Strengths, limitations and application in nutrition research. Br. J. Nutr. 2007, 99, 12–22. [Google Scholar] [CrossRef] [PubMed]
  145. Kondratyuk, T.P.; Pezzuto, J.M. Natural product polyphenols of relevance to human health. Pharm. Biol. 2004, 42 (Suppl. 1), 46–63. [Google Scholar] [CrossRef]
  146. Cory, H.; Passarelli, S.; Szeto, J.; Tamez, M.; Mattei, J. The role of polyphenols in human health and food systems: A mini-review. Front. Nutr. 2018, 5, 87. [Google Scholar] [CrossRef] [PubMed]
  147. Nunes, C.; Ferreira, E.; Freitas, V.; Almeida, L.; Barbosa, R.M.; Laranjinha, J. Intestinal anti-inflammatory activity of red wine extract: Unveiling the mechanisms in colonic epithelial cells. Food Funct. 2012, 4, 373–383. [Google Scholar] [CrossRef] [PubMed]
  148. Seeram, N.P.; Adams, L.S.; Henning, S.M.; Niu, Y.; Zhang, Y.; Nair, M.G.; Heber, D. In vitro antiproliferative, apoptotic and antioxidant activities of punicalagin, ellagic acid and a total pomegranate tannin extract are enhanced in combination with other poly-phenols as found in pomegranate juice. J. Nutr. Biochem. 2005, 16, 360–367. [Google Scholar] [CrossRef] [PubMed]
  149. Graziani, G.; D’argenio, G.; Tuccillo, C.; Loguercio, C.; Ritieni, A.; Morisco, F.; Blanco, C.D.V.; Fogliano, V.; Romano, M. Apple polyphenol extracts prevent damage to human gastric epithelial cells in vitro and to rat gastric mucosa in vivo. Gut 2005, 54, 193–200. [Google Scholar] [CrossRef] [PubMed]
  150. Schaefer, S.; Baum, M.; Eisenbrand, G.; Dietrich, H.; Will, F.; Janzowski, C. Polyphenolic apple juice extracts and their major constituents reduce oxidative damage in human colon cell lines. Mol. Nutr. Food Res. 2006, 50, 24–33. [Google Scholar] [CrossRef] [PubMed]
  151. Chedea, V.S.; Palade, L.M.; Marin, D.E.; Pelmus, R.S.; Habeanu, M.; Rotar, M.C.; Gras, M.A.; Pistol, G.C.; Taranu, I. Intestinal absorption and antioxidant activity of grape pomace polyphenols. Nutrients 2018, 10, 588. [Google Scholar] [CrossRef] [PubMed]
  152. Bornsek, S.M.; Ziberna, L.; Polak, T.; Vanzo, A.; Ulrih, N.P.; Abram, V.; Tramer, F.; Passamonti, S. Bilberry and blueberry anthocyanins act as powerful intracellular antioxidants in mammalian cells. Food Chem. 2012, 134, 1878–1884. [Google Scholar] [CrossRef] [PubMed]
  153. Oz, H.S.; Chen, T.S.; McClain, C.J.; de Villiers, W.J. Antioxidants as novel therapy in a murine model of colitis. J. Nutr. Biochem. 2005, 16, 297–304. [Google Scholar] [CrossRef]
  154. Scarano, A.; Butelli, E.; De Santis, S.; Cavalcanti, E.; Hill, L.; De Angelis, M.; Giovinazzo, G.; Chieppa, M.; Martin, C.; Santino, A. Combined dietary anthocyanins, flavonols, and stilbenoids alleviate inflammatory bowel disease symptoms in mice. Front. Nutr. 2018, 4, 75. [Google Scholar] [CrossRef]
  155. Sato, Y.; Itagaki, S.; Kurokawa, T.; Ogura, J.; Kobayashi, M.; Hirano, T.; Sugawara, M.; Iseki, K. In vitro and in vivo antioxidant properties of chlorogenic acid and caffeic acid. Int. J. Pharm. 2010, 403, 136–138. [Google Scholar] [CrossRef]
  156. Nakatani, N.; Kayano, S.I.; Kikuzaki, H.; Sumino, K.; Katagiri, K.; Mitani, T. Identification, quantitative determination, and antioxidative activities of chlorogenic acid isomers in prune (Prunus domestica L.). J. Agric. Food Chem. 2000, 48, 5512–5516. [Google Scholar] [CrossRef] [PubMed]
  157. Zhou, Y. The protective effects of cryptochlorogenic acid on β-cells function in diabetes in vivo and vitro via inhibition of ferroptosis. Diabetes Metab. Syndr. Obes. 2020, 13, 1921–1931. [Google Scholar] [CrossRef]
  158. Hirohashi, K.; Ohashi, S.; Amanuma, Y.; Nakai, Y.; Ida, T.; Baba, K.; Mitani, Y.; Mizumoto, A.; Yamamoto, Y.; Kikuchi, O.; et al. Protective effects of Alda-1, an ALDH2 activator, on alcohol-derived DNA damage in the esophagus of human ALDH2*2 (Glu504Lys) knock-in mice. Carcinogenesis 2019, 41, 194–202. [Google Scholar] [CrossRef] [PubMed]
  159. Xu, J.-G.; Hu, Q.-P.; Liu, Y. Antioxidant and DNA-Protective Activities of Chlorogenic Acid Isomers. J. Agric. Food Chem. 2012, 60, 11625–11630. [Google Scholar] [CrossRef] [PubMed]
  160. Stefanowicz-Hajduk, J.; Król-Kogus, B.; Sparzak-Stefanowska, B.; Kimel, K.; Ochocka, J.R.; Krauze-Baranowska, M. Cytotoxic activity of standardized extracts, a fraction, and individual secondary metabolites from fenugreek seeds against SKOV-3, HeLa and MOLT-4 cell lines. Pharm. Biol. 2021, 59, 422–435. [Google Scholar] [CrossRef]
  161. Yagasaki, K.; Miura, Y.; Okauchi, R.; Furuse, T. Inhibitory effects of chlorogenic acid and its related compounds on the invasion of hepatoma cells in culture. Cytotechnology 2000, 33, 229–235. [Google Scholar] [CrossRef]
  162. Morishita, H.; Kido, R. Antioxidant activities of chlorogenic acids. In Colloque Scientifique International sur le Café; ASIC Association Scientifique Internationale: Paris, France, 1995; pp. 119–124. [Google Scholar]
  163. Mehrzadi, S.; Khalili, H.; Fatemi, I.; Malayeri, A.; Siahpoosh, A.; Goudarzi, M. Zingerone mitigates carrageenan-induced inflammation through antioxidant and anti-inflammatory activities. Inflammation 2021, 44, 186–193. [Google Scholar] [CrossRef]
  164. Mitrea, D.R.; Malkey, R.; Florian, T.L.; Filip, A.; Clichici, S.; Bidian, C.; Baldea, I. Daily oral administration of chlorogenic acid prevents the experimental carrageenan-induced oxidative stress. J. Physiol. Pharmacol. 2020, 71, 55–65. [Google Scholar]
  165. El-Khadragy, M.F.; Al-Megrin, W.A.; Alomar, S.; Alkhuriji, A.F.; Metwally, D.M.; Mahgoub, S.; Albeltagy, R.S. Chlorogenic acid abates male reproductive dysfunction in arsenic-exposed mice via attenuation of testicular oxido-inflammatory stress and apoptotic responses. Chem. Biol. Interact. 2021, 333, 109333. [Google Scholar] [CrossRef]
  166. Huang, K.; Liang, X.C.; Zhong, Y.L.; He, W.Y.; Wang, Z. 5-Caffeoylquinic acid decreases diet-induced obesity in rats by modulating PPARα and LXRα transcription. J. Sci. Food Agric. 2015, 95, 1903–1910. [Google Scholar] [CrossRef] [PubMed]
  167. Jin, S.; Chang, C.; Zhang, L.; Liu, Y.; Huang, X.; Chen, Z. Chlorogenic acid improves late diabetes through adiponectin receptor signaling pathways in db/db mice. PLoS ONE 2015, 10, e0120842. [Google Scholar] [CrossRef]
  168. Karunanidhi, A.; Thomas, R.; Van Belkum, A.; Neela, V. In vitro antibacterial and antibiofilm activities of chlorogenic acid against clinical isolates of Stenotrophomonas maltophilia including the trimethoprim/sulfamethoxazole resistant strain. BioMed Res. Int. 2013, 2013, 392058. [Google Scholar] [CrossRef]
  169. Suzuki, A.; Kagawa, D.; Ochiai, R.; Tokimitsu, I.; Saito, I. Green coffee bean extract and its metabolites have a hypotensive effect in spontaneously hypertensive rats. Hypertens. Res. 2002, 25, 99–107. [Google Scholar] [CrossRef]
  170. Meng, S.; Cao, J.; Feng, Q.; Peng, J.; Hu, Y. Roles of chlorogenic acid on regulating glucose and lipids metabolism: A review. Evid.-Based Complement. Altern. Med. 2013, 2013, 801457. [Google Scholar] [CrossRef]
  171. Chen, S.; Li, M.; Li, Y.; Hu, H.; Li, Y.; Huang, Y.; Wang, Y. A UPLC-ESI-MS/MS method for simultaneous quantitation of chlorogenic acid, scutellarin, and scutellarein in rat plasma: Application to a comparative pharmacokinetic study in sham-operated and MCAO rats after oral administration of Erigeron breviscapus extract. Molecules 2018, 23, 1808. [Google Scholar] [CrossRef]
  172. Lee, T.-K.; Park, J.H.; Ahn, J.H.; Kim, H.; Song, M.; Lee, J.-C.; Kim, J.D.; Jeon, Y.H.; Choi, J.H.; Lee, C.H.; et al. Pretreatment of Populus tomentiglandulosa protects hippocampal CA1 pyramidal neurons from ischemia-reperfusion injury in gerbils via increasing SODs expressions and maintaining BDNF and IGF-I expressions. Chin. J. Nat. Med. 2019, 17, 424–434. [Google Scholar] [CrossRef] [PubMed]
  173. Park, Y.E.; Noh, Y.; Kim, D.W.; Lee, T.-K.; Ahn, J.H.; Kim, B.; Lee, J.-C.; Park, C.W.; Park, J.H.; Kim, J.D.; et al. Experimental pretreatment with YES-10®, a plant extract rich in scutellarin and chlorogenic acid, protects hippocampal neurons from ischemia/reperfusion injury via antioxidant role. Exp. Ther. Med. 2021, 21, 183. [Google Scholar] [CrossRef]
  174. Gaspar, A.; Martins, M.; Silva, P.; Garrido, E.M.; Garrido, J.; Firuzi, O.; Miri, R.; Saso, L.; Borges, F. Dietary phenolic acids and derivatives. evaluation of the antioxidant activity of sinapic acid and its alkyl esters. J. Agric. Food Chem. 2010, 58, 11273–11280. [Google Scholar] [CrossRef]
  175. Nićiforović, N.; Abramovič, H. Sinapic acid and its derivatives: Natural sources and bioactivity. Compr. Rev. Food Sci. Food Saf. 2014, 13, 34–51. [Google Scholar] [CrossRef]
  176. Cartea, M.E.; Francisco, M.; Soengas, P.; Velasco, P. Phenolic compounds in brassica vegetables. Molecules 2011, 16, 251–280. [Google Scholar] [CrossRef]
  177. Fabre, N.; Urizzi, P.; Souchard, J.; Fréchard, A.; Claparols, C.; Fourasté, I.; Moulis, C. An antioxidant sinapic acid ester isolated from Iberis amara. Fitoterapia 2000, 71, 425–428. [Google Scholar] [CrossRef]
  178. Siger, A.; Czubinski, J.; Dwiecki, K.; Kachlicki, P.; Nogala-Kalucka, M. Identification and antioxidant activity of sinapic acid derivatives in Brassica napus L. seed meal extracts. Eur. J. Lipid Sci. Technol. 2013, 115, 1130–1138. [Google Scholar] [CrossRef]
  179. Bešlo, D.; Golubić, N.; Rastija, V.; Agić, D.; Karnaš, M.; Šubarić, D.; Lučić, B. Antioxidant Activity, Metabolism, and Bioavailability of Polyphenols in the Diet of Animals. Antioxidants 2023, 12, 1141. [Google Scholar] [CrossRef]
  180. Silambarasan, T.; Manivannan, J.; Priya, M.K.; Suganya, N.; Chatterjee, S.; Raja, B. Sinapic acid prevents hypertension and cardiovascular remodeling in pharmacological model of nitric oxide inhibited rats. PLoS ONE 2014, 9, e115682. [Google Scholar] [CrossRef]
  181. Şahin, M.; Öncü, G.; Yılmaz, M.A.; Özkan, D.; Saybaşılı, H. Transformation of SH-SY5Y cell line into neuron-like cells: Investigation of electrophysiological and biomechanical changes. Neurosci. Lett. 2021, 745, 135628. [Google Scholar] [CrossRef]
  182. Tungalag, T.; Yang, D.K. Sinapic acid protects SH-SY5Y human neuroblastoma cells against 6-hydroxydopamine-induced neurotoxicity. Biomedicines 2021, 9, 295. [Google Scholar] [CrossRef]
