Next Article in Journal
Analysis of Structure and Antioxidant Activity of Polysaccharides from Aralia continentalis
Next Article in Special Issue
Gold Nanoparticles Encapsulated Resveratrol as an Anti-Aging Agent to Delay Cataract Development
Previous Article in Journal
Progranulin in Musculoskeletal Inflammatory and Degenerative Disorders, Focus on Rheumatoid Arthritis, Lupus and Intervertebral Disc Disease: A Systematic Review
Previous Article in Special Issue
Pharmacological Properties of Jaeumgeonbi-Tang on Redox System and Stress-Related Hormones in Chronic Subjective Dizziness: A Randomized, Double-Blind, Parallel-Group, Placebo-Controlled Trial
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploiting Polyphenol-Mediated Redox Reorientation in Cancer Therapy

1
School of Basic Medical Sciences, State Key Laboratory of Southwestern Chinese Medicine Resources, Chengdu University of Traditional Chinese Medicine, Chengdu 611137, China
2
State Key Laboratory of Biotherapy and Cancer Center, West China School of Basic Medical Sciences & Forensic Medicine, West China Hospital, Sichuan University, Collaborative Innovation Center for Biotherapy, Chengdu 610041, China
3
Department of Encephalopathy, Chongqing Hospital of Traditional Chinese Medicine, Chongqing 400021, China
4
Hainan Cancer Center and Tumor Institute, The First Affiliated Hospital of Hainan Medical University, Haikou 570102, China
5
Department of Otorhinolaryngology and Head and Neck Surgery, The Affiliated Lihuili Hospital, Ningbo University, Ningbo 315040, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceuticals 2022, 15(12), 1540; https://doi.org/10.3390/ph15121540
Submission received: 7 November 2022 / Revised: 6 December 2022 / Accepted: 7 December 2022 / Published: 12 December 2022
(This article belongs to the Special Issue Natural Products as Drug Candidates for Redox-Related Human Disease)

Abstract

:
Polyphenol, one of the major components that exert the therapeutic effect of Chinese herbal medicine (CHM), comprises several categories, including flavonoids, phenolic acids, lignans and stilbenes, and has long been studied in oncology due to its significant efficacy against cancers in vitro and in vivo. Recent evidence has linked this antitumor activity to the role of polyphenols in the modulation of redox homeostasis (e.g., pro/antioxidative effect) in cancer cells. Dysregulation of redox homeostasis could lead to the overproduction of reactive oxygen species (ROS), resulting in oxidative stress, which is essential for many aspects of tumors, such as tumorigenesis, progression, and drug resistance. Thus, investigating the ROS-mediated anticancer properties of polyphenols is beneficial for the discovery and development of novel pharmacologic agents. In this review, we summarized these extensively studied polyphenols and discussed the regulatory mechanisms related to the modulation of redox homeostasis that are involved in their antitumor property. In addition, we discussed novel technologies and strategies that could promote the development of CHM-derived polyphenols to improve their versatile anticancer properties, including the development of novel delivery systems, chemical modification, and combination with other agents.

1. Introduction

Polyphenols, the most common natural antioxidants extracted from plants, primarily function as secondary metabolites to protect plants against reactive oxygen species (ROS), ultraviolet radiation, mechanical damage, pathogens, parasites, and predators [1,2]. It commonly exists in a wide variety of dietary sources, including vegetables, fruits, spices, and natural beverages (e.g., tea, coffee, and wine). Notably, natural herbal medicine [3,4], especially Chinese herbal medicine (CHM), shares the holistic principles of “homology of medicine and food” in traditional Chinese medicine (TCM), which is rich in polyphenols [5,6]. A plethora of studies have identified the numerous activities of polyphenols, including anti-inflammatory, antioxidant, antiviral, and anti-aging activities; therefore, they have been applied and proven efficient in various disease models, including inflammatory bowel disease, coronary heart disease, coronavirus disease 2019, Alzheimer’s disease, and particularly, cancer [7,8,9,10,11].
Polyphenols can be classified into four major categories, including flavonoids, phenolic acids, lignans, and stilbenes, according to their chemical structure [12]. Flavonoids are characterized by a C6-C3-C6 carbon backbone, which is structured as a phenyl ring (ring A) fused with a pyran ring (ring C) that carries another phenyl ring (ring B) substitution (ring C) [13]. Based on this structure, flavonoids are subdivided into six major subclasses: flavonols, flavanols, anthocyanins, flavones, flavanones, and isoflavones [14,15] (Figure 1). Flavonols are ketones with a hydroxyl group at position 3 of ring C [16], while flavanols lack a ketone found in flavonols, instead of a hydroxyl group at position 3 of ring C [17]. Unlike flavonols, anthocyanidins lack a ketone group but contain positively charged oxygen at position 4 of ring C [18]. Phenolic acids, as one of the main classes of phenolic compounds, are produced through the phenylpropanoid pathway and possess phenol moieties and resonance-stabilized structures [19,20]. Phenolic acids are typically bound with esters, glycosides, or amides, but rarely exist in their free form [21]. They mainly consist of two subgroups: hydroxycinnamic and hydroxybenzoic acids [22,23]. Hydroxycinnamic acids are present in simple esters with quinic acid or glucose, containing a saturated tail followed by carboxylic acid [24,25]. Hydroxybenzoic acids, such as vanillic, gallic, protocatechuic, ellagic, and syringic acids, lack tail saturation [17]. Lignans are natural compounds polymerized by phenylpropanoid (C6-C3) derivatives in different ways, usually forming dimers [26] with a few trimers and tetramers [27,28]. The phenylpropane dimers are also divided into lignans or neolignans based on the presence or absence of the 8,8′-bond between phenylpropanoid monomers [26]. Stilbenes contain two benzene rings joined by a molecule of ethanol or ethylene to present a C6-C2-C6 structure [29,30,31]. They are characterized by the presence of a 1,2-diphenylethylene nucleus with hydroxyl groups substituted on the aromatic rings [32], in which resveratrol is well known for its versatile activities [33].
ROS contradictorily influence cancer evolution, either triggering tumorigenesis, promoting cancer cell transformation, or leading to tumor regression and even cell death [34]. Intracellular oxidative stress can lead to ROS elevation. To accommodate abnormal ROS levels, cancer cells modify different antioxidant systems, including antioxidant transcription factors, nicotinamide adenine dinucleotide phosphate (NADPH) generation, and metabolic reprogramming, to neutralize oxidative stress [35,36]. During their initiation, stimulating antioxidant transcription factors and increasing NADPH generation enhance the survival of cancer cells [37,38]. Cancer cells are able to activate the pentose phosphate pathway to increase the nucleic acid supply to maintain proliferative capacity. Furthermore, the activation of AMPK signaling and elevation in reductive glutathione and folate metabolism lead to an increase in NADPH synthesis, which protects cancer cells from oxidative stress during metastasis [39]. Moreover, activated transcription factors such as nuclear factor erythroid 2-related factor 2 (NRF2) and hypoxia inducible factor-1α (HIF-1α) promote cancer cells to adapt to drug-mediated oxidative stress (chemoresistance) [40,41]. Therefore, cancer cells exhibit aberrant redox homeostasis to survive the harsh tumor microenvironment (Figure 2).
Disrupting the homeostasis of oxidative stress and a reduced state have been proven to be significant strategies for cancer treatment [42,43]. Some chemotherapeutic agents, such as doxorubicin, oxaliplatin, and paclitaxel, have been shown to exert antitumor effects by disturbing intracellular redox homeostasis [44,45,46], but the side-effect, frequent occurrence of drug resistance and tumor recurrence impaired their use in clinical application. Currently, many natural products, such as natural vitamins, alkaloids, saponins, polypeptides, polysaccharides, and polyphenols, have been employed as antioxidants/prooxidants to treat cancer [47,48,49]. Among them, polyphenols have attracted considerable attention in cancer therapy due to their inherent characteristic of modulating oxidative stress, the critical player in cancer incidence and progression [14,50,51]. Even with the limitations of polyphenols in delivery, targeting, and preservation [52,53,54,55], an increasing amount of evidence has demonstrated that these viable anticancer candidates can regulate redox homeostasis in cancer cells by acting as antioxidants or pro-oxidants to prevent cancer [56,57]. Besides, the overwhelming advantages of polyphenols, including specificity of the response, negligible toxicity, and omnipresence, make them excellent antitumor agents [58]. Currently, researchers are devoted to developing new techniques and strategies for improving the application of polyphenols in cancer treatment [59,60].
In this review, we summarize the therapeutic mechanisms underlying the polyphenol-mediated modulation of redox homeostasis in cancer treatment and drug resistance surmounting. In addition, we discuss novel technologies or methods for optimizing the delivery, targeting, and preservation of polyphenols for effective cancer treatment.

2. Polyphenols Modulate Redox Homeostasis for Cancer Therapy

In cancer cells, moderate ROS (the ROS level that is beneficial for tissue turnover and cell proliferation) function as second messengers in cellular physiological processes to modulate cellular signaling and biological reactions to maintain endogenous homeostasis [61], thereby providing advantages for carcinogenesis, metastasis, and cell survival. However, excessive ROS beyond the toxic threshold (the maximum level suitable for cellular homeostasis, triggering the redox homeostasis to activate cell death) could impede tumor progression and trigger cell senescence, apoptosis, or ferroptosis [62,63,64,65]. Therefore, the regulation of cellular redox homeostasis via antioxidants or prooxidants holds significance for cancer therapy. In recent years, modulation of redox homeostasis using natural antioxidants/pro-oxidants for cancer treatment has been largely explored in preclinical research and clinical evaluations [66,67,68,69], and many clinical trials have been implemented (NCT01912820, NCT03493997, NCT00256334, NCT00433576, NCT01717066). As the major sources of natural antioxidants, polyphenols have attracted considerable interest as novel anticancer agents with the dual functions of regulating oxidative stress according to their properties and to the dosage in different tumor models [70,71], including colorectal cancer, breast cancer, lung cancer, gastric cancer, cervical cancer, etc. [72,73,74,75,76]. In this section, we will review the polyphenols that have been reported for their use in cancer therapy, with an emphasis on the underlying molecular mechanisms related to the modulation of redox homeostasis.

2.1. Polyphenols Mediate Antioxidant Effects in Cancer Therapy

A variety of polyphenols, including kaempferol, resveratrol, catechins, curcumin, wogonin, quercetin, etc., have been reported to suppress tumors by suspending oxidative stress via various mechanisms. Primarily, the hydroxyl group on polyphenols undergoes hydrogen atom transfer or single electron transfer reactions, thereby scavenging radicals such as hydroxyl, peroxyl, or peroxynitrite [77]. In addition, polyphenols can inhibit the activity of oxidases, including superoxide-producing enzymes, e.g., cyclooxygenase 2 (COX-2) and NADPH oxidase (NOX), or promote the activity of antioxidant enzymes, e.g., superoxide dismutase (SOD), GSH peroxidases (GPXs), peroxiredoxins (PRDXs), and catalase (CAT) [78,79,80,81,82], thus eliminating excessive ROS levels and recovering redox homeostasis. Additionally, polyphenols have been reported to activate classical antioxidant signaling pathways, such as NRF2 and forehead box class O (FOXO), and promote the transcription of antioxidant proteins, such as Skinhead-1, Heme Oxygenase-1 (HO-1), PRDX-2, SOD3, and Glutathione S-transferase-4 (GST-4) [83,84,85,86,87].