  183. Graf, E. Antioxidant potential of ferulic acid. Free Radic. Biol. Med. 1992, 13, 435–448. [Google Scholar] [CrossRef]
  184. Kikuzaki, H.; Hisamoto, M.; Hirose, K.; Akiyama, K.; Taniguchi, H. Antioxidant properties of ferulic acid and its related compounds. J. Agric. Food Chem. 2002, 50, 2161–2168. [Google Scholar] [CrossRef]
  185. Kumar, N.; Goel, N. Phenolic acids: Natural versatile molecules with promising therapeutic applications. Biotechnol. Rep. 2019, 24, e00370. [Google Scholar] [CrossRef]
  186. Linfert, D.; Chowdhry, T.; Rabb, H. Lymphocytes and ischemia-reperfusion injury. Transplant. Rev. 2009, 23, 1–10. [Google Scholar] [CrossRef]
  187. Ren, Z.; Zhang, R.; Li, Y.; Li, Y.; Yang, Z.; Yang, H. Ferulic acid exerts neuroprotective effects against cerebral ischemia/reperfusion-induced injury via antioxidant and anti-apoptotic mechanisms in vitro and in vivo. Int. J. Mol. Med. 2017, 40, 1444–1456. [Google Scholar] [CrossRef]
  188. Cuadrado, A.; Nebreda, A.R. Mechanisms and functions of p38 MAPK signalling. Biochem. J. 2010, 429, 403–417. [Google Scholar] [CrossRef]
  189. Barrajón-Catalán, E.; Herranz-López, M.; Joven, J.; Segura-Carretero, A.; Alonso-Villaverde, C.; Menendez, J.A.; Micol, V. Molecular promiscuity of plant polyphenols in the management of age-related diseases: Far beyond their antioxidant properties. Adv. Exp. Med. Biol. 2014, 824, 141–159. [Google Scholar] [CrossRef]
  190. Koh, P.-O. Ferulic acid modulates nitric oxide synthase expression in focal cerebral ischemia. Lab. Anim. Res. 2012, 28, 273–278. [Google Scholar] [CrossRef]
  191. Na Yin, Z.; Wu, W.J.; Sun, C.Z.; Liu, H.F.; Chen, W.B.; Zhan, Q.P.; Lei, Z.G.; Xin, X.; Ma, J.J.; Yao, K.; et al. Antioxidant and anti-inflammatory capacity of ferulic acid released from wheat bran by solid-state fermentation of Aspergillus niger. Biomed. Environ. Sci. 2019, 32, 11–21. [Google Scholar] [CrossRef]
  192. Li, D.; Liang, H.; Li, Y.; Zhang, J.; Qiao, L.; Luo, H. Allicin alleviates lead-induced bone loss by preventing oxidative stress and osteoclastogenesis via SIRT1/FOXO1 pathway in mice. Biol. Trace Elem. Res. 2021, 199, 237–243. [Google Scholar] [CrossRef]
  193. Cho, S.-Y.; Chung, Y.-S.; Kwak, Y.-S.; Roh, H.-T. Gender-Specific Changes of Plasma MDA, SOD, and Lymphocyte DNA Damage during High Intensity Exercise. J. Life Sci. 2011, 21, 838–844. [Google Scholar] [CrossRef]
  194. Kelainy, E.G.; Laila, I.M.I.; Ibrahim, S.R. The effect of ferulic acid against lead-induced oxidative stress and DNA damage in kidney and testes of rats. Environ. Sci. Pollut. Res. 2019, 26, 31675–31684. [Google Scholar] [CrossRef]
  195. Wang, Y.; Chen, X.; Huang, Z.; Chen, D.; Yu, B.; Yu, J.; Chen, H.; He, J.; Luo, Y.; Zheng, P. Dietary ferulic acid supplementation improves antioxidant capacity and lipid metabolism in weaned piglets. Nutrients 2020, 12, 3811. [Google Scholar] [CrossRef] [PubMed]
  196. Hinton, D.R.; Kannan, R. Sodium iodate induced retinal degeneration: New insights from an old model. Neural Regen. Res. 2014, 9, 2044–2045. [Google Scholar] [CrossRef]
  197. Kohno, M.; Musashi, K.; Ikeda, H.O.; Horibe, T.; Matsumoto, A.; Kawakami, K. Oral administration of ferulic acid or ethyl ferulate attenuates retinal damage in sodium iodate-induced retinal degeneration mice. Sci. Rep. 2020, 10, 8688. [Google Scholar] [CrossRef]
  198. Pal, S.; Maity, S.; Sardar, S.; Begum, S.; Dalui, R.; Parvej, H.; Bera, K.; Pradhan, A.; Sepay, N.; Paul, S.; et al. Antioxidant ferulic acid prevents the aggregation of bovine β-lactoglobulin in vitro. J. Chem. Sci. 2020, 132, 103. [Google Scholar] [CrossRef]
  199. Lambruschini, C.; Demori, I.; El Rashed, Z.; Rovegno, L.; Canessa, E.; Cortese, K.; Grasselli, E.; Moni, L. Synthesis, photoisomerization, antioxidant activity, and lipid-lowering effect of ferulic acid and feruloyl amides. Molecules 2020, 26, 89. [Google Scholar] [CrossRef] [PubMed]
  200. Yu, L.; Wen, H.; Jiang, M.; Wu, F.; Tian, J.; Lu, X.; Liu, W. Effects of ferulic acid on growth performance, immunity and antioxidant status in genetically improved farmed tilapia (Oreochromis niloticus) fed oxidized fish oil. Aquac. Nutr. 2020, 26, 1431–1442. [Google Scholar] [CrossRef]
  201. Wu, Y.; Shi, Y.G.; Zheng, X.L.; Dang, Y.L.; Zhu, C.M.; Zhang, R.R.; Li, J.H. Lipophilic ferulic acid derivatives protect PC12 cells against oxidative damage via modulating β-amyloid aggregation and activating Nrf2 enzymes. Food Funct. 2020, 11, 4707–4718. [Google Scholar] [CrossRef] [PubMed]
  202. Umaru, I.J.; Ahmed, F.B.; Umaru, K.I. Cinnamic Acid from Barringtonia asiatica Stem-Bark Extract. Its Cytotoxicity, Antioxidant and Bioactive Potentials on Some Selected Pathogens. Int. J. Recent Biotechnol. 2020, 8, 35–46. [Google Scholar] [CrossRef]
  203. Ruwizhi, N.; Aderibigbe, B.A. Cinnamic acid derivatives and their biological efficacy. Int. J. Mol. Sci. 2020, 21, 5712. [Google Scholar] [CrossRef]
  204. Lima, T.C.; Ferreira, A.R.; Silva, D.F.; Lima, E.O.; de Sousa, D.P. Antifungal activity of cinnamic acid and benzoic acid esters against Candida albicans strains. Nat. Prod. Res. 2018, 32, 572–575. [Google Scholar] [CrossRef]
  205. Rathod, N.B.; Elabed, N.; Punia, S.; Ozogul, F.; Kim, S.-K.; Rocha, J.M. Recent Developments in Polyphenol Applications on Human Health: A Review with Current Knowledge. Plants 2023, 12, 1217. [Google Scholar] [CrossRef] [PubMed]
  206. Kakkar, S.; Bais, S. A review on protocatechuic acid and its pharmacological potential. ISRN Pharmacol. 2014, 2014, 952943. [Google Scholar] [CrossRef] [PubMed]
  207. Khan, A.K.; Rashid, R.; Fatima, N.; Mahmood, S.; Mir, S.; Khan, S.; Jabeen, N.; Murtaza, G. Pharmacological activities of protocatechuic acid. Acta Pol. Pharm 2015, 72, 643–650. [Google Scholar] [PubMed]
  208. Zheng, J.; Li, Q.; He, L.; Weng, H.; Su, D.; Liu, X.; Ling, W.; Wang, D. Protocatechuic acid inhibits vulnerable atherosclerotic lesion progression in older Apoe-/- mice. J. Nutr. 2020, 150, 1167–1177. [Google Scholar] [CrossRef] [PubMed]
  209. Al Olayan, E.M.; Aloufi, A.S.; AlAmri, O.D.; El-Habit, O.H.; Moneim, A.E.A. Protocatechuic acid mitigates cadmium-induced neurotoxicity in rats: Role of oxidative stress, inflammation and apoptosis. Sci. Total Environ. 2020, 723, 137969. [Google Scholar] [CrossRef] [PubMed]
  210. Aruoma, O.I.; Murcia, A.; Butler, J.; Halliwell, B. Evaluation of the antioxidant and prooxidant actions of gallic acid and its derivatives. J. Agric. Food Chem. 1993, 41, 1880–1885. [Google Scholar] [CrossRef]
  211. Badhani, B.; Sharma, N.; Kakkar, R. Gallic acid: A versatile antioxidant with promising therapeutic and industrial applications. RSC Adv. 2015, 5, 27540–27557. [Google Scholar] [CrossRef]
  212. Zhou, D.; Yang, Q.; Tian, T.; Chang, Y.; Li, Y.; Duan, L.R.; Wang, S.W. Gastroprotective effect of gallic acid against ethanol-induced gastric ulcer in rats: Involvement of the Nrf2/HO-1 signaling and anti-apoptosis role. Biomed. Pharmacother. 2020, 126, 110075. [Google Scholar] [CrossRef]
  213. Ojo, O.A.; Rotimi, D.; Bright, A.E.; Kayode, O.T.; Ojo, A.B.; Alejolowo, O.O.; Ajiboye, B.O.; Oluba, O.M. Gallic acid protects against cadmium chloride-induced alterations in Wistar rats via the antioxidant defense mechanism. J. Pharm. Pharmacogn. Res. 2021, 9, 668–676. [Google Scholar] [CrossRef]
  214. Rudrapal, M.; Chetia, D. Plant flavonoids as potential source of future antimalarial leads. Syst. Rev. Pharm. 2017, 8, 13–18. [Google Scholar] [CrossRef]
  215. Pekkarinen, S.S.; Heinonen, I.M.; Hopia, A.I. Flavonoids quercetin, myricetin, kaemferol and (+)-catechin as antioxidants in methyl linoleate. J. Sci. Food Agric. 1999, 79, 499–506. [Google Scholar] [CrossRef]
  216. Wang, H.; Joseph, J.A. Structure–activity relationships of quercetin in antagonizing hydrogen peroxide-induced calcium dysregulation in PC12 cells. Free Radic. Biol. Med. 1999, 27, 683–694. [Google Scholar] [CrossRef] [PubMed]
  217. Burda, S.; Oleszek, W. Antioxidant and antiradical activities of flavonoids. J. Agric. Food Chem. 2001, 49, 2774–2779. [Google Scholar] [CrossRef]
  218. Rice-Evans, C. Flavonoid antioxidants. Curr. Med. Chem. 2001, 8, 797–807. [Google Scholar] [CrossRef]
  219. Arora, A.; Nair, M.G.; Strasburg, G.M. Structure–activity relationships for antioxidant activities of a series of flavonoids in a liposomal system. Free Radic. Biol. Med. 1998, 24, 1355–1363. [Google Scholar] [CrossRef]
  220. Leopoldini, M.; Russo, N.; Chiodo, S.; Toscano, M. Iron chelation by the powerful antioxidant flavonoid quercetin. J. Agric. Food Chem. 2006, 54, 6343–6351. [Google Scholar] [CrossRef]
  221. 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]
  222. Peng, Z.; Gong, X.; Yang, Y.; Huang, L.; Zhang, Q.; Zhang, P.; Zhang, B. Hepatoprotective effect of quercetin against LPS/d-GalN induced acute liver injury in mice by inhibiting the IKK/NF-κB and MAPK signal pathways. Int. Immunopharmacol. 2017, 52, 281–289. [Google Scholar] [CrossRef] [PubMed]
  223. Jin, Y.; Huang, Z.L.; Li, L.; Yang, Y.; Wang, C.H.; Wang, Z.T.; Ji, L.L. Quercetin attenuates toosendanin-induced hepatotoxicity through inducing the Nrf2/GCL/GSH antioxidant signaling pathway. Acta Pharmacol. Sin. 2019, 40, 75–85. [Google Scholar] [CrossRef]
  224. Zhu, M.; Zhou, X.; Zhao, J. Quercetin prevents alcohol induced liver injury through targeting of PI3K/Akt/nuclear factor κB and STAT3 signaling pathway. Exp. Ther. Med. 2017, 14, 6169–6175. [Google Scholar] [CrossRef]
  225. Ma, J.Q.; Li, Z.; Xie, W.R.; Liu, C.M.; Liu, S.S. Quercetin protects mouse liver against CCl4-induced inflammation by the TLR2/4 and MAPK/NF-κB pathway. Int. Immunopharmacol. 2015, 28, 531–539. [Google Scholar] [CrossRef] [PubMed]
  226. Pratheeshkumar, P.; Son, Y.-O.; Divya, S.P.; Wang, L.; Turcios, L.; Roy, R.V.; Hitron, J.A.; Kim, D.; Dai, J.; Asha, P.; et al. Quercetin inhibits Cr(VI)-induced malignant cell transformation by targeting miR-21-PDCD4 signaling pathway. Oncotarget 2016, 8, 52118–52131. [Google Scholar] [CrossRef] [PubMed]
  227. Dong, F.; Wang, S.; Wang, Y.; Yang, X.; Jiang, J.; Wu, D.; Qu, X.; Fan, H.; Yao, R. Quercetin ameliorates learning and memory via the Nrf2-ARE signaling pathway in d-galactose-induced neurotoxicity in mice. Biochem. Biophys. Res. Commun. 2017, 491, 636–641. [Google Scholar] [CrossRef] [PubMed]