2.1.1. Kaempferol

Kaempferol, a well-known natural polyphenol derived from CHMs (e.g., Tetrastigma hemsleyanum Diels et Gilg, Sparganii Rhizoma, Lysimachiae Herba (dried whole part of Lysimachia christinae Hance)) [88,89,90]), has obvious anticancer efficiency in multiple cancer cells [91,92,93]. Kaempferol can promote the expression of NOQ1, SOD1, and HO-1 in HL-60 cells, a leukemia cell line, thus influencing the antioxidant status [94]. According to a recent report [95], kaempferol could prevent the occurrence of hemolysis through the upregulation of antioxidant enzymes and the inhibition of ROS generation and lipid peroxidation. Moreover, it inhibited the function of phosphorylated AKT (p-AKT), cyclin D1, and cyclin-dependent kinases 4 while promoting the expression of phosphorylated breast cancer susceptibility protein, p-ATM, p53, p21, p38, Bax, and Bid, eventually evoking S phase arrest and apoptosis in bladder cancer cells. Kaempferol can also prevent carcinogenesis via modulating NRF2 signaling in MCF-10A cells [96]. Mechanistically, kaempferol upregulated the expression of NRF2 and its downstream enzyme NAD(P)H: quinone oxidoreductase 1 (NQO1), thus decreasing oxidative stress, recovering redox homeostasis, and preventing carcinogenesis. In addition, kaempferol suppressed ROS production in mouse bone marrow-derived neutrophils, which was related to the blockade of neutrophil extracellular trap formation, therefore reducing lung metastasis in a mouse breast cancer model [97]. Notably, kaempferol also decreases the ROS level by upregulating the expression of JAK/STAT3, MAPK, PI3K/AKT, and NF-κB [98,99], which are able to overcome ROS-mediated drug resistance and sensitize tumor cells to 5-fluorouracil (5-FU) [100,101]. As reported by a recent study [102], the combination of kaempferol and 5-FU blocked the production of ROS and downregulated the expression of ABC subfamily G member 2 and multidrug resistance-associated protein 1 in 5-FU-resistant LS174 cells, thus surmounting the efflux of 5-FU and drug resistance. Similarly, kaempferol could sensitize oxaliplatin (Ox)-resistant HCT116 (HCT116-OxR) cells to oxaliplatin by decreasing the transactivation activity of adaptor protein complex-1 [103]. These findings demonstrate the versatile properties of kaempferol as an antioxidant for cancer treatment.

2.1.2. Resveratrol

Resveratrol, one of the most investigated polyphenols from CHMs [104,105], has been widely applied in cancer prevention due to various biological activities, especially the regulation of oxidative stress [106,107,108]. The antioxidant activities of resveratrol in cancer therapy are also related to manipulating the expression of NRF2 and its target genes, such as NQO1, SOD3, and 8-oxoguanine DNA glycosylase 1 (OGG1) [109,110], thus inhibiting estrogen-mediated DNA damage and suppressing mammary carcinogenesis [111]. In azoxymethane (AOM)- and dextran sulfate sodium (DSS)-induced cancer models, resveratrol activated crosstalk between NRF2 and mitogen-activated protein kinase phosphatase 1 to inhibit oxidative stress and prevent carcinogenesis [112]. In addition, resveratrol suppresses and abolishes the phosphorylation of key regulators, such as transforming growth factor (TGF)-β1, Smad2 and Smad3, which play essential roles in epithelial mesenchymal transition (EMT) to retard EMT and lung metastasis processes in breast cancer [113]. Moreover, resveratrol also showed the potential to overcome drug resistance. An experiment indicated that resveratrol could enhance the expression of NRF2 and suppress the level of nutrient-deprivation autophagy factor-1 (NAF-1), which led to cell death in pancreatic cancer cells [114]. More importantly, resveratrol could target NAF-1 and improve the sensitivity of pancreatic cancer cells to gemcitabine.

2.1.3. Catechins

Catechins are the most abundant polyphenols in tea (the second most consumed beverage and the most commonly used CHM worldwide [115,116]), mainly including epigallocatechin-3-gallate (EGCG), epigallocatechin, epicatechin-3-gallate, and epicatechin [117,118]. Among them, EGCG is the most abundant tea polyphenol and has attracted considerable attention in cancer therapy due to its antioxidant activity [119]. Normally, EGCG can scavenge ROS and mitigate the damage of oxidative DNA and protein to protect against cancer [120]. Basically, the phenolic hydroxyl groups of EGCG endow it with the ability to combine with ROS by oxidizing the B and D ring to form stable phenoxy radicals, thereby downregulating cellular ROS levels [119]. In addition, ECCG can modulate the activity of antioxidative enzymes or oxidases to suppress the formation of byproducts such as malondialdehyde and 8-hydroxy-2′-deoxyguanosine [121,122,123]. For example, EGCG administration could inhibit the process of multistage mouse skin carcinogenesis by decreasing the expression of oxidative enzymes, such as inducible nitric oxide synthase (iNOS) and COX-2 [124]. Oral gavage of tea polyphenols, of which EGCG is the major component, could significantly prevent diethylnitrosamine/phenobarbital-induced hepatocarcinogenesis in a rat model, which was attributed to the increased total antioxidant capacity [125]. In addition, EGCG could upregulate the mRNA and protein levels of NRF2, uridine 5′-diphosphate-glucuronosyltransferase (UGT)1A, and UGT1A8, followed by the suppression of proliferation and liver and lung metastasis in an HT-29 cancer cell mouse model [126]. Similarly, oral administration of EGCG inhibited the miR483-3p-induced enhancement of human hepatocellular carcinoma cell migration and invasion by counteracting EMT markers [127]. Moreover, EGCG exhibited anticancer ability by maintaining an optimum level of NRF2 to surmount etoposide resistance in lung cancer cells by regulating the activity of KEAP1, thus inducing G2/M arrest and overcoming multidrug resistance [128].

2.1.4. Other Antioxidant Polyphenols

There are many other polyphenols that can be applied in cancer treatment due to their antioxidant capability. Curcumin, a well-known polyphenol extracted from the rhizomes of Curcuma longa, was reported to inhibit the proliferation of breast cancer cells through the NRF2-mediated downregulation of Flap endonuclease 1, a DNA repair-specific nuclease. In detail, curcumin could trigger NRF2 translocation from the cytoplasm to the nucleus to bind to the promoter of Fen1 and decrease its transcriptional activity [129]. In recent studies, curcumin has been proven to increase the overall survival of cancer patients in various clinical trials [NCT01042938, NCT01160302, NCT03211104]. Wogonin, another polyphenol extracted from the root of Scutellaria baicalensis Georgi (a kind of CHM), has been used to treat inflammatory diseases [130]. It is now reported to promote nuclear translocation of NRF2 and counteract the levels of interleukin (IL)-6 and IL-1β, therefore relieving inflammation-associated oxidative stress and preventing colorectal carcinogenesis induced by AOM and DSS in a mouse model [131]. Moreover, quercetin, a commonly used polyphenol derived from CHMs and other edible plants, has exhibited an excellent anticancer capability by modulating redox [132]. Quercetin inhibited the migration and invasion of SAS human oral cancer cells through the downregulation of matrix metalloproteinase (MMP)-2 and MMP-9. In detail, quercetin reduced the protein activity of vascular endothelial growth factor (VEGF), NF-κB, iNOS, and COX-2 to eliminate ROS and restrain metastasis [133]. In fact, there are too many kinds of natural polyphenols to list, but they all exhibit remarkable tumor suppressing efficiency by impairing cellular oxidative stress [134,135].
Taken together, polyphenols can directly eliminate ROS to relieve oxidative stress. Additionally, they manipulate the activity of some redox signaling pathways and antioxidant enzymes to retard tumor progression, including carcinogenesis, proliferation, metastasis, and drug resistance, in a variety of cancer types (Figure 3).

2.2. Polyphenols Suppressive Cancer by Promoting Oxidative Stress

Several kinds of polyphenols also exhibit prooxidant capabilities in cancer therapy. Intriguingly, some kinds of polyphenols exhibit antioxidant activity, but high doses of them induce pro-oxidative stress and lead to DNA damage, lipid peroxidation, inflammasomes, and autophagosome augmentation [136,137,138,139], thus causing apoptosis, ferroptosis, autophagy, and pyroptosis in cancer cells [140,141,142,143]. Obviously, the pro-oxidant mechanism is similarly involved in the modulation of the redox signaling pathway or the activity of oxidative enzymes [144,145].

2.2.1. Curcumin

A high dosage of curcumin also exhibits a pro-oxidant potential to suppress tumors. For example, curcumin could evoke cytotoxic cell death in non-small cell lung cancer (NSCLC) by promoting ROS accumulation and mitochondrial transmembrane potential reduction, thus causing DNA damage, G2/M arrest, and subsequent mitochondrial apoptosis [146]. In addition, several other ROS-mediated cell death models, such as ferroptosis, autophagy, and pyroptosis, could be induced by curcumin to kill cancer cells [147]. For instance, Tang and his coworkers revealed that curcumin administration led to iron overload, glutathione (GSH) depletion, and lipid peroxidation, causing the downregulation of Recombinant Solute Carrier Family 7, Member 11, and Glutathione peroxidase 4 (GPX4) in lung cancer cells [148]. Moreover, oxidative stress induced autolysosome accumulation and autophagy-dependent ferroptosis, consequently inhibiting the growth and proliferation of cancer cells. In hepatocellular carcinoma, curcumin intervention triggered the elevation of ROS levels to activate caspase-3 for cleavage of Gasdermin E, the executor protein of pore formation, thereby causing pyroptosis in HepG2 cells [143]. Intriguingly, curcumin could suppress the expression of key antioxidant enzymes, including SOD, CAT, GPX, and HO-1, and suspend the activity of the transcription factor NRF2, therefore increasing intracellular oxidative stress to reverse cisplatin resistance in SKOV-3 ovarian cancer cells [149]. The combined strategy of tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and curcumin use synergistically sensitized ACHN renal cancer cells to TRAIL in a mechanism of ROS-mediated activation of the JNK-CHOP pathway [150]. Similarly, its analog, L48H37, could induce ROS accumulation to evoke endoplasmic reticulum (ER) stress through phosphorylation of PERK and eIF2α and subsequently trigger cell cycle arrest and apoptosis in lung cancer cells [151].

2.2.2. Wogonin

Wogonin is another compound that shows prooxidant effects in cancer treatment. For example, wogonin could elicit abundant ROS to activate caspase 3/8/9 and subsequent apoptosis, thus contributing to the inhibition of lung cancer cell proliferation [152]. In some contexts, NRF2 exerts a protective effect against chemotherapy-mediated oxidative stress, which promotes cancer cell survival and drug resistance [153]. Wogonin could downregulate the expression of NRF2 as well as its target proteins HO-1, NADP(H), and NQO-1 to impair the NRF2-mediated antioxidant system and reverse DOX resistance in MCF-7ADR breast cancer cells [154]. Further exploration demonstrated that the downregulation of NRF2 by wogonin could potentiate the cytotoxicity of chemotherapeutic agents by inhibiting multidrug resistance-associated proteins (MRPs) and inducing phage II enzymes [155]. Similarly, the inhibition of MRP1 by wogonin relies on the dissociation of NRF2 from antioxidant response elements (AREs) in human myelogenous leukemia, in which the PI3K/Akt pathway and DNA-PKcs are activated [156]. A similar mechanism was also observed in NRF2-mediated cisplatin resistance in head and neck cancer, and wogonin surmounted this phenotype [157].