  228. Wattel, A.; Kamel, S.; Prouillet, C.; Petit, J.; Lorget, F.; Offord, E.; Brazier, M. Flavonoid quercetin decreases osteoclastic differentiation induced by RANKL via a mechanism involving NFκB and AP-1. J. Cell. Biochem. 2004, 92, 285–295. [Google Scholar] [CrossRef] [PubMed]
  229. Magar, R.T.; Sohng, J.K. A review on structure, modifications and structure-activity relation of quercetin and its derivatives. J. Microbiol. Biotechnol. 2019, 30, 11–20. [Google Scholar] [CrossRef] [PubMed]
  230. Xu, D.; Hu, M.-J.; Wang, Y.-Q.; Cui, Y.-L. Antioxidant activities of quercetin and its complexes for medicinal application. Molecules 2019, 24, 1123. [Google Scholar] [CrossRef]
  231. Zhou, J.; Wang, L.F.; Wang, J.Y.; Tang, N. Synthesis, characterization, antioxidative and antitumor activities of solid quercetin rare earth (III) complexes. J. Inorg. Biochem. 2001, 83, 41–48. [Google Scholar] [CrossRef] [PubMed]
  232. Jullian, C.; Moyano, L.; Yañez, C.; Olea-Azar, C. Complexation of quercetin with three kinds of cyclodextrins: An antioxidant study. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2007, 67, 230–234. [Google Scholar] [CrossRef]
  233. Bukhari, S.B.; Memon, S.; Mahroof-Tahir, M.; Bhanger, M.I. Synthesis, characterization and antioxidant activity copper–quercetin complex. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2009, 71, 1901–1906. [Google Scholar] [CrossRef]
  234. Wu, T.-H.; Yen, F.-L.; Lin, L.-T.; Tsai, T.-R.; Lin, C.-C.; Cham, T.-M. Preparation, physicochemical characterization, and antioxidant effects of quercetin nanoparticles. Int. J. Pharm. 2008, 346, 160–168. [Google Scholar] [CrossRef]
  235. Bukhari, S.B.; Memon, S.; Tahir, M.M.; Bhanger, M.I. Synthesis, characterization and investigation of antioxidant activity of cobalt–quercetin complex. J. Mol. Struct. 2008, 892, 39–46. [Google Scholar] [CrossRef]
  236. Dehghan, G.; Khoshkam, Z. Tin(II)–quercetin complex: Synthesis, spectral characterisation and antioxidant activity. Food Chem. 2012, 131, 422–426. [Google Scholar] [CrossRef]
  237. Manach, C.; Morand, C.; Crespy, V.; Demigné, C.; Texier, O.; Régérat, F.; Rémésy, C. Quercetin is recovered in human plasma as conjugated derivatives which retain antioxidant properties. FEBS Lett. 1998, 426, 331–336. [Google Scholar] [CrossRef] [PubMed]
  238. Ravichandran, R.; Rajendran, M.; Devapiriam, D. Antioxidant study of quercetin and their metal complex and determination of stability constant by spectrophotometry method. Food Chem. 2014, 146, 472–478. [Google Scholar] [CrossRef] [PubMed]
  239. Song, X.; Wang, Y.; Gao, L. Mechanism of antioxidant properties of quercetin and quercetin-DNA complex. J. Mol. Model. 2020, 26, 133. [Google Scholar] [CrossRef] [PubMed]
  240. Birinci, Y.; Niazi, J.H.; Aktay-Çetin, O.; Basaga, H. Quercetin in the form of a nano-antioxidant (QTiO2) provides stabilization of quercetin and maximizes its antioxidant capacity in the mouse fibroblast model. Enzym. Microb. Technol. 2020, 138, 109559. [Google Scholar] [CrossRef]
  241. Oyedemi, S.O.; Nwaogu, G.; Chukwuma, C.I.; Adeyemi, O.T.; Matsabisa, M.G.; Swain, S.S.; Aiyegoro, O.A. Quercetin modulates hyperglycemia by improving the pancreatic antioxidant status and enzymes activities linked with glucose metabolism in type 2 diabetes model of rats: In silico studies of molecular interaction of quercetin with hexokinase and catalase. J. Food Biochem. 2019, 44, e13127. [Google Scholar] [CrossRef] [PubMed]
  242. Halevas, E.; Pekou, A.; Papi, R.; Mavroidi, B.; Hatzidimitriou, A.G.; Zahariou, G.; Litsardakis, G.; Sagnou, M.; Pelecanou, M.; Pantazaki, A.A. Synthesis, physicochemical characterization and biological properties of two novel Cu(II) complexes based on natural products curcumin and quercetin. J. Inorg. Biochem. 2020, 208, 111083. [Google Scholar] [CrossRef] [PubMed]
  243. Tvrdá, E.; Debacker, M.; Ďuračka, M.; Kováč, J.; Bučko, O. Quercetin and naringenin provide functional and antioxidant protection to stored boar semen. Animals 2020, 10, 1930. [Google Scholar] [CrossRef]
  244. Yuan, F.; Wang, P.; Yang, Y.; Shi, P.; Cheng, L. Quercetin-albumin nano-complex as an antioxidant agent against hydrogen peroxide-induced death of spinal cord neurons as a model of preventive care study. Arab. J. Chem. 2020, 13, 8172–8182. [Google Scholar] [CrossRef]
  245. Pan, X.; Liu, X.; Zhao, H.; Wu, B.; Liu, G. Antioxidant, anti-inflammatory and neuroprotective effect of kaempferol on rotenone-induced Parkinson’s disease model of rats and SH-S5Y5 cells by preventing loss of tyrosine hydroxylase. J. Funct. Foods 2020, 74, 104140. [Google Scholar] [CrossRef]
  246. Oueslati, M.H.; Ben Tahar, L.; Harrath, A.H. Catalytic, antioxidant and anticancer activities of gold nanoparticles synthesized by kaempferol glucoside from Lotus leguminosae. Arab. J. Chem. 2018, 13, 3112–3122. [Google Scholar] [CrossRef]
  247. Hofer, S.; Geisler, S.; Lisandrelli, R.; Ngoc, H.N.; Ganzera, M.; Schennach, H.; Fuchs, D.; Fuchs, J.E.; Gostner, J.M.; Kurz, K. Pharmacological targets of kaempferol within inflammatory pathways—A hint towards the central role of tryptophan metabolism. Antioxidants 2020, 9, 180. [Google Scholar] [CrossRef] [PubMed]
  248. Tatsimo, S.J.N.; Tamokou, J.D.D.; Havyarimana, L.; Csupor, D.; Forgo, P.; Hohmann, J.; Kuiate, J.-R.; Tane, P. Antimicrobial and antioxidant activity of kaempferol rhamnoside derivatives from Bryophyllum pinnatum. BMC Res. Notes 2012, 5, 158. [Google Scholar] [CrossRef] [PubMed]
  249. Wang, J.; Fang, X.; Ge, L.; Cao, F.; Zhao, L.; Wang, Z.; Xiao, W. Antitumor, antioxidant and anti-inflammatory activities of kaempferol and its corresponding glycosides and the enzymatic preparation of kaempferol. PLoS ONE 2018, 13, e0197563. [Google Scholar] [CrossRef] [PubMed]
  250. Karim, N.; Shishir, M.R.I.; Gowd, V.; Chen, W. Hesperidin—An Emerging Bioactive Compound against Metabolic Diseases and Its Potential Biosynthesis Pathway in Microorganism. Food Rev. Int. 2021, 38 (Suppl. 1), 170–192. [Google Scholar] [CrossRef]
  251. Wilmsen, P.K.; Spada, D.S.; Salvador, M. Antioxidant activity of the flavonoid hesperidin in chemical and biological systems. J. Agric. Food Chem. 2005, 53, 4757–4761. [Google Scholar] [CrossRef]
  252. Cho, J. Antioxidant and neuroprotective effects of hesperidin and its aglycone hesperetin. Arch. Pharmacal Res. 2006, 29, 699–706. [Google Scholar] [CrossRef] [PubMed]
  253. Etcheverry, S.B.; Ferrer, E.G.; Naso, L.; Rivadeneira, J.; Salinas, V.; Williams, P.A.M. Antioxidant effects of the VO(IV) hesperidin complex and its role in cancer chemoprevention. JBIC J. Biol. Inorg. Chem. 2007, 13, 435–447. [Google Scholar] [CrossRef] [PubMed]
  254. Balakrishnan, A.; Menon, V.P. Antioxidant properties of hesperidin in nicotine-induced lung toxicity. Fundam. Clin. Pharmacol. 2007, 21, 535–546. [Google Scholar] [CrossRef]
  255. Kalpana, K.B.; Srinivasan, M.; Menon, V.P. Evaluation of antioxidant activity of hesperidin and its protective effect on H2O2 induced oxidative damage on pBR322 DNA and RBC cellular membrane. Mol. Cell. Biochem. 2008, 323, 21–29. [Google Scholar] [CrossRef]
  256. Heo, S.D.; Kim, J.; Choi, Y.; Ekanayake, P.; Ahn, M.; Shin, T. Hesperidin improves motor disability in rat spinal cord injury through anti-inflammatory and antioxidant mechanism via Nrf-2/HO-1 pathway. Neurosci. Lett. 2020, 715, 134619. [Google Scholar] [CrossRef]
  257. Wang, Z.H.; Kang, K.A.; Zhang, R.; Piao, M.J.; Jo, S.H.; Kim, J.S.; Kang, S.S.; Lee, J.S.; Park, D.H.; Hyun, J.W. Myricetin suppresses oxidative stress-induced cell damage via both direct and indirect antioxidant action. Environ. Toxicol. Pharmacol. 2010, 29, 12–18. [Google Scholar] [CrossRef]
  258. Mendes, R.A.; Almeida, S.K.C.; Soares, I.N.; Barboza, C.A.; Freitas, R.G.; Brown, A.; de Souza, G.L.C. A computational investigation on the antioxidant potential of myricetin 3,4′-di-O-α-L-rhamnopyranoside. J. Mol. Model. 2018, 24, 133. [Google Scholar] [CrossRef]
  259. Gupta, G.; Siddiqui, M.A.; Khan, M.M.; Ajmal, M.; Ahsan, R.; Rahaman, A.; Ahmad, A.; Arshad; Khushtar, M. Current pharmacological trends on myricetin. Drug Res. 2020, 70, 448–454. [Google Scholar] [CrossRef]
  260. Akhtar, S.; Najafzadeh, M.; Isreb, M.; Newton, L.; Gopalan, R.C.; Anderson, D. ROS-induced oxidative damage in lymphocytes ex vivo/in vitro from healthy individuals and MGUS patients: Protection by myricetin bulk and nanoforms. Arch. Toxicol. 2020, 94, 1229–1239. [Google Scholar] [CrossRef]
  261. Liu, P.; Deng, T.; Hou, X.; Wang, J. Antioxidant properties of isolated isorhamnetin from the sea buckthorn marc. Plant Foods Hum. Nutr. 2009, 64, 141–145. [Google Scholar] [CrossRef]
  262. Ishola, I.O.; Osele, M.O.; Chijioke, M.C.; Adeyemi, O.O. Isorhamnetin enhanced cortico-hippocampal learning and memory capability in mice with scopolamine-induced amnesia: Role of antioxidant defense, cholinergic and BDNF signaling. Brain Res. 2019, 1712, 188–196. [Google Scholar] [CrossRef]
  263. Yang, J.H.; Shin, B.Y.; Han, J.Y.; Kim, M.G.; Wi, J.E.; Kim, Y.W.; Cho, I.J.; Kim, S.C.; Shin, S.M.; Ki, S.H. Isorhamnetin protects against oxidative stress by activating Nrf2 and inducing the expression of its target genes. Toxicol. Appl. Pharmacol. 2014, 274, 293–301. [Google Scholar] [CrossRef]
  264. Xu, S.L.; Choi, R.C.Y.; Zhu, K.Y.; Leung, K.-W.; Guo, A.J.Y.; Bi, D.; Xu, H.; Lau, D.T.W.; Dong, T.T.X.; Tsim, K.W.K. Isorhamnetin, a flavonol aglycone from Ginkgo biloba L., induces neuronal differentiation of cultured PC12 cells: Potentiating the effect of nerve growth factor. Evid.-Based Complement. Altern. Med. 2012, 2012, 278273. [Google Scholar] [CrossRef]
  265. Iida, A.; Usui, T.; Zar Kalai, F.; Han, J.; Isoda, H.; Nagumo, Y. Protective effects of Nitraria retusa extract and its constituent iso-rhamnetin against amyloid β-induced cytotoxicity and amyloid β aggregation. Biosci. Biotechnol. Biochem. 2015, 79, 1548–1551. [Google Scholar] [CrossRef] [PubMed]