2.2.3. Resveratrol

Notably, resveratrol has been observed to have outstanding anticancer potential in many types of cancer cells, such as bladder, prostate, glioblastoma, colon, breast, lung, and ovarian cancer cells [158,159,160,161]. As an important source of CHM-derived polyphenols, resveratrol prevents carcinogenesis through its antioxidant ability in specific tumor types while also displaying its versatility in cancer treatment by exerting prooxidant effects. For example, a high concentration of resveratrol could induce ROS overproduction and LC3 expression and decrease mitochondrial membrane potential, which results in autophagy and proliferation suppression in HeLa cells [162]. Similarly, the increased ROS levels induced by resveratrol inhibit the activity of casein kinase 2, a protein related to proliferation and apoptosis, thus affecting mitochondrial function and the viability of breast cancer cells [163]. It has also been reported that resveratrol can inhibit the activity of antioxidant enzymes such as SOD and CAT to influence the outcome of cancer treatment. Recent studies have shown that resveratrol induces the elevation of intracellular ROS by attenuating SOD2 and CAT expression and promoting the activity of caspase-9 and caspase-3, thereby remarkably arresting proliferation and triggering extensive apoptosis in glioblastoma multiforme (GBM) cells [164]. Indeed, resveratrol could also manipulate the NRF2-related signaling pathway. In a related report, resveratrol significantly decreased the mRNA expression of Nrf2 to reverse drug resistance in HL-60ADR (Adriamycin resistant HL-60 cells) [110]. Together, these results provide a rational proposal to take advantage of the prooxidant capacity of resveratrol to combat cancer.

2.2.4. Other Prooxidant Polyphenols

Some other polyphenols could also display prooxidant potential in cancer treatment. Similar to resveratrol, kaempferol can act as an oxidative stress inducer. The downregulation of NRF2 by kaempferol impaired the transcription of its target genes, including NQO-1, HO-1, AKR1C1, and GST, to enhance ROS accumulation and apoptosis in NSCLC cells [165]. Rosmarinic acid (RA), a natural polyphenol extracted from Perilla frutescens and Glechoma hederacea L. [166,167], also exhibits pro-oxidant activity. Related studies have identified that RA can trigger intracellular ROS production to upregulate the cleavage rates of caspase-8/caspase-9/caspase-3 and inhibit the expression of MMP-9/MMP-2, thus promoting apoptosis and retarding EMT in osteosarcoma cells [168]. In addition, another study reported that RA promotes ER-mediated oxidative stress to induce bax translocation and caspase-3 cleavage, which eventually triggers apoptosis in oral cancer cells [169]. Similar to RA, the prooxidant function of quercetin can be attributed to the induction of ROS-mediated oxidative stress that results from the inhibition of antioxidant enzymes or activation of oxidase [170]. For example, quercetin could induce apoptosis by decreasing mitochondrial membrane potential and inducing ER-mediated activation of activating transcription factor (ATF)-6α and ATF-6β in oral cancer SAS cells [171]. Furthermore, quercetin can increase the activity of COX-2 and elevate the intracellular ROS level, which induces apoptosis and inhibits cell survival in human colon cancer [172]. Several other studies have also demonstrated the antitumor role of quercetin by exacerbating oxidative stress and inducing cell death in a variety of tumor models [173,174].
Taken together, the prooxidant function of polyphenols is indispensable to manipulating redox homeostasis in cancer cells. The aberration of redox signaling and intracellular oxidative stress induced by polyphenols can lead to macromolecule damage, such as DNA or protein damage, and lipid peroxidation, which triggers programmed cell death (PCD), such as apoptosis, ferroptosis, and pyroptosis, thereby playing an essential role in cancer prevention and treatment, as well as in overcoming drug resistance (Figure 4).

3. Novel Strategies Facilitate the Application of Polyphenols in Cancer Therapy

The characteristics of polyphenols in modulating redox homeostasis have been widely applied in cancer prevention and treatment, which lays the basis for the discovery and development of natural anticancer drugs. Indeed, many polyphenols have been explored in preclinical or clinical trials, but the drawbacks of polyphenols generally disturb their versatile properties in clinical settings [175]. For instance, specific structures, such as phenolic hydroxyl groups and the catechol ring of polyphenols, make them easy to oxidize [176], which greatly deteriorates their stability and increases the probability of degradation [177]. In addition, their poor water solubility, their inadequate bioavailability, and more importantly, the nonspecific selectivity of polyphenols limit their pharmacological applications [178,179]. Moreover, the single use of polyphenols always compromises their limited cytocidal effect [180]. Robust strategies have been developed to surmount these limitations to accelerate the efficient implementation of CHM-derived polyphenols for cancer treatment (Figure 5).

3.1. Modification

Structural modifications, such as esterification, methylation, and glycosylation, can avoid degradation and enhance the bioactivities of polyphenols. For example, the esterification of EGCG through the substitution of hydroxyl groups with the chain of fatty acids not only increased lipophilicity, but also promoted its antioxidant capacity via enhanced hydrogen atom donation [181]. Similarly, a lipophilized EGCG derivative (LEGCG) synthesized by a partial esterification reaction (an enzymatic esterification model) of EGCG with lauric acid improved its bioactivity, including anti-proliferation and pro-apoptosis effects [182]. In addition, the methylation of EGCG, which alters the phenolic hydroxyl groups of the EGCG benzene ring into methyl ether, could amend its oral absorption rate and blood stability. For instance, the in vivo bioavailability and stability of EGCG was greatly enhanced when the hydroxyl groups were replaced by more stable methoxy groups [183]. Similarly, the bioavailability of methylated EGCG was higher than that of free or unmethylated EGCG [184]. The glycosylation of polyphenols can improve their solubility and stability and protect these compounds from oxidants, light degradation, and hostile gastrointestinal conditions [185,186]. Intriguingly, the glycosylated EGCG could act as a prodrug and first be deglycosylated at the intestinal surface before diffusing into enterocytes, thereby increasing the stability of EGCG during processing, storage, and gut transit after ingestion [187]. Moreover, the modifications for polyphenols could enhance the purification efficiency, prolong the preservation time, and avoid degradation in elevated large-scale production and commercialization [188,189,190]. Taken together, the modification of polyphenols evades rapid degradation and enhances their accessibility, thus providing rational strategies to accelerate the application of polyphenols in cancer treatment.

3.2. Nano Strategies

Recent evidence has shown that nano strategies can aid in overcoming polyphenols’ inherent drawbacks, including their low water solubility, poor stability, and nontargeting ability [191,192]. A myriad of nano strategies has been well established for improving pharmacokinetic properties under polyphenol administration [193,194]. Some nanocarriers, such as nanoparticles, liposomes, hydrogels, and extracellular vehicles, are widely applied to deliver polyphenols for cancer treatment [195,196,197]. The nanostructured lipid carrier can encapsulate kaempferol to optimize its low aqueous solubility and poor bioavailability [198]. After being modified with hyaluronic acid (HA), the HA-KA-NLC nanoplatform could target NSCLC cells by recognizing highly expressed CD44, thus exhibiting more efficient inhibition of their proliferation and EMT than free kaempferol administration. Indeed, these nano strategies strengthened redox regulation, which synergizes with the enhanced bioavailability to greatly improve therapeutic efficiency. One recent study reported a solid self-nanoemulsifying drug delivery system (s-SNEDDS) loaded with resveratrol and tamoxifen to treat breast cancer [199]. The s-SNEDDS team not only improved the bioavailability of resveratrol, but also sensitized tamoxifen-mediated chemotherapy and exhibited satisfactory suppression of MCF-7 breast cancer cells by triggering resveratrol-induced ROS elimination and SOD activation. In addition, a tannic acid-loaded dual antioxidant-photosensitizing hydrogel system was established to protect against human melanoma. In this nanosystem, chitosan-based hydrogels were designed using tannic acid as an antioxidant cross-linker and loaded with photosensitizer PDI-Ala, in which the tannic acid controlled the ROS generation and minimized the side effects of singlet oxygen synergistically with PDI-Ala boosted photodynamic therapy [200]. Recently, natural exosome-like nanovesicles (ENs) have endured much investigation as novel carriers and therapeutic agents. In Zu’s study, ENs extracted from green tea leaves were rich in polyphenols such as EGCG, quercetin, and myricetin and exhibited distinct efficiency in preventing AOM- and DSS-induced colorectal carcinogenesis in a mouse model by maintaining intracellular redox homeostasis [201]. Similarly, tea flower-derived ENs could inhibit breast cancer metastasis by stimulating ROS amplification [202]. Taken together, nano strategies ideally tackle the limitations of polyphenols, making them promising agents for cancer treatment.

3.3. Combination with Other Agents

Strategies to combine polyphenols with other agents play important roles in preclinical evaluation and clinical implementation, including amplifying ROS-mediated therapeutic efficiency and avoiding chemotherapy-induced side effects [203]. The combinational use of EGCG with metformin, a classical antidiabetic drug, stimulated intracellular ROS accumulation induced by EGCG (100 μM) through the modulation of Sirtuin 1-dependent deacetylation on NRF2, thus augmenting the anticancer effect of EGCG in NSCLC treatment [204]. Indeed, the combination of polyphenols with first-line chemotherapies such as paclitaxel, 5-fluorouracil, and oxaliplatin could not only augment the cytocidal function but also reduce the side effects of these chemotherapies [205,206]. In addition, combined treatment with quercetin and paclitaxel could significantly inhibit proliferation and migration and evoke apoptosis via increased ROS generation, as well as attenuate the side effects of paclitaxel in PC-3 prostate cancer cells [207]. Similarly, the combination of curcumin also sensitizes cancer cells to oxaliplatin and alleviates oxaliplatin-induced peripheral neuropathic pain by inhibiting the oxidative stress-mediated activation of NF-κB [208,209,210]. Interestingly, a recent phase I trial showed the safety, tolerability, and feasibility of administering curcumin as an adjunct to FOLFOX (5-fluorouracil, folinic acid, and oxaliplatin) chemotherapy in patient-derived colorectal or liver metastases cancer [NCT01490996]. Alternatively, resveratrol exhibited a synergetic effect with 5-fluorouracil to induce an imbalance in cellular antioxidant activities and subsequent intracellular ROS accumulation and lipid peroxides, thus leading to a significant decrease in long-term colon cell survival [211]. In summary, these novel strategies ingeniously alleviate the predicament of polyphenols and enhance the feasibility of developing new agents with redox regulation’s ability to overcome cancer progression.

4. Conclusions and Perspective

Oxidative stress acts as a double-edged sword in tumor progression; thus, the strategy to manipulate ROS levels to influence redox homeostasis is significant for targeting tumors. Notably, polyphenols can regulate oxidative stress in cancer treatment. In this review, focusing on redox homeostasis modulation and the involved molecular mechanisms, we have summarized several polyphenols that can function at different stages of cancer progression. We have also discussed several novel strategies, including chemical modification, nanotechnology, and combined treatment, that polyphenols could facilitate as optimum therapeutic agents in clinical settings.
A myriad of preclinical and clinical trials has demonstrated the potential value of polyphenols in the prevention of carcinogenesis, limitation of proliferation, inhibition of metastasis, and overcoming of drug resistance. Given that a large number of polyphenols are derived from CHMs and are easily acquired from a common daily diet, polyphenols are recognized as promising candidates for cancer therapy. However, several issues remain to be resolved before they can actually be used in the clinic. Primarily, their poor solvability, targeting ability, and stability seriously impede their application. In addition, some reports revealed that several antioxidants lead to the progression of tumors [212,213,214]. Moreover, the precision of dosage and pharmacokinetics remains a realistic problem in gastrointestinal transit through oral administration [215]. For these considerations, several strategies, including chemical modification, nanotechnology, and combined treatment, have been adopted to optimize the project in daily application. It is worth noting that the antagonistic potential of polyphenols reminds us to establish rigorous in vitro and in vivo screening systems before creating daily use and depth methods to explore the accuracy of dosage and pharmacokinetics. Furthermore, novel strategies still need investment. Favorable clinical data are also largely needed to confirm their tangible benefit in cancer therapy. Only through accruing such knowledge and developing these strategies will polyphenols become an optimum therapeutic agent in daily application.