  266. Choi, Y.H. The cytoprotective effect of isorhamnetin against oxidative stress is mediated by the upregulation of the Nrf2-dependent HO-1 expression in C2C12 myoblasts through scavenging reactive oxygen species and ERK inactivation. Gen. Physiol. Biophys. 2016, 35, 145–154. [Google Scholar] [CrossRef] [PubMed]
  267. Chen, Q.; Song, S.; Wang, Z.; Shen, Y.; Xie, L.; Li, J.; Chen, Q. Isorhamnetin induces the paraptotic cell death through ROS and the ERK/MAPK pathway in OSCC cells. Oral Dis. 2021, 27, 240–250. [Google Scholar] [CrossRef] [PubMed]
  268. Cavia-Saiz, M.; Busto, M.D.; Pilar-Izquierdo, M.C.; Ortega, N.; Perez-Mateos, M.; Muñiz, P. Antioxidant properties, radical scavenging activity and biomolecule protection capacity of flavonoid naringenin and its glycoside naringin: A comparative study. J. Sci. Food Agric. 2010, 90, 1238–1244. [Google Scholar] [CrossRef] [PubMed]
  269. Yáñez, J.A.; Davies, N.M. Stereospecific high-performance liquid chromatographic analysis of naringenin in urine. J. Pharm. Biomed. Anal. 2005, 39, 164–169. [Google Scholar] [CrossRef] [PubMed]
  270. Patel, K.; Singh, G.K.; Patel, D.K. A review on pharmacological and analytical aspects of naringenin. Chin. J. Integr. Med. 2014, 24, 551–560. [Google Scholar] [CrossRef] [PubMed]
  271. El-Saad, A.M.A.; Abdel-Wahab, W.M. Naringenin Attenuates Toxicity and Oxidative Stress Induced by Lambda-cyhalothrin in Liver of Male Rats. Pak. J. Biol. Sci. PJBS 2020, 23, 510–517. [Google Scholar] [CrossRef] [PubMed]
  272. Singh, P.; Bansal, S.; Kuhad, A.; Kumar, A.; Chopra, K. Naringenin ameliorates diabetic neuropathic pain by modulation of oxidative-nitrosative stress, cytokines and MMP-9 levels. Food Funct. 2020, 11, 4548–4560. [Google Scholar] [CrossRef] [PubMed]
  273. Heo, H.J.; Kim, D.O.; Shin, S.C.; Kim, M.J.; Kim, B.G.; Shin, D.H. Effect of antioxidant flavanone, naringenin, from Citrus junos on neuroprotection. J. Agric. Food Chem. 2004, 52, 1520–1525. [Google Scholar] [CrossRef]
  274. Lee, M.-K.; Bok, S.-H.; Jeong, T.-S.; Moon, S.-S.; Lee, S.-E.; Park, Y.B.; Choi, M.-S. Supplementation of naringenin and its synthetic derivative alters antioxidant enzyme activities of erythrocyte and liver in high cholesterol-fed rats. Bioorg. Med. Chem. 2002, 10, 2239–2244. [Google Scholar] [CrossRef]
  275. Türkkan, B.; Özyürek, M.; Bener, M.; Güçlü, K.; Apak, R. Synthesis, characterization and antioxidant capacity of naringenin-oxime. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2012, 85, 235–240. [Google Scholar] [CrossRef] [PubMed]
  276. Liaquat, L.; Batool, Z.; Sadir, S.; Rafiq, S.; Shahzad, S.; Perveen, T.; Haider, S. Naringenin-induced enhanced antioxidant defence system meliorates cholinergic neurotransmission and consolidates memory in male rats. Life Sci. 2018, 194, 213–223. [Google Scholar] [CrossRef]
  277. Li, T.R.; Yang, Z.Y.; Wang, B.D.; Qin, D.D. Synthesis, characterization, antioxidant activity and DNA-binding studies of two rare earth (III) complexes with naringenin-2-hydroxy benzoyl hydrazone ligand. Eur. J. Med. Chem. 2008, 43, 1688–1695. [Google Scholar] [CrossRef]
  278. Rehman, K.; Khan, I.I.; Akash, M.S.H.; Jabeen, K.; Haider, K. Naringenin downregulates inflammation-mediated nitric oxide overproduction and potentiates endogenous antioxidant status during hyperglycemia. J. Food Biochem. 2020, 44, e13422. [Google Scholar] [CrossRef]
  279. Wali, A.F.; Rashid, S.; Rashid, S.M.; Ansari, M.A.; Khan, M.R.; Haq, N.; Alhareth, D.Y.; Ahmad, A.; Rehman, M.U. Naringenin regulates doxorubicin-induced liver dysfunction: Impact on oxidative stress and inflammation. Plants 2020, 9, 550. [Google Scholar] [CrossRef] [PubMed]
  280. Tseng, Y.T.; Hsu, H.T.; Lee, T.Y.; Chang, W.H.; Lo, Y.C. Naringenin, a dietary flavanone, enhances insulin-like growth factor 1 receptor-mediated antioxidant defense and attenuates methylglyoxal-induced neurite damage and apoptotic death. Nutr. Neurosci. 2021, 24, 71–81. [Google Scholar] [CrossRef] [PubMed]
  281. Zhou, X.; Wang, F.; Zhou, R.; Song, X.; Xie, M. Apigenin: A current review on its beneficial biological activities. J. Food Biochem. 2017, 41, e12376. [Google Scholar] [CrossRef]
  282. Singh, J.P.V.; Selvendiran, K.; Banu, S.M.; Padmavathi, R.; Sakthisekaran, D. Protective role of Apigenin on the status of lipid peroxidation and antioxidant defense against hepatocarcinogenesis in Wistar albino rats. Phytomedicine 2004, 11, 309–314. [Google Scholar] [CrossRef] [PubMed]
  283. Darabi, P.; Khazali, H.; Natanzi, M.M. Therapeutic potentials of the natural plant flavonoid apigenin in polycystic ovary syndrome in rat model: Via modulation of pro-inflammatory cytokines and antioxidant activity. Gynecol. Endocrinol. 2019, 36, 582–587. [Google Scholar] [CrossRef]
  284. Goudarzi, M.; Kalantar, M.; Sadeghi, E.; Karamallah, M.H.; Kalantar, H. Protective effects of apigenin on altered lipid peroxidation, inflammation, and antioxidant factors in methotrexate-induced hepatotoxicity. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2020, 394, 523–531. [Google Scholar] [CrossRef]
  285. Wang, N.; Yi, W.J.; Tan, L.; Zhang, J.H.; Xu, J.; Chen, Y.; Qin, M.; Yu, S.; Guan, J.; Zhang, R. Apigenin attenuates streptozotocin-induced pancreatic β cell damage by its protective effects on cellular antioxidant defense. In Vitro Cell. Dev. Biol. Anim. 2017, 53, 554–563. [Google Scholar] [CrossRef] [PubMed]
  286. Ashokkumar, P.; Sudhandiran, G. Protective role of luteolin on the status of lipid peroxidation and antioxidant defense against azoxymethane-induced experimental colon carcinogenesis. Biomed. Pharmacother. 2008, 62, 590–597. [Google Scholar] [CrossRef] [PubMed]
  287. Khan, N.; Syed, D.N.; Ahmad, N.; Mukhtar, H. Fisetin: A dietary antioxidant for health promotion. Antioxid. Redox Signal. 2013, 19, 151–162. [Google Scholar] [CrossRef]
  288. Park, B.-S.; Choi, N.-E.; Lee, J.H.; Kang, H.-M.; Yu, S.-B.; Kim, H.-J.; Kang, H.-K.; Kim, I.-R. Crosstalk between fisetin-induced apoptosis and autophagy in human oral squamous cell carcinoma. J. Cancer 2019, 10, 138–146. [Google Scholar] [CrossRef] [PubMed]
  289. Naeimi, A.F.; Alizadeh, M. Antioxidant properties of the flavonoid fisetin: An updated review of in vivo and in vitro studies. Trends Food Sci. Technol. 2017, 70, 34–44. [Google Scholar] [CrossRef]
  290. Sahu, B.D.; Kalvala, A.K.; Koneru, M.; Mahesh Kumar, J.; Kuncha, M.; Rachamalla, S.S.; Sistla, R. Ameliorative effect of fisetin on cisplatin-induced nephrotoxicity in rats via modulation of NF-κB activation and antioxidant defence. PLoS ONE. 2014, 9, e105070. [Google Scholar] [CrossRef]
  291. Alikatte, K.; Palle, S.; Rajendra Kumar, J.; Pathakala, N. Fisetin improved rotenone-induced behavioral deficits, oxidative changes, and mitochondrial dysfunctions in rat model of Parkinson’s disease. J. Diet. Suppl. 2021, 18, 57–71. [Google Scholar] [CrossRef] [PubMed]
  292. Maher, P. Modulation of the neuroprotective and anti-inflammatory activities of the flavonolfisetin by the transition metals iron and copper. Antioxidants 2020, 9, 1113. [Google Scholar] [CrossRef]
  293. Olaso, E.; Badiola, I.; Villacé, P.; Basaldua, I. Downregulation of discoidin domain receptor 2 in A375 human melanoma cells reduces its experimental liver metastasis ability. Oncol. Rep. 2011, 26, 971–978. [Google Scholar] [CrossRef]
  294. Ahmad, S.; Khan, A.; Ali, W.; Jo, M.H.; Park, J.; Ikram, M.; Kim, M.O. Fisetin rescues the mice brains against D-galactose-induced oxidative stress, neuroinflammation and memory impairment. Front. Pharmacol. 2021, 12, 612078. [Google Scholar] [CrossRef]
  295. Pei, F.; Pei, H.; Su, C.; Du, L.; Wang, J.; Xie, F.; Yin, Q.; Gao, Z. Fisetin alleviates neointimal hyperplasia via PPARγ/PON2 antioxidative pathway in shr rat artery injury model. Oxid. Med. Cell. Longev. 2021, 2021, 6625517. [Google Scholar] [CrossRef] [PubMed]
  296. Kubina, R.; Iriti, M.; Kabała-Dzik, A. Anticancer potential of selected flavonols: Fisetin, kaempferol, and quercetin on head and neck cancers. Nutrients 2021, 13, 845. [Google Scholar] [CrossRef] [PubMed]
  297. Sim, H.; Choo, S.; Kim, J.; Baek, M.-C.; Bae, J.-S. Fisetin suppresses pulmonary inflammatory responses through heme oxygenase-1 mediated downregulation of inducible nitric oxide synthase. J. Med. Food 2020, 23, 1163–1168. [Google Scholar] [CrossRef] [PubMed]
  298. Shanmugam, K.; Boovarahan, S.R.; Prem, P.; Sivakumar, B.; Kurian, G.A. Fisetin attenuates myocardial ischemia-reperfusion injury by activating the reperfusion injury salvage kinase (RISK) signaling pathway. Front. Pharmacol. 2021, 12, 566470. [Google Scholar] [CrossRef] [PubMed]
  299. Yao, X.; Li, L.; Kandhare, A.D.; Mukherjee-Kandhare, A.A.; Bodhankar, S.L. Attenuation of reserpine induced fibromyalgia via ROS and serotonergic pathway modulation by fisetin, a plant flavonoid polyphenol. Exp. Ther. Med. 2020, 19, 1343–1355. [Google Scholar] [CrossRef] [PubMed]
  300. Kumar, R.; Kumar, R.; Khurana, N.; Singh, S.K.; Khurana, S.; Verma, S.; Sharma, N.; Kapoor, B.; Vyas, M.; Khursheed, R.; et al. Enhanced oral bioavailability and neuroprotective effect of fisetin through its SNEDDS against rotenone-induced Parkinson’s disease rat model. Food Chem. Toxicol. 2020, 144, 111590. [Google Scholar] [CrossRef]
  301. Pop, R.; Ştefănut, M.N.; Căta, A.; Tănasie, C.; Medeleanu, M. Ab initio study regarding the evaluation of the antioxidant character of cyanidin, delphinidin and malvidin. Open Chem. 2012, 10, 180–186. [Google Scholar] [CrossRef]
  302. Bognar, E.; Sarszegi, Z.; Szabo, A.; Debreceni, B.; Kalman, N.; Tucsek, Z.; Sumegi, B.; Gallyas, F. Antioxidant and anti-inflammatory effects in RAW264.7 macrophages of malvidin, a major red wine polyphenol. PLoS ONE 2013, 8, e65355. [Google Scholar] [CrossRef]
  303. Bastin, A.; Sadeghi, A.; Nematollahi, M.H.; Abolhassani, M.; Mohammadi, A.; Akbari, H. The effects of malvidin on oxidative stress parameters and inflammatory cytokines in LPS-induced human THP-1 cells. J. Cell. Physiol. 2021, 236, 2790–2799. [Google Scholar] [CrossRef]
  304. Xu, J.; Zhang, Y.; Ren, G.; Yang, R.; Chen, J.; Xiang, X.; Qin, H.; Chen, J. Inhibitory effect of delphinidin on oxidative stress induced by H2O2 in HepG2 cells. Oxid. Med. Cell. Longev. 2020, 2020, 4694760. [Google Scholar] [CrossRef]