Author Contributions

Conceptualization, Z.S. and J.L.; writing—original draft preparation, L.L. and P.J.; writing—review and editing, M.L., Y.W., B.H. and B.L.; visualization, Y.G., J.C. and K.H.; supervision, C.H.; funding acquisition, Z.S. and J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the Hainan Province Science and Technology special fund (ZDYF2020132, ZDYF2020137, ZDYF2022SHFZ065), the specific research fund of the Innovation Platform for Academicians of Hainan Province (YSPTZX202208), the Hainan Province Clinical Medical Center (QWYH202175), Ningbo Clinical Research Center for Otolaryngology Head and Neck Disease (No. 2022L005), and the Ningbo Medical and Health Brand Discipline (No. PPXK2018-02).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Figures in this review were created with the BioRender platform.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Di Meo, F.; Valentino, A.; Petillo, O.; Peluso, G.; Filosa, S.; Crispi, S. Bioactive Polyphenols and Neuromodulation: Molecular Mechanisms in Neurodegeneration. Int. J. Mol. Sci. 2020, 21, 2564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Obrador, E.; Salvador-Palmer, R.; Pellicer, B.; López-Blanch, R.; Sirerol, J.A.; Villaescusa, J.I.; Montoro, A.; Dellinger, R.W.; Estrela, J.M. Combination of natural polyphenols with a precursor of NAD(+) and a TLR2/6 ligand lipopeptide protects mice against lethal γ radiation. J. Adv. Res. 2022. [Google Scholar] [CrossRef] [PubMed]
  3. Zhou, P.; Zhao, X.N.; Ma, Y.Y.; Tang, T.J.; Wang, S.S.; Wang, L.; Huang, J.L. Virtual screening analysis of natural flavonoids as trimethylamine (TMA)-lyase inhib-itors for coronary heart disease. J. Food Biochem. 2022, e14376. [Google Scholar]
  4. Curti, V.; Di Lorenzo, A.; Dacrema, M.; Xiao, J.; Nabavi, S.M.; Daglia, M. In Vitro polyphenol effects on apoptosis: An update of literature data. Semin. Cancer Biol. 2017, 46, 119–131. [Google Scholar] [CrossRef] [PubMed]
  5. Deng, J.; Feng, X.; Zhou, L.; He, C.; Li, H.; Xia, J.; Ge, Y.; Zhao, Y.; Song, C.; Chen, L.; et al. Heterophyllin B, a cyclopeptide from Pseudostellaria heterophylla, improves memory via immunomodulation and neurite regeneration in i.c.v.Aβ-induced mice. Food Res. Int. 2022, 158, 111576. [Google Scholar] [CrossRef] [PubMed]
  6. Ge, S.; Duo, L.; Wang, J.; Zhula, G.; Yang, J.; Li, Z.; Tu, Y. A unique understanding of traditional medicine of pomegranate, Punica granatum L. and its current research status. J. Ethnopharmacol. 2021, 271, 113877. [Google Scholar] [CrossRef]
  7. Zhang, B.; Zhang, Y.; Liu, X.; Zhao, C.; Yin, J.; Li, X.; Zhang, X.; Wang, J.; Wang, S. Distinctive anti-inflammatory effects of resveratrol, dihydroresveratrol, and 3-(4-hydroxyphenyl)-propionic acid on DSS-induced colitis in pseudo-germ-free mice. Food Chem. 2023, 400, 133904. [Google Scholar] [CrossRef]
  8. Wu, Z.; Huang, S.; Li, T.; Li, N.; Han, D.; Zhang, B.; Xu, Z.Z.; Zhang, S.; Pang, J.; Wang, S.; et al. Gut microbiota from green tea polyphenol-dosed mice improves intestinal epithelial home-ostasis and ameliorates experimental colitis. Microbiome 2021, 9, 184. [Google Scholar] [CrossRef]
  9. Hoseini, A.; Namazi, G.; Farrokhian, A.; Reiner, Ž.; Aghadavod, E.; Bahmani, F.; Asemi, Z. The effects of resveratrol on metabolic status in patients with type 2 diabetes mellitus and coronary heart disease. Food Funct. 2019, 10, 6042–6051. [Google Scholar] [CrossRef]
  10. Singh, A.P.; Singh, R.; Verma, S.S.; Rai, V.; Kaschula, C.H.; Maiti, P.; Gupta, S.C. Health benefits of resveratrol: Evidence from clinical studies. Med. Res. Rev. 2019, 39, 1851–1891. [Google Scholar] [CrossRef]
  11. Zhang, Z.; Zhang, X.; Bi, K.; He, Y.; Yan, W.; Yang, C.S.; Zhang, J. Potential protective mechanisms of green tea polyphenol EGCG against COVID-19. Trends Food Sci. Technol. 2021, 114, 11–24. [Google Scholar] [CrossRef] [PubMed]
  12. Luo, J.; Li, Z.; Wang, J.; Weng, Q.; Chen, S.; Hu, M. Antifungal Activity of Isoliquiritin and Its Inhibitory Effect against Peronophythora litchi Chen through a Membrane Damage Mechanism. Molecules 2016, 21, 237. [Google Scholar] [CrossRef] [PubMed]
  13. Najjar, R.; Turner, C.; Wong, B.; Feresin, R. Berry-Derived Polyphenols in Cardiovascular Pathologies: Mechanisms of Disease and the Role of Diet and Sex. Nutrients 2021, 13, 387. [Google Scholar] [CrossRef] [PubMed]
  14. Slika, H.; Mansour, H.; Wehbe, N.; Nasser, S.A.; Iratni, R.; Nasrallah, G.; Shaito, A.; Ghaddar, T.; Kobeissy, F.; Eid, A.H. Therapeutic potential of flavonoids in cancer: ROS-mediated mechanisms. Biomed. Pharmacother. 2022, 146, 112442. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, Z.; Shi, J.; Nice, E.C.; Huang, C.; Shi, Z. The Multifaceted Role of Flavonoids in Cancer Therapy: Leveraging Autophagy with a Double-Edged Sword. Antioxidants 2021, 10, 1138. [Google Scholar] [CrossRef] [PubMed]
  16. Ku, Y.S.; Ng, M.S.; Cheng, S.S.; Lo, A.W.; Xiao, Z.; Shin, T.S.; Chung, G.; Lam, H.M. Understanding the Composition, Biosynthesis, Accumulation and Transport of Fla-vonoids in Crops for the Promotion of Crops as Healthy Sources of Flavonoids for Human Consumption. Nutrients 2020, 12, 1717. [Google Scholar] [CrossRef] [PubMed]
  17. Najjar, R.; Feresin, R. Protective Role of Polyphenols in Heart Failure: Molecular Targets and Cellular Mechanisms Underlying Their Therapeutic Potential. Int. J. Mol. Sci. 2021, 22, 1668. [Google Scholar] [CrossRef]
  18. Ngamsamer, C.; Sirivarasai, J.; Sutjarit, N. The Benefits of Anthocyanins against Obesity-Induced Inflammation. Biomolecules 2022, 12, 852. [Google Scholar] [CrossRef]
  19. Chialva, C.; Blein, T.; Crespi, M.; Lijavetzky, D. Insights into long non-coding RNA regulation of anthocyanin carrot root pigmentation. Sci. Rep. 2021, 11, 4093. [Google Scholar] [CrossRef]
  20. Kumar, N.; Goel, N. Phenolic acids: Natural versatile molecules with promising therapeutic applications. Biotechnol. Rep. 2019, 24, e00370. [Google Scholar] [CrossRef]
  21. Obidiegwu, J.E.; Lyons, J.B.; Chilaka, C.A. The Dioscorea Genus (Yam)—An Appraisal of Nutritional and Therapeutic Potentials. Foods 2020, 9, 1304. [Google Scholar] [CrossRef] [PubMed]
  22. Collins, A.E.; Saleh, T.M.; Kalisch, B.E. Naturally Occurring Antioxidant Therapy in Alzheimer’s Disease. Antioxidants 2022, 11, 213. [Google Scholar] [CrossRef]
  23. Montenegro-Landívar, M.F.; Tapia-Quirós, P.; Vecino, X.; Reig, M.; Granados, M.; Farran, A.; Cortina, J.L.; Saurina, J.; Valderrama, C. Recovery of Natural Polyphenols from Spinach and Orange By-Products by Pressure-Driven Membrane Processes. Membranes 2022, 12, 669. [Google Scholar] [CrossRef]
  24. Oliva, E.; Fanti, F.; Palmieri, S.; Viteritti, E.; Eugelio, F.; Pepe, A.; Compagnone, D.; Sergi, M. Predictive Multi Experiment Approach for the Determination of Conjugated Phenolic Compounds in Vegetal Matrices by Means of LC-MS/MS. Molecules 2022, 27, 3089. [Google Scholar] [CrossRef] [PubMed]
  25. Zeiss, D.R.; Mhlongo, M.I.; Tugizimana, F.; Steenkamp, P.A.; Dubery, I.A. Metabolomic Profiling of the Host Response of Tomato (Solanum lycopersicum) Following Infection by Ralstonia solanacearum. Int. J. Mol. Sci. 2019, 20, 3945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zálešák, F.; Bon, D.J.-Y.D.; Pospíšil, J. Lignans and Neolignans: Plant secondary metabolites as a reservoir of biologically active substances. Pharmacol. Res. 2019, 146, 104284. [Google Scholar] [CrossRef] [PubMed]
  27. Li, B.; Fan, S.; Hu, J.; Ma, Y.; Feng, Y.; Wang, F.; Wang, X.; Niu, L. Phytochemical Analysis Using UPLC-MS/MS Combined with Network Pharmacology Methods to Explore the Biomarkers for the Quality Control of Lingguizhugan Decoction. Evidence-Based Complement. Altern. Med. 2021, 2021, 7849032. [Google Scholar] [CrossRef]
  28. Wang, L.-X.; Wang, H.-L.; Huang, J.; Chu, T.-Z.; Peng, C.; Zhang, H.; Chen, H.-L.; Xiong, Y.-A.; Tan, Y.-Z. Review of lignans from 2019 to 2021: Newly reported compounds, diverse activities, structure-activity relationships and clinical applications. Phytochemistry 2022, 202, 113326. [Google Scholar] [CrossRef]
  29. Qayyum, Z.; Noureen, F.; Khan, M.; Khan, M.; Haider, G.; Munir, F.; Gul, A.; Amir, R. Identification and Expression Analysis of Stilbene Synthase Genes in Arachis hypogaea in Response to Methyl Jasmonate and Salicylic Acid Induction. Plants 2022, 11, 1776. [Google Scholar] [CrossRef]
  30. Cosme, P.; Rodríguez, A.B.; Espino, J.; Garrido, M. Plant Phenolics: Bioavailability as a Key Determinant of Their Potential Health-Promoting Applications. Antioxidants 2020, 9, 1263. [Google Scholar] [CrossRef]
  31. Sharifi-Rad, J.; Quispe, C.; Zam, W.; Kumar, M.; Cardoso, S.M.; Pereira, O.R.; Ademiluyi, A.O.; Adeleke, O.; Moreira, A.C.; Živković, J.; et al. Phenolic Bioactives as Antiplatelet Aggregation Factors: The Pivotal Ingredients in Maintaining Cardiovascular Health. Oxidative Med. Cell. Longev. 2021, 2021, 2195902. [Google Scholar] [CrossRef] [PubMed]