  305. Rodríguez-Bonilla, P.; Gandía-Herrero, F.; Matencio, A.; García-Carmona, F.; López-Nicolás, J.M. Comparative study of the anti-oxidant capacity of four stilbenes using ORAC, ABTS+, and FRAP techniques. Food Anal. Methods 2017, 10, 2994–3000. [Google Scholar] [CrossRef]
  306. Gülçin, I. Antioxidant properties of resveratrol: A structure–activity insight. Innov. Food Sci. Emerg. Technol. 2010, 11, 210–218. [Google Scholar] [CrossRef]
  307. de la Lastra, C.A.; Villegas, I. Resveratrol as an antioxidant and pro-oxidant agent: Mechanisms and clinical implications. Biochem. Soc. Trans. 2007, 35, 1156–1160. [Google Scholar] [CrossRef]
  308. Carrizzo, A.; Forte, M.; Damato, A.; Trimarco, V.; Salzano, F.; Bartolo, M.; Maciag, A.; Puca, A.A.; Vecchione, C. Antioxidant effects of resveratrol in cardiovascular, cerebral and metabolic diseases. Food Chem. Toxicol. 2013, 61, 215–226. [Google Scholar] [CrossRef] [PubMed]
  309. Olas, B.; Wachowicz, B. Resveratrol and vitamin C as antioxidants in blood platelets. Thromb. Res. 2002, 106, 143–148. [Google Scholar] [CrossRef] [PubMed]
  310. Frankel, E.; Waterhouse, A.; Kinsella, J. Inhibition of human LDL oxidation by resveratrol. Lancet 1993, 341, 1103–1104. [Google Scholar] [CrossRef] [PubMed]
  311. Franco, J.G.; Lisboa, P.C.; Lima, N.S.; Amaral, T.A.; Peixoto-Silva, N.; Resende, A.C.; Oliveira, E.; Passos, M.C.; Moura, E.G. Resveratrol attenuates oxidative stress and prevents steatosis and hypertension in obese rats programmed by early weaning. J. Nutr. Biochem. 2013, 24, 960–966. [Google Scholar] [CrossRef]
  312. Kim, Y.; Lim, S.Y.; Rhee, S.H.; Park, K.Y.; Kim, C.H.; Choi, B.T.; Choi, Y.H. Resveratrol inhibits inducible nitric oxide synthase and cyclooxygenase-2 expression in β-amyloid-treated C6 glioma cells. Int. J. Mol. Med. 2006, 17, 1069–1675. [Google Scholar] [CrossRef] [PubMed]
  313. Inoue, H.; Jiang, X.-F.; Katayama, T.; Osada, S.; Umesono, K.; Namura, S. Brain protection by resveratrol and fenofibrate against stroke requires peroxisome proliferator-activated receptor α in mice. Neurosci. Lett. 2003, 352, 203–206. [Google Scholar] [CrossRef] [PubMed]
  314. Kelsey, N.A.; Wilkins, H.M.; Linseman, D.A. Nutraceutical antioxidants as novel neuroprotective agents. Molecules 2010, 15, 7792–7814. [Google Scholar] [CrossRef]
  315. Lee, S.-M.; Yang, H.; Tartar, D.M.; Gao, B.; Luo, X.; Ye, S.Q.; Zaghouani, H.; Fang, D. Prevention and treatment of diabetes with resveratrol in a non-obese mouse model of type 1 diabetes. Diabetologia 2011, 54, 1136–1146. [Google Scholar] [CrossRef] [PubMed]
  316. Ko, J.-H.; Sethi, G.; Um, J.-Y.; Shanmugam, M.K.; Arfuso, F.; Kumar, A.P.; Bishayee, A.; Ahn, K.S. The role of resveratrol in cancer therapy. Int. J. Mol. Sci. 2017, 18, 2589. [Google Scholar] [CrossRef] [PubMed]
  317. Likhitwitayawuid, K. Oxyresveratrol: Sources, productions, biological activities, pharmacokinetics, and delivery systems. Molecules 2021, 26, 4212. [Google Scholar] [CrossRef] [PubMed]
  318. Wang, L.; Zhao, H.; Wang, L.; Tao, Y.; Du, G.; Guan, W.; Liu, J.; Brennan, C.; Ho, C.-T.; Li, S. Effects of selected resveratrol analogues on activation and polarization of lipopolysaccharide-stimulated BV-2 microglial cells. J. Agric. Food Chem. 2020, 68, 3750–3757. [Google Scholar] [CrossRef] [PubMed]
  319. Ban, J.Y.; Jeon, S.Y.; Nguyen, T.T.H.; Bae, K.; Song, K.S.; Seonga, Y.H. Neuroprotective effect of oxyresveratrol from Smilacis chinae rhizome on amyloid β protein (25–35)-induced neurotoxicity in cultured rat cortical neurons. Biol. Pharm. Bull. 2006, 29, 2419–2424. [Google Scholar] [CrossRef] [PubMed]
  320. Ferlinahayati; Syah, Y.M.; Juliawaty, L.D.; Achmad, S.A.; Hakim, E.H.; Takayama, H.; Said, I.M.; Latip, J. Phenolic constituents from the wood of morus australis with cytotoxic activity. Z. Naturforschung C 2008, 63, 35–39. [Google Scholar] [CrossRef]
  321. He, W.; Rui, J.; Wang, Q.; Chen, Z. Antioxidant activity of piceatannol in canola oil. Eur. J. Lipid Sci. Technol. 2021, 123, 2000398. [Google Scholar] [CrossRef]
  322. Zhou, Y.; Khan, H.; Hoi, M.P.M.; Cheang, W.S. Piceatannol Protects Brain Endothelial Cell Line (bEnd. 3) against Lipopolysaccharide-Induced Inflammation and Oxidative Stress. Molecules 2022, 27, 1206. [Google Scholar] [CrossRef]
  323. Tang, Y.; Chan, S. A review of the pharmacological effects of piceatannol on cardiovascular diseases. Phytother. Res. 2014, 28, 1581–1588. [Google Scholar] [CrossRef]
  324. Koskela, A.; Reinisalo, M.; Hyttinen, J.M.T.; Kaarniranta, K.; Karjalainen, R.O. Pinosylvin-mediated protection against oxidative stress in human retinal pigment epithelial cells. Mol. Vis. 2014, 20, 760–769. [Google Scholar]
  325. Jančinová, V.; Perečko, T.; Nosáľ, R.; Harmatha, J.; Drábiková, K. The natural stilbenoidpinosylvin and activated neutrophils: Effects on oxidative burst, protein kinase C, apoptosis and efficiency in adjuvant arthritis. Acta Pharmacol. Sin. 2012, 33, 1285–1292. [Google Scholar] [CrossRef]
  326. Bauerova, K.; Acquaviva, A.; Ponist, S.; Gardi, C.; Vecchio, D.; Drafi, F.; Arezzini, B.; Bezakova, L.; Kuncirova, V.; Mihalova, D.; et al. Markers of inflammation and oxidative stress studied in adjuvant-induced arthritis in the rat on systemic and local level affected by pinosylvin and methotrexate and their combi-nation. Autoimmunity 2015, 48, 46–56. [Google Scholar] [CrossRef]
  327. Kolodziejczyk-Czepas, J.; Czepas, J. Rhaponticin as an anti-inflammatory component of rhubarb: A minireview of the current state of the art and prospects for future research. Phytochem. Rev. 2019, 18, 1375–1386. [Google Scholar] [CrossRef]
  328. Kawakami, S.; Kinoshita, Y.; Maruki-Uchida, H.; Yanae, K.; Sai, M.; Ito, T. Piceatannol and its metabolite, isorhapontigenin, induce SIRT1 expression in THP-1 human monocytic cell line. Nutrients 2014, 6, 4794–4804. [Google Scholar] [CrossRef]
  329. Kauppinen, A.; Suuronen, T.; Ojala, J.; Kaarniranta, K.; Salminen, A. Antagonistic crosstalk between NF-κB and SIRT1 in the regulation of inflammation and metabolic disorders. Cell. Signal. 2013, 25, 1939–1948. [Google Scholar] [CrossRef]
  330. Zanwar, A.A.; Badole, S.L.; Shende, P.S.; Hegde, M.V.; Bodhankar, S.L. Antioxidant role of catechin in health and disease. In Polyphenols in Human Health and Disease; Watson, R.R., Preedy, V.R., Eds.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 267–271. [Google Scholar]
  331. Ahmadi, S.M.; Farhoosh, R.; Sharif, A.; Rezaie, M. Structure-antioxidant activity relationships of luteolin and catechin. J. Food Sci. 2020, 85, 298–305. [Google Scholar] [CrossRef]
  332. Chen, W.; Zhu, X.; Lu, Q.; Zhang, L.; Wang, X.; Liu, R. C-ring cleavage metabolites of catechin and epicatechin enhanced antioxidant activities through intestinal microbiota. Food Res. Int. 2020, 135, 109271. [Google Scholar] [CrossRef]
  333. Wang, L.; Yang, Q.; Li, Y.; Wang, S.; Yang, F.; Zhao, X. How the functional group substitution and solvent effects affect the antioxidant activity of (+)-catechin? J. Mol. Liq. 2021, 327, 114818. [Google Scholar] [CrossRef]
  334. Taslimi, P.; Kocyigit, U.M.; Tüzün, B.; Kirici, M. Biological effects and molecular docking studies of Catechin 5-O-gallate: Antioxidant, anticholinergics, antiepileptic and antidiabetic potentials. J. Biomol. Struct. Dyn. 2022, 40, 2489–2497. [Google Scholar] [CrossRef]
  335. Pushp, P.; Sharma, N.; Joseph, G.S.; Singh, R.P. Antioxidant activity and detection of (−)epicatechin in the methanolic extract of stem of Tinospora cordifolia. J. Food Sci. Technol. 2011, 50, 567–572. [Google Scholar] [CrossRef]
  336. Jug, U.; Naumoska, K.; Vovk, I. (−)-Epicatechin—An Important contributor to the antioxidant activity of Japanese knotweed rhizome bark extract as determined by antioxidant activity-guided fractionation. Antioxidants 2021, 10, 133. [Google Scholar] [CrossRef] [PubMed]
  337. Shi, X.; Ye, J.; Leonard, S.; Ding, M.; Vallyathan, V.; Castranova, V.; Rojanasakul, Y.; Dong, Z. Antioxidant properties of (-)-epicatechin-3-gallate and its inhibition of Cr (VI)-induced DNA damage and Cr (IV)- or TPA-stimulated NF-κB activation. Mol. Cell. Biochem. 2000, 206, 125–132. [Google Scholar] [CrossRef] [PubMed]
  338. Minnelli, C.; Galeazzi, R.; Laudadio, E.; Amici, A.; Rusciano, D.; Armeni, T.; Cantarini, M.; Stipa, P.; Mobbili, G. Monoalkylated epigallocatechin-3-gallate (C18-EGCG) as novel lipophilic EGCG derivative: Characterization and antioxidant evaluation. Antioxidants 2020, 9, 208. [Google Scholar] [CrossRef] [PubMed]
  339. Ding, Y.; Ren, D.; Xu, H.; Liu, W.; Liu, T.; Li, L.; Li, J.; Li, Y.; Wen, A. Antioxidant and pro-angiogenic effects of corilagin in rat cerebral ischemia via Nrf2 activation. Oncotarget 2017, 8, 114816–114828. [Google Scholar] [CrossRef] [PubMed]
  340. Yuan, H.; Zhang, J.; Yin, X.; Liu, T.; Yue, X.; Li, C.; Wang, Y.; Li, D.; Wang, Q. The protective role of corilagin on renal calcium oxalate crystal-induced oxidative stress, inflammatory response, and apoptosis via PPAR-γ and PI3K/Akt pathway in rats. Biotechnol. Appl. Biochem. 2021, 68, 1323–1331. [Google Scholar] [CrossRef] [PubMed]
  341. Chen, Y.; Chen, C. Corilagin prevents tert-butyl hydroperoxide-induced oxidative stress injury in cultured N9 murine microglia cells. Neurochem. Int. 2011, 59, 290–296. [Google Scholar] [CrossRef] [PubMed]
  342. Kinoshita, S.; Inoue, Y.; Nakama, S.; Ichiba, T.; Aniya, Y. Antioxidant and hepatoprotective actions of medicinal herb, Terminalia catappa L. from Okinawa Island and its tannin corilagin. Phytomedicine 2007, 14, 755–762. [Google Scholar] [CrossRef] [PubMed]
  343. Wang, W.; Yang, L.; Liu, T.; Ma, Y.; Huang, S.; He, M.; Wang, J.; Wen, A.; Ding, Y. Corilagin ameliorates sleep deprivation-induced memory impairments by inhibiting NOX2 and activating Nrf2. Brain Res. Bull. 2020, 160, 141–149. [Google Scholar] [CrossRef] [PubMed]
  344. Liu, F.C.; Yu, H.P.; Chou, A.H.; Lee, H.C.; Liao, C.C. Corilagin reduces acetaminophen-induced hepatotoxicity through MAPK and NF-κB signaling pathway in a mouse model. Am. J. Transl. Res. 2020, 12, 5597–5607. [Google Scholar]
  345. Kitts, D.D.; Yuan, Y.V.; Wijewickreme, A.N.; Thompson, L.U. Antioxidant activity of the flaxseed lignan secoisolariciresinol diglycoside and its mammalian lignan metabolites enterodiol and enterolactone. Mol. Cell. Biochem. 1999, 202, 91–100. [Google Scholar] [CrossRef]