  32. Aryal, S.; Skinner, T.; Bridges, B.; Weber, J.T. The Pathology of Parkinson’s Disease and Potential Benefit of Dietary Polyphenols. Molecules 2020, 25, 4382. [Google Scholar] [CrossRef] [PubMed]
  33. Malaguarnera, L. Influence of Resveratrol on the Immune Response. Nutrients 2019, 11, 946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hayes, J.D.; Dinkova-Kostova, A.T.; Tew, K.D. Oxidative Stress in Cancer. Cancer Cell 2020, 38, 167–197. [Google Scholar] [CrossRef]
  35. Sies, H.; Jones, D.P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 2020, 21, 363–383. [Google Scholar] [CrossRef]
  36. Kalyanaraman, B.; Cheng, G.; Hardy, M.; Ouari, O.; Bennett, B.; Zielonka, J. Teaching the basics of reactive oxygen species and their relevance to cancer biology: Mitochondrial reactive oxygen species detection, redox signaling, and targeted therapies. Redox Biol. 2018, 15, 347–362. [Google Scholar] [CrossRef]
  37. Heurtaux, T.; Bouvier, D.S.; Benani, A.; Romero, S.H.; Frauenknecht, K.B.M.; Mittelbronn, M.; Sinkkonen, L. Normal and Pathological NRF2 Signalling in the Central Nervous System. Antioxidants 2022, 11, 1426. [Google Scholar] [CrossRef]
  38. Liu, X.; Zhang, Y.; Zhuang, L.; Olszewski, K.; Gan, B. NADPH debt drives redox bankruptcy: SLC7A11/xCT-mediated cystine uptake as a double-edged sword in cellular redox regulation. Genes Dis. 2021, 8, 731–745. [Google Scholar] [CrossRef]
  39. Qin, Z.; Xiang, C.; Zhong, F.; Liu, Y.; Dong, Q.; Li, K.; Shi, W.; Ding, C.; Qin, L.; He, F. Transketolase (TKT) activity and nuclear localization promote hepatocellular carcinoma in a metabolic and a non-metabolic manner. J. Exp. Clin. Cancer Res. 2019, 38, 1–21. [Google Scholar] [CrossRef] [Green Version]
  40. Rojo de la Vega, M.; Chapman, E.; Zhang, D.D. NRF2 and the Hallmarks of Cancer. Cancer Cell 2018, 34, 21–43. [Google Scholar] [CrossRef]
  41. Tang, Y.-A.; Chen, Y.-F.; Bao, Y.; Mahara, S.; Yatim, S.M.J.M.; Oguz, G.; Lee, P.L.; Feng, M.; Cai, Y.; Tan, E.Y.; et al. Hypoxic tumor microenvironment activates GLI2 via HIF-1α and TGF-β2 to promote chemoresistance in colorectal cancer. Proc. Natl. Acad. Sci. USA 2018, 115, E5990–E5999. [Google Scholar] [CrossRef] [PubMed]
  42. Perillo, B.; Di Donato, M.; Pezone, A.; Di Zazzo, E.; Giovannelli, P.; Galasso, G.; Castoria, G.; Migliaccio, A. ROS in cancer therapy: The bright side of the moon. Exp. Mol. Med. 2020, 52, 192–203. [Google Scholar] [CrossRef]
  43. Jin, P.; Jiang, J.; Zhou, L.; Huang, Z.; Nice, E.C.; Huang, C.; Fu, L. Mitochondrial adaptation in cancer drug resistance: Prevalence, mechanisms, and management. J. Hematol. Oncol. 2022, 15, 97. [Google Scholar] [CrossRef]
  44. Peng, L.; Jiang, J.; Chen, H.N.; Zhou, L.; Huang, Z.; Qin, S.; Jin, P.; Luo, M.; Li, B.; Shi, J.; et al. Redox-sensitive cyclophilin A elicits chemoresistance through realigning cellular oxi-dative status in colorectal cancer. Cell Rep. 2021, 37, 110069. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, Z.; Qin, S.; Chen, Y.; Zhou, L.; Yang, M.; Tang, Y.; Zuo, J.; Zhang, J.; Mizokami, A.; Nice, E.C.; et al. Inhibition of NPC1L1 disrupts adaptive responses of drug-tolerant persister cells to chemotherapy. EMBO Mol. Med. 2022, 14, e14903. [Google Scholar] [CrossRef] [PubMed]
  46. Jin, P.; Jiang, J.; Zhou, L.; Huang, Z.; Qin, S.; Chen, H.N.; Peng, L.; Zhang, Z.; Li, B.; Luo, M.; et al. Disrupting metformin adaptation of liver cancer cells by targeting the TOMM34/ATP5B axis. EMBO Mol. Med. 2022, 14, e16082. [Google Scholar] [CrossRef]
  47. Abdelmohsen, U.R.; Balasubramanian, S.; Oelschlaeger, A.T.; Grkovic, T.; Pham, N.B.; Quinn, R.J.; Hentschel, U. Potential of marine natural products against drug-resistant fungal, viral, and parasitic infections. Lancet Infect. Dis. 2017, 17, e30–e41. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, X.; Jia, W.; Zhu, L.; Mao, L.; Zhang, Y. Recent advances in heterocyclic aromatic amines: An update on food safety and hazardous control from food processing to dietary intake. Compr. Rev. Food Sci. Food Saf. 2020, 19, 124–148. [Google Scholar] [CrossRef] [Green Version]
  49. Mao, X.-Y.; Jin, M.-Z.; Chen, J.-F.; Zhou, H.-H.; Jin, W.-L. Live or let die: Neuroprotective and anti-cancer effects of nutraceutical antioxidants. Pharmacol. Ther. 2018, 183, 137–151. [Google Scholar] [CrossRef]
  50. Olcha, P.; Winiarska-Mieczan, A.; Kwiecień, M.; Nowakowski, Ł.; Miturski, A.; Semczuk, A.; Kiczorowska, B.; Gałczyński, K. Antioxidative, Anti-Inflammatory, Anti-Obesogenic, and Antidia-betic Properties of Tea Polyphenols-The Positive Impact of Regular Tea Consumption as an Element of Prophylaxis and Pharmacotherapy Support in Endometrial Cancer. Int. J. Mol. Sci. 2022, 23, 6703. [Google Scholar] [CrossRef]
  51. Wang, M.; Jiang, S.; Zhou, L.; Yu, F.; Ding, H.; Li, P.; Zhou, M.; Wang, K. Potential Mechanisms of Action of Curcumin for Cancer Prevention: Focus on Cellular Signaling Pathways and miRNAs. Int. J. Biol. Sci. 2019, 15, 1200–1214. [Google Scholar] [CrossRef] [PubMed]
  52. Purpura, V.; Benedetti, S.; Bondioli, E.; Scarpellini, F.; Giacometti, A.; Albertini, M.C.; Melandri, D. The Use of Quercetin to Improve the Antioxidant and Regenerative Properties of Frozen or Cryopreserved Human Amniotic Membrane. Antioxidants 2022, 11, 1250. [Google Scholar] [CrossRef] [PubMed]
  53. Vieira, I.R.S.; Conte-Junior, C.A. Nano-delivery systems for food bioactive compounds in cancer: Prevention, therapy, and clinical applications. Crit. Rev. Food Sci. Nutr. 2022, 1–26. [Google Scholar] [CrossRef]
  54. Bangar, S.P.; Chaudhary, V.; Sharma, N.; Bansal, V.; Ozogul, F.; Lorenzo, J.M. Kaempferol: A flavonoid with wider biological activities and its applications. Crit. Rev. Food Sci. Nutr. 2022, 1–25. [Google Scholar] [CrossRef] [PubMed]
  55. Munot, N.; Kandekar, U.; Giram, P.S.; Khot, K.; Patil, A.; Cavalu, S. A Comparative Study of Quercetin-Loaded Nanococh-leates and Liposomes: Formulation, Characterization, Assessment of Degradation and In Vitro Anticancer Potential. Pharmaceutics 2022, 14, 1601. [Google Scholar] [CrossRef] [PubMed]
  56. Farhan, M.; Rizvi, A. Understanding the Prooxidant Action of Plant Polyphenols in the Cellular Microenvironment of Malignant Cells: Role of Copper and Therapeutic Implications. Front. Pharmacol. 2022, 13, 929853. [Google Scholar] [CrossRef] [PubMed]
  57. Scuto, M.; Ontario, M.L.; Salinaro, A.T.; Caligiuri, I.; Rampulla, F.; Zimbone, V.; Modafferi, S.; Rizzolio, F.; Canzonieri, V.; Calabrese, E.J.; et al. Redox modulation by plant polyphenols targeting vitagenes for chemopre-vention and therapy: Relevance to novel anti-cancer interventions and mini-brain organoid technology. Free. Radic. Biol. Med. 2022, 179, 59–75. [Google Scholar] [CrossRef]
  58. Lendzion, K.; Gornowicz, A.; Strawa, J.W.; Bielawska, K.; Czarnomysy, R.; Popławska, B.; Bielawski, K.; Tomczyk, M.; Miltyk, W.; Bielawska, A. LC-PDA-MS and GC-MS Analysis of Scorzonera hispanica Seeds and Their Effects on Human Breast Cancer Cell Lines. Int. J. Mol. Sci. 2022, 23, 11584. [Google Scholar] [CrossRef]
  59. Guo, Y.; Sun, Q.; Wu, F.G.; Dai, Y.; Chen, X. Polyphenol-Containing Nanoparticles: Synthesis, Properties, and Therapeutic Delivery. Adv. Mater. 2021, 33, e2007356. [Google Scholar] [CrossRef]
  60. Ju, Y.; Liao, H.; Richardson, J.J.; Guo, J.; Caruso, F. Nanostructured particles assembled from natural building blocks for advanced therapies. Chem. Soc. Rev. 2022, 51, 4287–4336. [Google Scholar] [CrossRef]
  61. Mani, S.; Swargiary, G.; Ralph, S.J. Targeting the redox imbalance in mitochondria: A novel mode for cancer therapy. Mitochondrion 2022, 62, 50–73. [Google Scholar] [CrossRef] [PubMed]
  62. Cheung, E.C.; DeNicola, G.M.; Nixon, C.; Blyth, K.; Labuschagne, C.F.; Tuveson, D.A.; Vousden, K.H. Dynamic ROS Control by TIGAR Regulates the Initiation and Progression of Pancreatic Cancer. Cancer Cell 2020, 37, 168–182.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Reczek, C.R.; Birsoy, K.; Kong, H.; Martínez-Reyes, I.; Wang, T.; Gao, P.; Sabatini, D.M.; Chandel, N.S. A CRISPR screen identifies a pathway required for paraquat-induced cell death. Nat. Chem. Biol. 2017, 13, 1274–1279. [Google Scholar] [CrossRef] [Green Version]
  64. Cencioni, C.; Comunanza, V.; Middonti, E.; Vallariello, E.; Bussolino, F. The role of redox system in metastasis formation. Angiogenesis 2021, 24, 435–450. [Google Scholar] [CrossRef] [PubMed]
  65. Dodson, M.; Castro-Portuguez, R.; Zhang, D.D. NRF2 plays a critical role in mitigating lipid peroxidation and ferroptosis. Redox Biol. 2019, 23, 101107. [Google Scholar] [CrossRef]
  66. Paller, C.J.; Zhou, X.C.; Heath, E.I.; Taplin, M.-E.; Mayer, T.; Stein, M.N.; Bubley, G.J.; Pili, R.; Hudson, T.; Kakarla, R.; et al. Muscadine Grape Skin Extract (MPX) in Men with Biochemically Recurrent Prostate Cancer: A Randomized, Multicenter, Placebo-Controlled Clinical Trial. Clin. Cancer Res. 2018, 24, 306–315. [Google Scholar] [CrossRef] [Green Version]
  67. vila-Gálvez, M.; González-Sarrías, A.; Martínez-Díaz, F.; Abellán, B.; Martínez-Torrano, A.J.; Fernández-López, A.J.; Giménez-Bastida, J.A.; Espín, J.C. Disposition of Dietary Polyphenols in Breast Cancer Pa-tients’ Tumors, and Their Associated Anticancer Activity: The Particular Case of Curcumin. Mol. Nutr. Food Res. 2021, 65, e2100163. [Google Scholar] [CrossRef]
  68. Henning, S.M.; Wang, P.; Lee, R.-P.; Trang, A.; Husari, G.; Yang, J.; Grojean, E.M.; Ly, A.; Hsu, M.; Heber, D.; et al. Prospective randomized trial evaluating blood and prostate tissue concentrations of green tea polyphenols and quercetin in men with prostate cancer. Food Funct. 2020, 11, 4114–4122. [Google Scholar] [CrossRef]
  69. Liu, Y.; Yu, H.; Zhang, C.; Cheng, Y.; Hu, L.; Meng, X.; Zhao, Y. Protective effects of berberine on radiation-induced lung injury via intercellular adhesion molecular-1 and transforming growth factor-beta-1 in patients with lung cancer. Eur. J. Cancer 2008, 44, 2425–2432. [Google Scholar] [CrossRef]
  70. Tuli, H.S.; Mittal, S.; Aggarwal, D.; Parashar, G.; Parashar, N.C.; Upadhyay, S.K.; Barwal, T.S.; Jain, A.; Kaur, G.; Savla, R.; et al. Path of Silibinin from diet to medicine: A dietary polyphenolic flavonoid having potential anti-cancer therapeutic significance. Semin. Cancer Biol. 2021, 73, 196–218. [Google Scholar] [CrossRef]
  71. Rahaiee, S.; Assadpour, E.; Esfanjani, A.F.; Silva, A.S.; Jafari, S.M. Application of nano/microencapsulated phenolic compounds against cancer. Adv. Colloid Interface Sci. 2020, 279, 102153. [Google Scholar] [CrossRef] [PubMed]
  72. Wu, M.-F.; Huang, Y.-H.; Chiu, L.-Y.; Cherng, S.-H.; Sheu, G.-T.; Yang, T.-Y. Curcumin Induces Apoptosis of Chemoresistant Lung Cancer Cells via ROS-Regulated p38 MAPK Phosphorylation. Int. J. Mol. Sci. 2022, 23, 8248. [Google Scholar] [CrossRef] [PubMed]
  73. Kostrzewa, T.; Wołosewicz, K.; Jamrozik, M.; Drzeżdżon, J.; Siemińska, J.; Jacewicz, D.; Górska-Ponikowska, M.; Kołaczkowski, M.; Łaźny, R.; Kuban-Jankowska, A. Curcumin and Its New Derivatives: Correlation between Cytotoxicity against Breast Cancer Cell Lines, Degradation of PTP1B Phosphatase and ROS Generation. Int. J. Mol. Sci. 2021, 22, 10368. [Google Scholar] [CrossRef] [PubMed]
  74. Li, G.; Fang, S.; Shao, X.; Li, Y.; Tong, Q.; Kong, B.; Chen, L.; Wang, Y.; Yang, J.; Yu, H.; et al. Curcumin Reverses NNMT-Induced 5-Fluorouracil Resistance via Increasing ROS and Cell Cycle Arrest in Colorectal Cancer Cells. Biomolecules 2021, 11, 1295. [Google Scholar] [CrossRef]
  75. Wang, T.; Wu, X.; Al Rudaisat, M.; Song, Y.; Cheng, H. Curcumin induces G2/M arrest and triggers autophagy, ROS gen-eration and cell senescence in cervical cancer cells. J. Cancer 2020, 11, 6704–6715. [Google Scholar] [CrossRef]
  76. Lin, X.; Wang, L.; Zhao, L.; Zhu, Z.; Chen, T.; Chen, S.; Tao, Y.; Zeng, T.; Zhong, Y.; Sun, H.; et al. Curcumin micelles suppress gastric tumor cell growth by upregulating ROS generation, disrupting redox equilibrium and affecting mitochondrial bioenergetics. Food Funct. 2020, 11, 4146–4159. [Google Scholar] [CrossRef]
  77. Mahgoub, S.; Hashad, N.; Ali, S.; Ibrahim, R.; Said, A.M.; Moharram, F.A.; Mady, M. Polyphenolic Profile of Callistemon viminalis Aerial Parts: Antioxidant, Anticancer and In Silico 5-LOX Inhibitory Evaluations. Molecules 2021, 26, 2481. [Google Scholar] [CrossRef]
  78. Del Mar Rivas-Chacón, L.; Yanes-Díaz, J.; de Lucas, B.; Riestra-Ayora, J.I.; Madrid-García, R.; Sanz-Fernández, R.; Sánchez-Rodríguez, C. Preventive Effect of Cocoa Flavonoids via Suppression of Oxidative Stress-Induced Apoptosis in Auditory Senescent Cells. Antioxidants 2022, 11, 1450. [Google Scholar] [CrossRef]
  79. Li, Z.; Zhang, S.; Xue, J.; Mu, B.; Song, H.; Liu, Y. Exogenous Melatonin Treatment Induces Disease Resistance against Botrytis cinerea on Post-Harvest Grapes by Activating Defence Responses. Foods 2022, 11, 2231. [Google Scholar] [CrossRef]
  80. Liu, Z.; Jiang, F.; Mo, Y.; Liao, H.; Chen, P.; Zhang, H. Effects of Ethanol Treatment on Storage Quality and Antioxidant System of Postharvest Papaya. Front. Plant Sci. 2022, 13, 856499. [Google Scholar] [CrossRef]
  81. Khalil Alyahya, H.; Subash-Babu, P.; Mohammad Salamatullah, A.; Hayat, K.; Albader, N.; Alkaltham, M.S.; Ahmed, M.A.; Arzoo, S.; Bourhia, M. Quantification of Chlorogenic Acid and Vanillin from Coffee Peel Extract and its Effect on α-Amylase Activity, Immunoregulation, Mitochondrial Oxidative Stress, and Tumor Suppressor Gene Expression Levels in H(2)O(2)-Induced Human Mesenchymal Stem Cells. Front Pharmacol. 2021, 12, 760242. [Google Scholar] [CrossRef]
  82. Heydarzadeh, S.; Kia, S.K.; Zarkesh, M.; Pakizehkar, S.; Hosseinzadeh, S.; Hedayati, M. The Cross-Talk between Polyphenols and the Target Enzymes Related to Oxidative Stress-Induced Thyroid Cancer. Oxidative Med. Cell. Longev. 2022, 2022, 2724324. [Google Scholar] [CrossRef]
  83. Oláhová, M.; Veal, E.A. A peroxiredoxin, PRDX-2, is required for insulin secretion and insulin/IIS-dependent regulation of stress resistance and longevity. Aging Cell 2015, 14, 558–568. [Google Scholar] [CrossRef]
  84. Qi, Z.; Ji, H.; Le, M.; Li, H.; Wieland, A.; Bauer, S.; Liu, L.; Wink, M.; Herr, I. Sulforaphane promotes C. elegans longevity and healthspan via DAF-16/DAF-2 insulin/IGF-1 signaling. Aging 2021, 13, 1649–1670. [Google Scholar] [CrossRef]
  85. Kittimongkolsuk, P.; Roxo, M.; Li, H.; Chuchawankul, S.; Wink, M.; Tencomnao, T. Extracts of the Tiger Milk Mushroom (Lignosus rhinocerus) Enhance Stress Resistance and Extend Lifespan in Caenorhabditis elegans via the DAF-16/FoxO Sig-naling Pathway. Pharmaceuticals 2021, 14, 93. [Google Scholar] [CrossRef]
  86. Li, H.; Roxo, M.; Cheng, X.; Zhang, S.; Cheng, H.; Wink, M. Pro-oxidant and lifespan extension effects of caffeine and related methylxanthines in Caenorhabditis elegans. Food Chem. X 2019, 1, 100005. [Google Scholar] [CrossRef]
  87. Duangjan, C.; Rangsinth, P.; Gu, X.; Zhang, S.; Wink, M.; Tencomnao, T. Glochidion zeylanicum leaf extracts exhibit lifespan extending and oxidative stress resistance properties in Caenorhabditis elegans via DAF-16/FoxO and SKN-1/Nrf-2 signaling pathways. Phytomedicine 2019, 64, 153061. [Google Scholar] [CrossRef]
  88. Zhai, Y.; Sun, J.; Sun, C.; Zhao, H.; Li, X.; Yao, J.; Su, J.; Xu, X.; Xu, X.; Hu, J.; et al. Total flavonoids from the dried root of Tetrastigma hemsleyanum Diels et Gilg inhibit colo-rectal cancer growth through PI3K/AKT/mTOR signaling pathway. Phytother. Res. 2022, 36, 4263–4277. [Google Scholar] [CrossRef]
  89. Zhou, Y.; Chen, H.; Xue, J.; Yuan, J.; Cai, Z.; Wu, N.; Zou, L.; Yin, S.; Yang, W.; Liu, X.; et al. Qualitative Analysis and Componential Differences of Chemical Constituents in Lysim-achiae Herba from Different Habitats (Sichuan Basin) by UFLC-Triple TOF-MS/MS. Molecules 2022, 27, 4600. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, X.; Wu, Y.; Chen, G.; Yue, W.; Liang, Q.; Wu, Q. Optimisation of ultrasound assisted extraction of phenolic compounds from Sparganii rhizoma with response surface methodology. Ultrason. Sonochem. 2013, 20, 846–854. [Google Scholar] [CrossRef]
  91. Harrath, A.H.; Jalouli, M.; Oueslati, M.H.; Farah, M.A.; Feriani, A.; Aldahmash, W.; Aldawood, N.; Al-Anazi, K.; Falodah, F.; Swelum, A.; et al. The flavonoid, kaempferol-3-O-apiofuranosyl-7-O-rhamnopyranosyl, as a potential therapeutic agent for breast cancer with a promoting effect on ovarian function. Phytotherapy Res. 2021, 35, 6170–6180. [Google Scholar] [CrossRef]
  92. Felice, M.R.; Maugeri, A.; De Sarro, G.; Navarra, M.; Barreca, D. Molecular Pathways Involved in the Anti-Cancer Activity of Flavonols: A Focus on Myricetin and Kaempferol. Int. J. Mol. Sci. 2022, 23, 4411. [Google Scholar] [CrossRef]
  93. Amjad, E.; Sokouti, B.; Asnaashari, S. A systematic review of anti-cancer roles and mechanisms of kaempferol as a natural compound. Cancer Cell Int. 2022, 22, 260. [Google Scholar] [CrossRef]
  94. Kluska, M.; Juszczak, M.; Żuchowski, J.; Stochmal, A.; Woźniak, K. Effect of Kaempferol and Its Glycoside Derivatives on Antioxidant Status of HL-60 Cells Treated with Etoposide. Molecules 2022, 27, 333. [Google Scholar] [CrossRef]
  95. Wu, P.; Meng, X.; Zheng, H.; Zeng, Q.; Chen, T.; Wang, W.; Zhang, X.; Su, J. Kaempferol Attenuates ROS-Induced Hemolysis and the Molecular Mechanism of Its Induction of Apoptosis on Bladder Cancer. Molecules 2018, 23, 2592. [Google Scholar] [CrossRef] [Green Version]