  346. Tannous, S.; Haykal, T.; Dhaini, J.; Hodroj, M.H.; Rizk, S. The anti-cancer effect of flaxseed lignan derivatives on different acute myeloid leukemia cancer cells. Biomed. Pharmacother. 2020, 132, 110884. [Google Scholar] [CrossRef]
  347. Fauré, M.; Lissi, E.; Torres, R.; Videla, L.A. Antioxidant activities of lignans and flavonoids. Phytochemistry 1990, 29, 3773–3775. [Google Scholar] [CrossRef]
  348. Choi, Y.W.; Takamatsu, S.; Khan, S.I.; Srinivas, P.V.; Ferreira, D.; Zhao, J.; Khan, I.A. Schisandrene, a dibenzocyclooctadiene lignan from Schisandra chinensis: Structure−antioxidant activity relationships of dibenzocyclooctadiene lignans. J. Nat. Prod. 2006, 69, 356–359. [Google Scholar] [CrossRef]
  349. De Silva, S.F.; Alcorn, J. Flaxseed lignans as important dietary polyphenols for cancer prevention and treatment: Chemistry, pharmacokinetics, and molecular targets. Pharmaceuticals 2019, 12, 68. [Google Scholar] [CrossRef]
  350. Baderschneider, B.; Winterhalter, P. Isolation and characterization of novel benzoates, cinnamates, flavonoids, and lignans from riesling wine and screening for antioxidant activity. J. Agric. Food Chem. 2001, 49, 2788–2798. [Google Scholar] [CrossRef]
  351. Ghafoorunissa; Hemalatha, S.; Rao, M.V.V. Sesame lignans enhance antioxidant activity of vitamin E in lipid peroxidation systems. Mol. Cell. Biochem. 2004, 262, 195–202. [Google Scholar] [CrossRef]
  352. Mazimba, O.; Majinda, R.R.; Modibedi, C.; Masesane, I.B.; Cencič, A.; Chingwaru, W. Tylosema esculentum extractives and their bioactivity. Bioorg. Med. Chem. 2011, 19, 5225–5230. [Google Scholar] [CrossRef]
  353. Singh, R.; Singh, B.; Singh, S.; Kumar, N.; Kumar, S.; Arora, S. Umbelliferone—An antioxidant isolated from Acacia nilotica (L.) Willd. ex. Del. Food Chem. 2010, 120, 825–830. [Google Scholar] [CrossRef]
  354. Kanimozhi, G.; Prasad, N.R.; Ramachandran, S.; Pugalendi, K.V. Umbelliferone modulates gamma-radiation induced reactive oxygen species generation and subsequent oxidative damage in human blood lymphocytes. Eur. J. Pharmacol. 2011, 672, 20–29. [Google Scholar] [CrossRef] [PubMed]
  355. Kanimozhi, G.; Prasad, N.R.; Ramachandran, S.; Pugalendi, K. Umbelliferone protects whole-body irradiated Swiss albino mice: Study on animal survival, tissue antioxidant status and DNA damage. Biomed. Prev. Nutr. 2012, 2, 186–192. [Google Scholar] [CrossRef]
  356. Rauf, A.; Khan, R.; Khan, H.; Pervez, S.; Pirzada, A.S. In vivo antinociceptive and anti-inflammatory activities of umbelliferone isolated from Potentilla evestita. Nat. Prod. Res. 2014, 28, 1371–1374. [Google Scholar] [CrossRef]
  357. Ramu, R.; Shirahatti, P.S.; Zameer, F.; Ranganatha, L.V.; Prasad, M.N. Inhibitory effect of banana (Musa sp. var Nanjangud rasa bale) flower extract and its constituents Umbelliferone and Lupeol on α-glucosidase, aldose reductase and glycation at multiple stages. S. Afr. J. Bot. 2014, 95, 54–63. [Google Scholar]
  358. Gao, D.; Zhang, Y.L.; Xu, P.; Lin, Y.X.; Yang, F.Q.; Liu, J.H.; Zhu, H.W.; Xia, Z.N. In vitro evaluation of dual agonists for PPARγ/β from the flower of Edgeworthiagardneri (wall.) Meisn. J. Ethnopharmacol. 2015, 162, 14–19. [Google Scholar] [CrossRef]
  359. Ramesh, B.; Pugalendi, K. Antihyperglycemic effect of umbelliferone in streptozotocin-diabetic rats. J. Med. Food 2006, 9, 562–566. [Google Scholar] [CrossRef]
  360. Yu, S.-M.; Hu, D.-H.; Zhang, J.-J. Umbelliferone exhibits anticancer activity via the induction of apoptosis and cell cycle arrest in HepG2 hepatocellular carcinoma cells. Mol. Med. Rep. 2015, 12, 3869–3873. [Google Scholar] [CrossRef]
  361. Hong, S.; Kim, M.-M. Aesculetin inhibits cell invasion through inhibition of MMP-9 activity and antioxidant activity. J. Life Sci. 2016, 26, 673–679. [Google Scholar] [CrossRef]
  362. Thuong, P.T.; Hung, T.M.; Ngoc, T.M.; Ha, D.T.; Min, B.S.; Kwack, S.J.; Kang, T.S.; Choi, J.S.; Bae, K. Antioxidant activities of coumarins from Korean medicinal plants and their structure–activity relationships. Phytother. Res. 2010, 24, 101–106. [Google Scholar] [CrossRef]
  363. Dong, Y.; Hou, Q.; Sun, M.; Sun, J.; Zhang, B. Targeted isolation of antioxidant constituents from Plantago asiatica L. and in vitro activity assay. Molecules 2020, 25, 1825. [Google Scholar] [CrossRef]
  364. Jeon, H.Y.; Seo, D.B.; Shin, H.-J.; Lee, S.-J. Effect of Aspergillus oryzae-challenged germination on soybean isoflavone content and antioxidant activity. J. Agric. Food Chem. 2012, 60, 2807–2814. [Google Scholar] [CrossRef] [PubMed]
  365. Jantaratnotai, N.; Utaisincharoen, P.; Sanvarinda, P.; Thampithak, A.; Sanvarinda, Y. Phytoestrogens mediated anti-inflammatory effect through suppression of IRF-1 and pSTAT1 expressions in lipopolysaccharide-activated microglia. Int. Immunopharmacol. 2013, 17, 483–488. [Google Scholar] [CrossRef] [PubMed]
  366. Yuk, H.J.; Lee, J.W.; Park, H.A.; Kwon, O.K.; Seo, K.H.; Ahn, K.S.; Oh, S.R.; Ryu, H.W. Protective effects of coumestrol on lipopolysaccharide-induced acute lung injury via the inhibition of proinflammatory mediators and NF-κB activation. J. Funct. Foods 2017, 34, 181–188. [Google Scholar] [CrossRef]
  367. Castro, C.C.; Pagnussat, A.S.; Orlandi, L.; Worm, P.; Moura, N.; Etgen, A.M.; Netto, C.A. Coumestrol has neuroprotective effects before and after global cerebral ischemia in female rats. Brain Res. 2012, 1474, 82–90. [Google Scholar] [CrossRef] [PubMed]
  368. Yuk, H.J.; Lee, J.H.; Curtis-Long, M.J.; Lee, J.W.; Kim, Y.S.; Ryu, H.W.; Park, C.G.; Jeong, T.S.; Park, K.H. The most abundant polyphenol of soy leaves, coumestrol, displays potent α-glucosidase inhibitory activity. Food Chem. 2011, 126, 1057–1063. [Google Scholar] [CrossRef]
  369. Souri, E.; Farsam, H.; Sarkheil, P.; Ebadi, F. Antioxidant activity of some furanocoumarins isolated from Heracleum persicum. Pharm. Biol. 2004, 42, 396–399. [Google Scholar] [CrossRef]
  370. Song, L.; Chen, H.; Gao, H.; Fang, X.; Mu, H.; Yuan, Y.; Yang, Q.; Jiang, Y. Chemical quantification and antioxidant assay of four active components in Ficus hirta root using UPLC-PAD-MS fingerprinting combined with cluster analysis. Chem. Cent. J. 2013, 7, 147. [Google Scholar] [CrossRef] [PubMed]
  371. Kowalczyk, J.; Boguszewska-Czubara, A.; Skalicka-Woźniak, K.; Kurzepa, J.; Budzyńska, B. Bergapten improves scopolamine-induced memory impairment in mice via cholinergic and antioxidative mechanisms. Front. Neurosci. 2020, 14, 730. [Google Scholar] [CrossRef] [PubMed]
  372. Khan, A.U.; Ijaz, M.U.; Shah, F.A.; Khan, A.W.; Li, S. Neuroprotective Effects of Berbamine, Bergepten, and Carveol on Paclitaxel-Induced Peripheral Neuropathy. Res. Sq. 2022, preprint. [Google Scholar] [CrossRef]
  373. Adakudugu, E.A.; Ameyaw, E.O.; Obese, E.; Biney, R.P.; Henneh, I.T.; Aidoo, D.B.; Oge, E.N.; Attah, I.Y.; Obiri, D.D. Protective effect of bergapten in acetic acid-induced colitis in rats. Heliyon 2020, 6, e04710. [Google Scholar] [CrossRef]
  374. Potapenko, A.Y.; Kyagova, A.A. The application of antioxidants in investigations and optimization of photochemotherapy. Membr. Cell Biol. 1998, 12, 269–278. [Google Scholar]
  375. Ren, Y.; Song, X.; Tan, L.; Guo, C.; Wang, M.; Liu, H.; Cao, Z.; Li, Y.; Peng, C. A review of the pharmacological properties of psoralen. Front. Pharmacol. 2020, 11, 571535. [Google Scholar] [CrossRef]
  376. Wang, X.; Peng, P.; Pan, Z.; Fang, Z.; Lu, W.; Liu, X. Psoralen inhibits malignant proliferation and induces apoptosis through triggering endoplasmic reticulum stress in human SMMC7721 hepatoma cells. Biol. Res. 2019, 52, 1–13. [Google Scholar] [CrossRef]
  377. Wang, Y.; Hong, C.; Zhou, C.; Xu, D.; Qu, H.-B. Screening antitumor compounds psoralen and isopsoralen from Psoralea corylifolia L. seeds. Evid.-Based Complement. Altern. Med. 2011, 2011, 363052. [Google Scholar] [CrossRef]
  378. Deng, Z.; Hassan, S.; Rafiq, M.; Li, H.; He, Y.; Cai, Y.; Kang, X.; Liu, Z.; Yan, T. Pharmacological activity of eriodictyol: The Major natural polyphenolic flavanone. Evid.-Based Complement. Altern. Med. 2020, 2020, 6681352. [Google Scholar] [CrossRef]
  379. Lee, E.-R.; Kim, J.-H.; Kang, Y.-J.; Cho, S.-G. The anti-apoptotic and anti-oxidant effect of eriodictyol on UV-induced apoptosis in keratinocytes. Biol. Pharm. Bull. 2007, 30, 32–37. [Google Scholar] [CrossRef]
  380. Rossato, M.F.; Trevisan, G.; Walker, C.I.B.; Klafke, J.Z.; de Oliveira, A.P.; Villarinho, J.G.; Zanon, R.B.; Royes, L.F.F.; Athayde, M.L.; Gomez, M.V.; et al. Eriodictyol: A flavonoid antagonist of the TRPV1 receptor with antioxidant activity. Biochem. Pharmacol. 2011, 81, 544–551. [Google Scholar] [CrossRef]
  381. Lou, H.; Jing, X.; Ren, D.; Wei, X.; Zhang, X. Eriodictyol protects against H2O2-induced neuron-like PC12 cell death through activation of Nrf2/ARE signaling pathway. Neurochem. Int. 2012, 61, 251–257. [Google Scholar] [CrossRef]
  382. Habtemariam, S.; Dagne, E. Comparative antioxidant, prooxidant and cytotoxic activity of sigmoidin A and eriodictyol. Planta Medica 2010, 76, 589–594. [Google Scholar] [CrossRef]
  383. Lou, Z.; Wang, H.; Rao, S.; Sun, J.; Ma, C.; Li, J. p-Coumaric acid kills bacteria through dual damage mechanisms. Food Control 2012, 25, 550–554. [Google Scholar] [CrossRef]
  384. Li, C.Z.; Jin, H.H.; Sun, H.X.; Zhang, Z.Z.; Zheng, J.X.; Li, S.H.; Han, S.H. Eriodictyol attenuates cisplatin-induced kidney injury by inhibiting oxidative stress and inflammation. Eur. J. Pharmacol. 2016, 772, 124–130. [Google Scholar] [CrossRef]
  385. Roy, S.; Mallick, S.; Chakraborty, T.; Ghosh, N.; Singh, A.K.; Manna, S.; Majumdar, S. Synthesis, characterisation and antioxidant activity of luteolin–vanadium (II) complex. Food Chem. 2015, 173, 1172–1178. [Google Scholar] [CrossRef]
  386. Xu, K.; Liu, B.; Ma, Y.; Du, J.; Li, G.; Gao, H.; Zhang, Y.; Ning, Z. Physicochemical properties and antioxidant activities of luteolin-phospholipid complex. Molecules 2009, 14, 3486–3493. [Google Scholar] [CrossRef] [PubMed]