  96. Yang, Y.; Wang, Y.; Wang, T.; Jiang, X.; Wang, L. Screening active components of modified Xiaoyao powder as NRF2 ago-nists. Cell Biochem. Funct. 2017, 35, 518–526. [Google Scholar] [CrossRef]
  97. Zeng, J.; Xu, H.; Fan, P.-Z.; Xie, J.; He, J.; Yu, J.; Gu, X.; Zhang, C.-J. Kaempferol blocks neutrophil extracellular traps formation and reduces tumour metastasis by inhibiting ROS-PAD4 pathway. J. Cell. Mol. Med. 2020, 24, 7590–7599. [Google Scholar] [CrossRef]
  98. Park, S.J.; Kim, D.W.; Lim, S.R.; Sung, J.; Kim, T.H.; Min, I.S.; Choi, C.H.; Lee, S.J. Kaempferol Blocks the Skin Fibroblastic Interleukin 1β Expression and Cytotoxicity Induced by 12-O-tetradecanoylphorbol-13-acetate by Suppressing c-Jun N-terminal Kinase. Nutrients 2021, 13, 3079. [Google Scholar] [CrossRef]
  99. Liu, Z.; Yao, X.; Sun, B.; Jiang, W.; Liao, C.; Dai, X.; Chen, Y.; Chen, J.; Ding, R. Pretreatment with kaempferol attenuates microglia-mediate neuroinflammation by inhibiting MAPKs-NF-κB signaling pathway and pyroptosis after secondary spinal cord injury. Free. Radic. Biol. Med. 2021, 168, 142–154. [Google Scholar] [CrossRef]
  100. Dong, S.; Liang, S.; Cheng, Z.; Zhang, X.; Luo, L.; Li, L.; Zhang, W.; Li, S.; Xu, Q.; Zhong, M.; et al. ROS/PI3K/Akt and Wnt/β-catenin signalings activate HIF-1α-induced metabolic re-programming to impart 5-fluorouracil resistance in colorectal cancer. J. Exp. Clin. Cancer Res. 2022, 41, 15. [Google Scholar] [CrossRef] [PubMed]
  101. Zhang, D.; Zhou, Q.; Huang, D.; He, L.; Zhang, H.; Hu, B.; Peng, H.; Ren, D. ROS/JNK/c-Jun axis is involved in oridonin-induced caspase-dependent apoptosis in human colorectal cancer cells. Biochem. Biophys. Res. Commun. 2019, 513, 594–601. [Google Scholar] [CrossRef] [PubMed]
  102. Riahi-Chebbi, I.; Souid, S.; Othman, H.; Haoues, M.; Karoui, H.; Morel, A.; Srairi-Abid, N.; Essafi, M.; Essafi-Benkhadir, K. The Phenolic compound Kaempferol overcomes 5-fluorouracil resistance in human resistant LS174 colon cancer cells. Sci. Rep. 2019, 9, 195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Park, J.; Lee, G.-E.; An, H.-J.; Lee, C.-J.; Cho, E.S.; Kang, H.C.; Lee, J.Y.; Lee, H.S.; Choi, J.-S.; Kim, D.J.; et al. Kaempferol sensitizes cell proliferation inhibition in oxaliplatin-resistant colon cancer cells. Arch. Pharmacal Res. 2021, 44, 1091–1108. [Google Scholar] [CrossRef] [PubMed]
  104. Hu, Y.; Wang, S.; Wu, X.; Zhang, J.; Chen, R.; Chen, M.; Wang, Y. Chinese herbal medicine-derived compounds for cancer therapy: A focus on hepatocellular carcinoma. J. Ethnopharmacol. 2013, 149, 601–612. [Google Scholar] [CrossRef] [PubMed]
  105. Ma, B.-N.; Li, X.-J. Resveratrol extracted from Chinese herbal medicines: A novel therapeutic strategy for lung diseases. Chin. Herb. Med. 2020, 12, 349–358. [Google Scholar] [CrossRef] [PubMed]
  106. Rodriguez-Garcia, A.; Hevia, D.; Mayo, J.C.; Gonzalez-Menendez, P.; Coppo, L.; Lu, J.; Holmgren, A.; Sainz, R.M. Thioredoxin 1 modulates apoptosis induced by bioactive compounds in prostate cancer cells. Redox Biol. 2017, 12, 634–647. [Google Scholar] [CrossRef]
  107. Dolinsky, V.W.; Chan, A.Y.; Frayne, I.R.; Light, P.E.; Rosiers, C.D.; Dyck, J.R. Resveratrol Prevents the Prohypertrophic Effects of Oxidative Stress on LKB1. Circulation 2009, 119, 1643–1652. [Google Scholar] [CrossRef] [Green Version]
  108. Zhao, Y.; Song, W.; Wang, Z.; Wang, Z.; Jin, X.; Xu, J.; Bai, L.; Li, Y.; Cui, J.; Cai, L. Resveratrol attenuates testicular apoptosis in type 1 diabetic mice: Role of Akt-mediated Nrf2 activation and p62-dependent Keap1 degradation. Redox Biol. 2018, 14, 609–617. [Google Scholar] [CrossRef]
  109. Elshaer, M.; Chen, Y.; Wang, X.J.; Tang, X. Resveratrol: An overview of its anti-cancer mechanisms. Life Sci. 2018, 207, 340–349. [Google Scholar] [CrossRef]
  110. Li, Y.; Guo, Y.; Feng, Z.; Bergan, R.; Li, B.; Qin, Y.; Zhao, L.; Zhang, Z.; Shi, M. Involvement of the PI3K/Akt/Nrf2 Signaling Pathway in Resveratrol-Mediated Reversal of Drug Resistance in HL-60/ADR Cells. Nutr. Cancer 2019, 71, 1007–1018. [Google Scholar] [CrossRef]
  111. Singh, B.; Shoulson, R.; Chatterjee, A.; Ronghe, A.; Bhat, N.K.; Dim, D.C.; Bhat, H.K. Resveratrol inhibits estrogen-induced breast carcinogenesis through induction of NRF2-mediated protective pathways. Carcinogenesis 2014, 35, 1872–1880. [Google Scholar] [CrossRef]
  112. Zheng, Z.; Chen, Y.; Huang, J.; Deng, H.; Tang, X.; Wang, X.J. Mkp-1 is required for chemopreventive activity of butylated hydroxyanisole and resveratrol against colitis-associated colon tumorigenesis. Food Chem. Toxicol. 2019, 127, 72–80. [Google Scholar] [CrossRef] [PubMed]
  113. Sun, Y.; Zhou, Q.-M.; Lu, Y.-Y.; Zhang, H.; Chen, Q.-L.; Zhao, M.; Su, S.-B. Resveratrol Inhibits the Migration and Metastasis of MDA-MB-231 Human Breast Cancer by Reversing TGF-β1-Induced Epithelial-Mesenchymal Transition. Molecules 2019, 24, 1131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Cheng, L.; Yan, B.; Chen, K.; Jiang, Z.; Zhou, C.; Cao, J.; Qian, W.; Li, J.; Sun, L.; Ma, J.; et al. Resveratrol-Induced Downregulation of NAF-1 Enhances the Sensitivity of Pancreatic Cancer Cells to Gemcitabine via the ROS/Nrf2 Signaling Pathways. Oxidative Med. Cell. Longev. 2018, 2018, 9482018. [Google Scholar] [CrossRef]
  115. Pan, S.-Y.; Nie, Q.; Tai, H.-C.; Song, X.-L.; Tong, Y.-F.; Zhang, L.-J.; Wu, X.-W.; Lin, Z.-H.; Zhang, Y.-Y.; Ye, D.-Y.; et al. Tea and tea drinking: China’s outstanding contributions to the mankind. Chin. Med. 2022, 17, 27. [Google Scholar] [CrossRef] [PubMed]
  116. Hung, S.W.; Li, Y.; Chen, X.; Chu, K.O.; Zhao, Y.; Liu, Y.; Guo, X.; Man, G.C.; Wang, C.C. Green Tea Epigallocatechin-3-Gallate Regulates Autophagy in Male and Female Re-productive Cancer. Front Pharmacol. 2022, 13, 906746. [Google Scholar] [CrossRef]
  117. Kochman, J.; Jakubczyk, K.; Antoniewicz, J.; Mruk, H.; Janda, K. Health Benefits and Chemical Composition of Matcha Green Tea: A Review. Molecules 2020, 26, 85. [Google Scholar] [CrossRef]
  118. Dai, Y.; Yang, R.; Yan, Y.; Wu, Y.; Meng, X.; Yang, A.; Wu, Z.; Shi, L.; Li, X.; Chen, H. Digestive stability and transport ability changes of β-lactoglobulin–catechin complexes by M cell model in vitro. Front. Nutr. 2022, 9, 955135. [Google Scholar] [CrossRef]
  119. Lou, J.; Wang, W.; Zhu, L. Occurrence, Formation, and Oxidative Stress of Emerging Disinfection Byproducts, Halobenzoquinones, in Tea. Environ. Sci. Technol. 2019, 53, 11860–11868. [Google Scholar] [CrossRef]
  120. Henning, S.M.; Wang, P.; Said, J.; Magyar, C.; Castor, B.; Doan, N.; Tosity, C.; Moro, A.; Gao, K.; Li, L.; et al. Polyphenols in brewed green tea inhibit prostate tumor xenograft growth by local-izing to the tumor and decreasing oxidative stress and angiogenesis. J. Nutr. Biochem. 2012, 23, 1537–1542. [Google Scholar] [CrossRef] [Green Version]
  121. Wubetu, G.Y.; Shimada, M.; Morine, Y.; Ikemoto, T.; Ishikawa, D.; Iwahashi, S.; Yamada, S.; Saito, Y.; Arakawa, Y.; Imura, S. Epigallocatechin gallate hinders human hepatoma and colon cancer sphere formation. J. Gastroenterol. Hepatol. 2016, 31, 256–264. [Google Scholar] [CrossRef] [PubMed]
  122. Ohishi, T.; Hayakawa, S.; Miyoshi, N. Involvement of microRNA modifications in anticancer effects of major polyphenols from green tea, coffee, wine, and curry. Crit. Rev. Food Sci. Nutr. 2022, 1–32. [Google Scholar] [CrossRef] [PubMed]
  123. Alam, M.; Ali, S.; Ashraf, G.M.; Bilgrami, A.L.; Yadav, D.K.; Hassan, I. Epigallocatechin 3-gallate: From green tea to cancer therapeutics. Food Chem. 2022, 379, 132135. [Google Scholar] [CrossRef]
  124. Chiou, Y.S.; Sang, S.; Cheng, K.H.; Ho, C.T.; Wang, Y.J.; Pan, M.H. Peracetylated (−)-epigallocatechin-3-gallate (AcEGCG) po-tently prevents skin carcinogenesis by suppressing the PKD1-dependent signaling pathway in CD34+ skin stem cells and skin tumors. Carcinogenesis 2013, 34, 1315–1322. [Google Scholar] [CrossRef]
  125. Zhou, F.; Shen, T.; Duan, T.; Xu, Y.Y.; Khor, S.C.; Li, J.; Ge, J.; Zheng, Y.F.; Hsu, S.; DE Stefano, J.; et al. Antioxidant effects of lipophilic tea polyphenols on diethylnitrosa-mine/phenobarbital-induced hepatocarcinogenesis in rats. In Vivo 2014, 28, 495–503. [Google Scholar] [PubMed]
  126. Yuan, J.-H.; Li, Y.-Q.; Yang, X.-Y. Inhibition of Epigallocatechin Gallate on Orthotopic Colon Cancer by Upregulating the Nrf2-UGT1A Signal Pathway in Nude Mice. Pharmacology 2007, 80, 269–278. [Google Scholar] [CrossRef] [PubMed]
  127. Kang, Q.; Tong, Y.; Gowd, V.; Wang, M.; Chen, F.; Cheng, K.-W. Oral administration of EGCG solution equivalent to daily achievable dosages of regular tea drinkers effectively suppresses miR483-3p induced metastasis of hepatocellular carcinoma cells in mice. Food Funct. 2021, 12, 3381–3392. [Google Scholar] [CrossRef]