  387. Kang, K.A.; Piao, M.J.; Ryu, Y.S.; Hyun, Y.J.; Park, J.E.; Shilnikova, K.; Zhen, A.X.; Kang, H.K.; Koh, Y.S.; Jeong, Y.J.; et al. Luteolin induces apoptotic cell death via antioxidant activity in human colon cancer cells. Int. J. Oncol. 2017, 51, 1169–1178. [Google Scholar] [CrossRef] [PubMed]
  388. Olayinka, E.; Ore, A.; Olotu, O.; Ogbuji, V.; Adeyemo, O.; Ola, O. Preservation of antioxidant defense system by Morin in bicalutamide-induced rat testicular toxicity. Ife J. Sci. 2021, 23, 115–122. [Google Scholar] [CrossRef]
  389. Wu, T.-W.; Zeng, L.; Wu, J.; Fung, K. Morin hydrate is a plant-derived and antioxidant-based hepatoprotector. Life Sci. 1993, 53, PL213–PL218. [Google Scholar] [CrossRef] [PubMed]
  390. Kok, L.; Wong, Y.; Wu, T.; Chan, H.; Kwok, T.; Fung, K. Morin hydrate: A potential antioxidant in minimizing the free-radicals-mediated damage to cardiovascular cells by anti-tumor drugs. Life Sci. 2000, 67, 91–99. [Google Scholar] [CrossRef] [PubMed]
  391. Das, D.; Nath, B.C.; Phukon, P.; Dolui, S.K. Synthesis and evaluation of antioxidant and antibacterial behavior of CuO nanoparticles. Colloids Surf. B Biointerfaces 2013, 101, 430–433. [Google Scholar] [CrossRef] [PubMed]
  392. Plumb, G.W.; de Pascual-Teresa, S.; Santos-Buelga, C.; Rivas-Gonzalo, J.C.; Williamson, G. Antioxidant properties of gallocatechin and prodelphinidins from pomegranate peel. Redox Rep. 2002, 7, 41–46. [Google Scholar] [CrossRef] [PubMed]
  393. Borrás, C.; Gambini, J.; Gómez-Cabrera, M.C.; Sastre, J.; Pallardó, F.V.; Mann, G.E.; Vina, J.; Borrás, C.; Gambini, J.; Gómez-Cabrera, M.C.; et al. Genistein, a soy isoflavone, up-regulates expression of antioxidant genes: Involvement of estrogen receptors, ERK1/2, and NFκB. FASEB J. 2006, 20, 2136–2138. [Google Scholar] [CrossRef] [PubMed]
  394. Gholampour, F.; Mohammadi, Z.; Karimi, Z.; Owji, S.M. Protective effect of genistein in a rat model of ischemic acute kidney injury. Gene 2020, 753, 144789. [Google Scholar] [CrossRef] [PubMed]
  395. Poasakate, A.; Maneesai, P.; Rattanakanokchai, S.; Bunbupha, S.; Tong-Un, T.; Pakdeechote, P. Genistein prevents nitric oxide deficiency-induced cardiac dysfunction and remodeling in rats. Antioxidants 2021, 10, 237. [Google Scholar] [CrossRef]
  396. Gao, Y.; Xu, J.J.; Xu, M.L.; Shi, S.S.; Xiong, J.F. Genistein and daidzein reduced chlorpyrifos induced damage in PC12 cell. Oil Crop Sci. 2020, 5, 1–5. [Google Scholar] [CrossRef]
  397. Lu, C.; Lv, J.; Jiang, N.; Wang, H.; Huang, H.; Zhang, L.; Li, S.; Zhang, N.; Fan, B.; Liu, X.; et al. Protective effects of Genistein on the cognitive deficits induced by chronic sleep deprivation. Phytother. Res. 2020, 34, 846–858. [Google Scholar] [CrossRef] [PubMed]
  398. Salahshoor, M.R.; Jalili, C.; Rashidi, I.; Roshankhah, S.; Jalili, F. Protective Effect of Genistein on the Morphine-Induced Kidney Disorders in Male Mice. Electron. J. Gen. Med. 2020, 17, em213. [Google Scholar] [CrossRef] [PubMed]
  399. Foti, P.; Erba, D.; Riso, P.; Spadafranca, A.; Criscuoli, F.; Testolin, G. Comparison between daidzein and genistein antioxidant activity in primary and cancer lymphocytes. Arch. Biochem. Biophys. 2005, 433, 421–427. [Google Scholar] [CrossRef] [PubMed]
  400. Xie, K.; Li, Y.; Chen, D.; Yu, B.; Luo, Y.; Mao, X.; Huang, Z.; Yu, J.; Luo, J.; Zheng, P.; et al. Daidzein supplementation enhances embryo survival by improving hormones, antioxidant capacity, and metabolic profiles of amniotic fluid in sows. Food Funct. 2020, 11, 10588–10600. [Google Scholar] [CrossRef]
  401. Tomar, A.; Kaushik, S.; Khan, S.I.; Bisht, K.; Nag, T.C.; Arya, D.S.; Bhatia, J. The dietary isoflavone daidzein mitigates oxidative stress, apoptosis, and inflammation in CDDP-induced kidney injury in rats: Impact of the MAPK signaling pathway. J. Biochem. Mol. Toxicol. 2020, 34, e22431. [Google Scholar] [CrossRef] [PubMed]
  402. Li, Y.; Jiang, X.; Cai, L.; Zhang, Y.; Ding, H.; Yin, J.; Li, X. Dietary Daidzein Supplementation Improved Growth Performance and Antioxidant Properties in Weaned and Growing Pigs. Res. Sq. 2021, preprint. [Google Scholar] [CrossRef]
  403. Choi, E.J. The prooxidant, rather than antioxidant, acts of daidzein in vivo and in vitro: Daidzein suppresses glutathione metabolism. Eur. J. Pharmacol. 2006, 542, 162–169. [Google Scholar] [CrossRef] [PubMed]
  404. Yang, J.; Guo, J.; Yuan, J. In vitro antioxidant properties of rutin. LWT Food Sci. Technol. 2008, 41, 1060–1066. [Google Scholar] [CrossRef]
  405. Boyle, S.P.; Dobson, V.L.; Duthie, S.J.; Hinselwood, D.C.; Kyle, J.A.M.; Collins, A.R. Bioavailability and efficiency of rutin as an antioxidant: A human supplementation study. Eur. J. Clin. Nutr. 2000, 54, 774–782. [Google Scholar] [CrossRef]
  406. Kamalakkannan, N.; Prince, P.S.M. Antihyperglycaemic and antioxidant effect of rutin, a polyphenolic flavonoid, in streptozotocin-induced diabetic wistar rats. Basic Clin. Pharmacol. Toxicol. 2006, 98, 97–103. [Google Scholar] [CrossRef] [PubMed]
  407. Girsang, E.; Lister, I.N.E.; Ginting, C.N.; Sholihah, I.A.; Raif, M.A.; Kunardi, S.; Million, H.; Widowati, W. Antioxidant and antiaging activity of rutin and caffeic acid. Pharmaciana 2020, 10, 147–156. [Google Scholar] [CrossRef]
  408. Gęgotek, A.; Jarocka-Karpowicz, I.; Skrzydlewska, E. Cytoprotective effect of ascorbic acid and rutin against oxidative changes in the proteome of skin fibroblasts cultured in a three-dimensional system. Nutrients 2020, 12, 1074. [Google Scholar] [CrossRef] [PubMed]
  409. Aslan, A.; Hussein, Y.T.; Gok, O.; Beyaz, S.; Erman, O.; Baspinar, S. Ellagic acid ameliorates lung damage in rats via modulating antioxidant activities, inhibitory effects on inflammatory mediators and apoptosis-inducing activities. Environ. Sci. Pollut. Res. 2020, 27, 7526–7537. [Google Scholar] [CrossRef]
  410. Priyadarsini, K.I.; Khopde, S.M.; Kumar, S.S.; Mohan, H. Free radical studies of ellagic acid, a natural phenolic antioxidant. J. Agric. Food Chem. 2002, 50, 2200–2206. [Google Scholar] [CrossRef] [PubMed]
  411. Manning, B.D.; Toker, A. AKT/PKB signaling: Navigating the network. Cell 2017, 169, 381–405. [Google Scholar] [CrossRef]
  412. Duan, J.; Guan, Y.; Mu, F.; Guo, C.; Zhang, E.; Yin, Y.; Wei, G.; Zhu, Y.; Cui, J.; Cao, J.; et al. Protective effect of butin against ischemia/reperfusion-induced myocardial injury in diabetic mice: Involvement of the AMPK/GSK-3β/Nrf2 signaling pathway. Sci. Rep. 2017, 7, srep41491. [Google Scholar] [CrossRef]
  413. Kang, G.G.; Francis, N.; Hill, R.; Waters, D.; Blanchard, C.; Santhakumar, A.B. Dietary polyphenols and gene expression in molecular pathways associated with type 2 diabetes mellitus: A review. Int. J. Mol. Sci. 2019, 21, 140. [Google Scholar] [CrossRef]
  414. Fu, Y.-S.; Kang, N.; Yu, Y.; Mi, Y.; Guo, J.; Wu, J.; Weng, C.-F. Polyphenols, flavonoids and inflammasomes: The role of cigarette smoke in COPD. Eur. Respir. Rev. 2022, 31, 220028. [Google Scholar] [CrossRef]
  415. Morton, L.; Braakhuis, A.J. The Effects of Fruit-Derived Polyphenols on Cognition and Lung Function in Healthy Adults: A Systematic Review and Meta-Analysis. Nutrients 2021, 13, 4273. [Google Scholar] [CrossRef]
  416. Harris, C.; Hansen, J.M. Nrf2-Mediated Resistance to Oxidant-Induced Redox Disruption in Embryos. Birth Defects Res. Part B Dev. Reprod. Toxicol. 2012, 95, 213–218. [Google Scholar] [CrossRef] [PubMed]
  417. Hussain, T.; Tan, B.; Liu, G.; Murtaza, G.; Rahu, N.; Saleem, M.; Yin, Y. Modulatory mechanism of polyphenols and Nrf2 signaling pathway in LPS challenged pregnancy disorders. Oxid. Med. Cell. Longev. 2017, 2017, 8254289. [Google Scholar] [CrossRef] [PubMed]
  418. Guerby, P.; Tasta, O.; Swiader, A.; Pont, F.; Bujold, E.; Parant, O.; Vayssiere, C.; Salvayre, R.; Negre-Salvayre, A. Role of oxidative stress in the dysfunction of the placental endothelial nitric oxide synthase in preeclampsia. Redox Biol. 2021, 40, 101861. [Google Scholar] [CrossRef] [PubMed]
  419. Filosa, S.; Di Meo, F.; Crispi, S. Polyphenols-gut microbiota interplay and brain neuromodulation. Neural Regen. Res. 2018, 13, 2055–2059. [Google Scholar] [CrossRef] [PubMed]
  420. Liu, S.; Cheng, L.; Liu, Y.; Zhan, S.; Wu, Z.; Zhang, X. Relationship between Dietary Polyphenols and Gut Microbiota: New Clues to Improve Cognitive Disorders, Mood Disorders and Circadian Rhythms. Foods 2023, 12, 1309. [Google Scholar] [CrossRef] [PubMed]
  421. Sarubbo, F.; Moranta, D.; Tejada, S.; Jiménez, M.; Esteban, S. Impact of Gut Microbiota in Brain Ageing: Polyphenols as Beneficial Modulators. Antioxidants 2023, 12, 812. [Google Scholar] [CrossRef] [PubMed]
  422. Oh, S.; Gwak, J.; Park, S.; Yang, C.S. Green tea polyphenol EGCG suppresses Wnt/β-catenin signaling by promoting GSK-3β-and PP2A-independent β-catenin phosphorylation/degradation. Biofactors 2014, 40, 586–595. [Google Scholar] [CrossRef] [PubMed]
  423. Flowers, A.; Lee, J.-Y.; Acosta, S.; Hudson, C.; Small, B.; Sanberg, C.D.; Bickford, P.C.; Grimmig, B. NT-020 treatment reduces inflammation and augments Nrf-2 and Wnt signaling in aged rats. J. Neuroinflamm. 2015, 12, 174. [Google Scholar] [CrossRef]
  424. Shakoor, H.; Feehan, J.; Apostolopoulos, V.; Platat, C.; Al Dhaheri, A.S.; Ali, H.I.; Ismail, L.C.; Bosevski, M.; Stojanovska, L. Immunomodulatory effects of dietary poly-phenols. Nutrients 2021, 13, 728. [Google Scholar] [CrossRef]
  425. Khan, H.; Sureda, A.; Belwal, T.; Çetinkaya, S.; Süntar, I.; Tejada, S.; Devkota, H.P.; Ullah, H.; Aschner, M. Polyphenols in the treatment of autoimmune diseases. Autoimmun. Rev. 2019, 18, 647–657. [Google Scholar] [CrossRef]
  426. Omoigui, S. The Interleukin-6 inflammation pathway from cholesterol to aging–Role of statins, bisphosphonates and plant polyphenols in aging and age-related diseases. Immun. Ageing 2007, 4, 1. [Google Scholar] [CrossRef] [PubMed]
  427. Zhang, S.; Xu, M.; Zhang, W.; Liu, C.; Chen, S. Natural polyphenols in metabolic syndrome: Protective mechanisms and clinical applications. Int. J. Mol. Sci. 2021, 22, 6110. [Google Scholar] [CrossRef] [PubMed]
  428. Cheng, M.; He, J.; Wang, H.; Li, C.; Wu, G.; Zhu, K.; Chen, X.; Zhang, Y.; Tan, L. Comparison of microwave, ultrasound and ultrasound-microwave assisted solvent extraction methods on phenolic profile and antioxidant activity of extracts from jackfruit (Artocarpus heterophyllus Lam.) pulp. LWT 2023, 173, 114395. [Google Scholar] [CrossRef]