  128. Datta, S.; Sinha, D. EGCG maintained Nrf2-mediated redox homeostasis and minimized etoposide resistance in lung cancer cells. J. Funct. Foods 2019, 62, 103553. [Google Scholar] [CrossRef]
  129. Chen, B.; Zhang, Y.; Wang, Y.; Rao, J.; Jiang, X.; Xu, Z. Curcumin inhibits proliferation of breast cancer cells through Nrf2-mediated down-regulation of Fen1 expression. J. Steroid Biochem. Mol. Biol. 2014, 143, 11–18. [Google Scholar] [CrossRef]
  130. Rajagopal, C.; Lankadasari, M.B.; Aranjani, J.M.; Harikumar, K. Targeting oncogenic transcription factors by polyphenols: A novel approach for cancer therapy. Pharmacol. Res. 2018, 130, 273–291. [Google Scholar] [CrossRef]
  131. Yao, J.; Zhao, L.; Zhao, Q.; Zhao, Y.; Sun, Y.; Zhang, Y.; Miao, H.; You, Q.D.; Hu, R.; Guo, Q.L. NF-κB and Nrf2 signaling pathways contribute to wogonin-mediated inhibition of in-flammation-associated colorectal carcinogenesis. Cell Death Dis. 2014, 5, e1283. [Google Scholar] [CrossRef] [PubMed]
  132. Sassi, N.; Mattarei, A.; Espina, V.; Liotta, L.; Zoratti, M.; Paradisi, C.; Biasutto, L. Potential anti-cancer activity of 7-O-pentyl quercetin: Efficient, membrane-targeted kinase inhibition and pro-oxidant effect. Pharmacol. Res. 2017, 124, 9–19. [Google Scholar] [CrossRef] [PubMed]
  133. Lai, W.W.; Hsu, S.C.; Chueh, F.S.; Chen, Y.Y.; Yang, J.S.; Lin, J.P.; Lien, J.C.; Tsai, C.H.; Chung, J.G. Quercetin inhibits migration and invasion of SAS human oral cancer cells through inhibition of NF-κB and matrix metalloproteinase-2/-9 signaling pathways. Anticancer. Res. 2013, 33, 1941–1950. [Google Scholar] [PubMed]
  134. Pintha, K.; Chaiwangyen, W.; Yodkeeree, S.; Suttajit, M.; Tantipaiboonwong, P. Suppressive Effects of Rosmarinic Acid Rich Fraction from Perilla on Oxidative Stress, Inflammation and Metastasis Ability in A549 Cells Exposed to PM via C-Jun, P-65-Nf-Κb and Akt Signaling Pathways. Biomolecules 2021, 11, 1090. [Google Scholar] [CrossRef]
  135. Sekar, V.; Anandasadagopan, S.K.; Ganapasam, S. Genistein regulates tumor microenvironment and exhibits anticancer effect in dimethyl hydrazine-induced experimental colon carcinogenesis. BioFactors 2016, 42, 623–637. [Google Scholar] [CrossRef]
  136. Xi, X.; Wang, J.; Qin, Y.; You, Y.; Huang, W.; Zhan, J. The Biphasic Effect of Flavonoids on Oxidative Stress and Cell Prolif-eration in Breast Cancer Cells. Antioxidants 2022, 11, 622. [Google Scholar] [CrossRef]
  137. Souza, K.S.; Moreira, L.S.; Silva, B.T.; Oliveira, B.P.; Carvalho, A.S.; Silva, P.S.; Verri, W.A.; Sá-Nakanishi, A.B.; Bracht, L.; Zanoni, J.N.; et al. Low dose of quercetin-loaded pectin/casein microparticles reduces the oxidative stress in arthritic rats. Life Sci. 2021, 284, 119910. [Google Scholar] [CrossRef]
  138. Lee, H.S.; Santana, Á.L.; Peterson, J.; Yucel, U.; Perumal, R.; De Leon, J.; Lee, S.H.; Smolensky, D. Anti-Adipogenic Activity of High-Phenolic Sorghum Brans in Pre-Adipocytes. Nutrients 2022, 14, 1493. [Google Scholar] [CrossRef]
  139. Han, Y.; Dong, Z.; Wang, C.; Li, Q.; Hao, Y.; Yang, Z.; Zhu, W.; Zhang, Y.; Liu, Z.; Feng, L. Ferrous ions doped calcium carbonate nanoparticles potentiate chemotherapy by in-ducing ferroptosis. J. Control. Release 2022, 348, 346–356. [Google Scholar] [CrossRef]
  140. Han, L.; Li, L.; Wu, G. Induction of ferroptosis by carnosic acid-mediated inactivation of Nrf2/HO-1 potentiates cisplatin responsiveness in OSCC cells. Mol. Cell. Probes 2022, 64, 101821. [Google Scholar] [CrossRef]
  141. Nile, A.; Nile, S.H.; Shin, J.; Park, G.; Oh, J.-W. Quercetin-3-Glucoside Extracted from Apple Pomace Induces Cell Cycle Arrest and Apoptosis by Increasing Intracellular ROS Levels. Int. J. Mol. Sci. 2021, 22, 10749. [Google Scholar] [CrossRef] [PubMed]
  142. Zheng, N.; Liu, L.; Liu, W.-W.; Li, F.; Hayashi, T.; Tashiro, S.-I.; Onodera, S.; Ikejima, T. Crosstalk of ROS/RNS and autophagy in silibinin-induced apoptosis of MCF-7 human breast cancer cells in vitro. Acta Pharmacol. Sin. 2017, 38, 277–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Liang, W.-F.; Gong, Y.-X.; Li, H.-F.; Sun, F.-L.; Li, W.-L.; Chen, D.-Q.; Xie, D.-P.; Ren, C.-X.; Guo, X.-Y.; Wang, Z.-Y.; et al. Curcumin Activates ROS Signaling to Promote Pyroptosis in Hepatocellular Carcinoma HepG2 Cells. In Vivo 2021, 35, 249–257. [Google Scholar] [CrossRef]
  144. Ghosh, S.; Dutta, N.; Banerjee, P.; Gajbhiye, R.L.; Sareng, H.R.; Kapse, P.; Pal, S.; Burdelya, L.; Mandal, N.C.; Ravichandiran, V.; et al. Induction of monoamine oxidase A-mediated oxidative stress and impairment of NRF2-antioxidant defence response by polyphenol-rich fraction of Bergenia ligulata sensitizes prostate cancer cells in vitro and in vivo. Free. Radic. Biol. Med. 2021, 172, 136–151. [Google Scholar] [CrossRef] [PubMed]
  145. Rupasinghe, H.P.V.; Dellaire, G.; Xu, Z. Regulation of Nrf2/ARE Pathway by Dietary Flavonoids: A Friend or Foe for Cancer Management? Antioxidants 2020, 9, 973. [Google Scholar]
  146. Wang, C.; Song, X.; Shang, M.; Zou, W.; Zhang, M.; Wei, H.; Shao, H. Curcumin exerts cytotoxicity dependent on reactive oxygen species accumulation in non-small-cell lung cancer cells. Futur. Oncol. 2019, 15, 1243–1253. [Google Scholar] [CrossRef]
  147. Cao, X.; Li, Y.; Wang, Y.; Yu, T.; Zhu, C.; Zhang, X.; Guan, J. Curcumin suppresses tumorigenesis by ferroptosis in breast cancer. PLoS ONE 2022, 17, e0261370. [Google Scholar] [CrossRef]
  148. Tang, X.; Ding, H.; Liang, M.; Chen, X.; Yan, Y.; Wan, N.; Chen, Q.; Zhang, J.; Cao, J. Curcumin induces ferroptosis in non-small-cell lung cancer via activating autophagy. Thorac. Cancer 2021, 12, 1219–1230. [Google Scholar] [CrossRef]
  149. Kalinina, E.V.; Hasan, A.A.S.; Tatarskiy, V.V.; Volodina, Y.L.; Petrova, A.S.; Novichkova, M.D.; Zhdanov, D.D.; Nurmuradov, N.K.; Chernov, N.N.; Shtil, A.A. Suppression of PI3K/Akt/mTOR Signaling Pathway and Antioxidant System and Reversal of Cancer Cells Resistance to Cisplatin under the Effect of Curcumin. Bull. Exp. Biol. Med. 2022, 173, 371–375. [Google Scholar] [CrossRef]
  150. Obaidi, I.; Cassidy, H.; Gaspar, V.I.; McCaul, J.; Higgins, M.; Halász, M.; Reynolds, A.L.; Kennedy, B.N.; McMorrow, T. Curcumin Sensitizes Kidney Cancer Cells to TRAIL-Induced Apoptosis via ROS Mediated Activation of JNK-CHOP Pathway and Upregulation of DR4. Biology 2020, 9, 92. [Google Scholar] [CrossRef]
  151. Feng, C.; Xia, Y.; Zou, P.; Shen, M.; Hu, J.; Ying, S.; Pan, J.; Liu, Z.; Dai, X.; Zhuge, W.; et al. Curcumin analog L48H37 induces apoptosis through ROS-mediated endoplasmic reticulum stress and STAT3 pathways in human lung cancer cells. Mol. Carcinog. 2017, 56, 1765–1777. [Google Scholar] [CrossRef] [PubMed]
  152. Wang, C.; Cui, C. Inhibition of Lung Cancer Proliferation by Wogonin is Associated with the Activation of Apoptosis and Generation of Reactive Oxygen Species. Balk. Med. J. 2019, 37, 29–33. [Google Scholar] [CrossRef] [PubMed]
  153. Mirzaei, S.; Zarrabi, A.; Hashemi, F.; Zabolian, A.; Saleki, H.; Azami, N.; Hamzehlou, S.; Farahani, M.V.; Hushmandi, K.; Ashrafizadeh, M.; et al. Nrf2 Signaling Pathway in Chemoprotection and Doxorubicin Resistance: Po-tential Application in Drug Discovery. Antioxidants 2021, 10, 349. [Google Scholar] [CrossRef] [PubMed]
  154. Zhong, Y.; Zhang, F.; Sun, Z.; Zhou, W.; Li, Z.Y.; You, Q.D.; Guo, Q.L.; Hu, R. Drug resistance associates with activation of Nrf2 in MCF-7/DOX cells, and wogonin reverses it by down-regulating Nrf2-mediated cellular defense response. Mol. Carcinog. 2013, 52, 824–834. [Google Scholar] [CrossRef] [PubMed]
  155. Qian, C.; Wang, Y.; Zhong, Y.; Tang, J.; Zhang, J.; Li, Z.; Wang, Q.; Hu, R. Wogonin-enhanced reactive oxygen species-induced apoptosis and potentiated cyto-toxic effects of chemotherapeutic agents by suppression Nrf2-mediated signaling in HepG2 cells. Free. Radic. Res. 2014, 48, 607–621. [Google Scholar] [CrossRef] [PubMed]
  156. Xu, X.; Zhang, Y.; Li, W.; Miao, H.; Zhang, H.; Zhou, Y.; Li, Z.; You, Q.; Zhao, L.; Guo, Q. Wogonin reverses multi-drug resistance of human myelogenous leukemia K562/A02 cells via downregulation of MRP1 expression by inhibiting Nrf2/ARE signaling pathway. Biochem. Pharmacol. 2014, 92, 220–234. [Google Scholar] [CrossRef] [PubMed]
  157. Kim, E.H.; Jang, H.; Shin, D.; Baek, S.H.; Roh, J.-L. Targeting Nrf2 with wogonin overcomes cisplatin resistance in head and neck cancer. Apoptosis 2016, 21, 1265–1278. [Google Scholar] [CrossRef]
  158. Chen, X.; Li, W.; Xu, C.; Wang, J.; Zhu, B.; Huang, Q.; Chen, D.; Sheng, J.; Zou, Y.; Lee, Y.M.; et al. Comparative profiling of analog targets: A case study on resveratrol for mouse melanoma metastasis suppression. Theranostics 2018, 8, 3