  429. Olaimat, A.N.; Al-Holy, M.A.; Abughoush, M.H.; Daseh, L.; Al-Nabulsi, A.A.; Osaili, T.M.; Al-Rousan, W.; Maghaydah, S.; Ayyash, M.; Holley, R.A. Survival of Salmonella enterica and Listeria monocytogenes in date palm paste and syrup at different storage temperatures. J. Food Sci. 2023, 88, 2950–2959. [Google Scholar] [CrossRef]
  430. Khan, M.K.; Ahmad, K.; Hassan, S.; Imran, M.; Ahmad, N.; Xu, C. Effect of novel technologies on polyphenols during food processing. Innov. Food Sci. Emerg. Technol. 2018, 45, 361–381. [Google Scholar] [CrossRef]
  431. Barbera, M. Reuse of food waste and wastewater as a source of polyphenolic compounds to use as food additives. J. AOAC Int. 2020, 103, 906–914. [Google Scholar] [CrossRef]
  432. Cai, H.; You, S.; Xu, Z.; Li, Z.; Guo, J.; Ren, Z.; Fu, C. Novel extraction methods and potential applications of polyphenols in fruit waste: A review. J. Food Meas. Charact. 2021, 15, 3250–3261. [Google Scholar] [CrossRef]
  433. Ben-Othman, S.; Jõudu, I.; Bhat, R. Bioactives from Agri-Food Wastes: Present Insights and Future Challenges. Molecules 2020, 25, 510. [Google Scholar] [CrossRef] [PubMed]
  434. Kuppusamy, S.; Venkateswarlu, K.; Megharaj, M. Examining the polyphenol content, antioxidant activity and fatty acid composition of twenty-one different wastes of fruits, vegetables, oilseeds and beverages. SN Appl. Sci. 2020, 2, 673. [Google Scholar] [CrossRef]
  435. Di Donato, P.; Taurisano, V.; Tommonaro, G.; Pasquale, V.; Jiménez, J.M.S.; de Pascual-Teresa, S.; Poli, A.; Nicolaus, B. Biological properties of polyphenols extracts from agro industry’s wastes. Waste Biomass Valoriz. 2018, 9, 1567–1578. [Google Scholar] [CrossRef]
  436. Vijayalaxmi, S.; Jayalakshmi, S.K.; Sreeramulu, K. Polyphenols from different agricultural residues: Extraction, identification and their antioxidant properties. J. Food Sci. Technol. 2014, 52, 2761–2769. [Google Scholar] [CrossRef]
  437. Teixeira, A.; Baenas, N.; Dominguez-Perles, R.; Barros, A.; Rosa, E.; Moreno, D.A.; Garcia-Viguera, C. Natural bioactive compounds from winery by-products as health promoters: A review. Int. J. Mol. Sci. 2014, 15, 15638–15678. [Google Scholar] [CrossRef]
  438. Ramón-Gonçalves, M.; Alcaraz, L.; Pérez-Ferreras, S.; León-González, M.E.; Rosales-Conrado, N.; López, F.A. Extraction of polyphenols and synthesis of new activated carbon from spent coffee grounds. Sci. Rep. 2019, 9, 17706. [Google Scholar] [CrossRef]
  439. Vamanu, E.; Gatea, F.; Pelinescu, D.R. Bioavailability and Bioactivities of Polyphenols Eco Extracts from Coffee Grounds after In Vitro Digestion. Foods 2020, 9, 1281. [Google Scholar] [CrossRef]
  440. Pasrija, D.; Anandharamakrishnan, C. Techniques for extraction of green tea polyphenols: A review. Food Bioprocess Technol. 2015, 8, 935–950. [Google Scholar] [CrossRef]
  441. Üstündağ, G.; Erşan, S.; Özcan, E.; Özan, G.; Kayra, N.; Ekinci, F.Y. Black tea processing waste as a source of antioxidant and antimicrobial phenolic compounds. Eur. Food Res. Technol. 2016, 242, 1523–1532. [Google Scholar] [CrossRef]
  442. Rasouli, H.; Farzei, M.H.; Khodarahmi, R. Polyphenols and their benefits: A review. Int. J. Food Prop. 2017, 20, 1700–1741. [Google Scholar] [CrossRef]
  443. e Silva, K.F.C.; Strieder, M.M.; Pinto, M.B.C.; Rostagno, M.A.; Hubinger, M.D. Processing Strategies for Extraction and Concentration of Bitter Acids and Polyphenols from Brewing By-Products: A Comprehensive Review. Processes 2023, 11, 921. [Google Scholar] [CrossRef]
  444. Bravi, E.; De Francesco, G.; Sileoni, V.; Perretti, G.; Galgano, F.; Marconi, O. Brewing by-product upcycling potential: Nutritionally valuable compounds and antioxidant activity evaluation. Antioxidants 2021, 10, 165. [Google Scholar] [CrossRef]
  445. 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]
  446. Cuevas, A.; Saavedra, N.; Salazar, L.A.; Abdalla, D.S.P. Modulation of immune function by polyphenols: Possible contribution of epigenetic factors. Nutrients 2013, 5, 2314–2332. [Google Scholar] [CrossRef]
  447. Russo, G.L.; Vastolo, V.; Ciccarelli, M.; Albano, L.; Macchia, P.E.; Ungaro, P. Dietary polyphenols and chromatin remodeling. Crit. Rev. Food Sci. Nutr. 2017, 57, 2589–2599. [Google Scholar] [CrossRef]
  448. Rajendran, P.; Abdelsalam, S.A.; Renu, K.; Veeraraghavan, V.; Ben Ammar, R.; Ahmed, E.A. Polyphenols as potent epigenetics agents for cancer. Int. J. Mol. Sci. 2022, 23, 11712. [Google Scholar] [CrossRef]
  449. Arora, I.; Sharma, M.; Sun, L.Y.; Tollefsbol, T.O. The epigenetic link between polyphenols, aging and age-related diseases. Genes 2020, 11, 1094. [Google Scholar] [CrossRef]
  450. Borsoi, F.T.; Neri-Numa, I.A.; de Oliveira, W.Q.; de Araújo, F.F.; Pastore, G.M. Dietary polyphenols and their relationship to the modulation of non-communicable chronic diseases and epigenetic mechanisms: A mini-review. Food Chem. Mol. Sci. 2023, 6, 100155. [Google Scholar] [CrossRef]
  451. Silva, L.B.A.R.; Pinheiro-Castro, N.; Novaes, G.M.; Pascoal, G.d.F.L.; Ong, T.P. Bioactive food compounds, epigenetics and chronic disease prevention: Focus on early-life interventions with polyphenols. Food Res. Int. 2019, 125, 108646. [Google Scholar] [CrossRef]
Figure 1. PRISMA diagram of literature search and study selection.
Figure 1. PRISMA diagram of literature search and study selection.
Antioxidants 13 00429 g001
Figure 2. Chemistry of polyphenols.
Figure 2. Chemistry of polyphenols.
Antioxidants 13 00429 g002
Figure 3. Various sources of dietary polyphenols.
Figure 3. Various sources of dietary polyphenols.
Antioxidants 13 00429 g003
Figure 4. Uptake of dietary polyphenols (absorption) and their further metabolism.
Figure 4. Uptake of dietary polyphenols (absorption) and their further metabolism.
Antioxidants 13 00429 g004
Figure 5. Pharmacological properties of antioxidant polyphenols in various disease models.
Figure 5. Pharmacological properties of antioxidant polyphenols in various disease models.
Antioxidants 13 00429 g005
Figure 6. Anti-inflammatory actions of polyphenols counteracting oxidative stress in diabetes mellitus.
Figure 6. Anti-inflammatory actions of polyphenols counteracting oxidative stress in diabetes mellitus.
Antioxidants 13 00429 g006
Figure 7. Role of polyphenols in respiratory health.
Figure 7. Role of polyphenols in respiratory health.
Antioxidants 13 00429 g007
Table 1. Classification of polyphenols.
Table 1. Classification of polyphenols.
Sr. No.Polyphenolic ClassCategoryExampleReferences
1.Phenolic acidHydroxybenzoic acidGallic acid[37]
Hydroxycinnamic acidsp-coumaric acid[38]
Caffeic acid[39]
Ferulic acid[40]
Sinapic acid[41]
2.FlavanoidsIsoflavones, Neoflavones and ChalconesGenistein and daidzein[42]
Flavones, Flavonols, Flavanones and FlavanonolsQuercetin and Kaempferol[43]
Flavanols and ProanthocyanidinsCatechins[44]
Anthocyanins [45]
3.Stilbenes Resveratrol[28,46]
4.TaninsCondensed tannins [30,31]
Hydrolysable tannins
Complex tannins
Phlorotannins
5.Lignans Matairesinol[47]
Secoisolariciresinol[48]
6.Coumarins Umbelliferone[36]
Aesculin
Psoralen
Table 3. Pharmacological activity of polyphenols with in vitro/in vivo models.
Table 3. Pharmacological activity of polyphenols with in vitro/in vivo models.
PolyphenolsIn Vitro/In Vivo ModelPharmacological ActivityReferences
Red Wine Extracts:
  • Malvidin-3-glucoside
  • Petunidin-3-glucoside
  • Malvidin-3-cumaroylglucoside
HT-29 cells
  • Red wine extract inhibits cytokine-induced IkB degradation.
  • Red wine extract inhibits COX-2 induction by cytokines.
  • Red wine extract inhibits iNOS induction by cytokines.
[147]
Pomegranate Juice Extract:
  • Punicalagin
  • Ellagic acid
Liposome model
HT-29 cells
  • Inhibition of cell proliferation of nonmetastatic colon cancer cells.
  • Induced apoptosis in HT-29 cells.
[148]
Apple Polyphenol Extracts:
  • Catechins
  • Chlorogenic acid
MKN 28 cells
Male Wistar rats
  • Prevention of xanthine-xanthine oxidase-induced cell injury.
  • Helps inhibit ROS-induced lipid peroxidation.
  • Significantly prevents indomethacin injury in vivo as well as in vitro.
[149]
Phenolic Acids:
  • Chlorogenic acid
  • 4-coumaroylquinic acid
Flavonols:
  • Quercetin-3-rutinoside
  • Quercetin-3-rhamnoside
HT-29 cells


CaCo-2 cells
  • Exhibit preventive effectiveness in colon cell lines.
[150]
Procyanidins:
  • Procyanidin dimer
  • Procyanidin trimer C1

IPEC-1 cells
TOPIG hybrid pigs
  • Resulted in a decrease in Lipid peroxidation.
  • It helps increase the total antioxidant status.
[151]
Anthocyanins:
  • Delphinidin 3-galactoside
  • Cyanidin 3-galactoside
CaCo-2 cells
  • Exhibit an intracellular antioxidant activity, thereby protecting cells from ROS.
[152]
Green Tea Polyphenols:
  • Epigallocatechin gallate
BALB/c mice with
DSS-induced colitis
  • Inhibit the activity of nuclear factor-κB.
  • Help lower the concentration of TNF-α.
[153]
Tomato Extract:
  • Resveratrol
  • Quercetin
C57/BL6 mice with
DSS-induced colitis
  • May be involved in the inhibition of inflammatory factors.
  • May help promote the production of proteins utilized in tissue repair.
[154]
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

Rudrapal, M.; Rakshit, G.; Singh, R.P.; Garse, S.; Khan, J.; Chakraborty, S. Dietary Polyphenols: Review on Chemistry/Sources, Bioavailability/Metabolism, Antioxidant Effects, and Their Role in Disease Management. Antioxidants 2024, 13, 429. https://doi.org/10.3390/antiox13040429

AMA Style

Rudrapal M, Rakshit G, Singh RP, Garse S, Khan J, Chakraborty S. Dietary Polyphenols: Review on Chemistry/Sources, Bioavailability/Metabolism, Antioxidant Effects, and Their Role in Disease Management. Antioxidants. 2024; 13(4):429. https://doi.org/10.3390/antiox13040429

Chicago/Turabian Style

Rudrapal, Mithun, Gourav Rakshit, Ravi Pratap Singh, Samiksha Garse, Johra Khan, and Soumi Chakraborty. 2024. "Dietary Polyphenols: Review on Chemistry/Sources, Bioavailability/Metabolism, Antioxidant Effects, and Their Role in Disease Management" Antioxidants 13, no. 4: 429. https://doi.org/10.3390/antiox13040429

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop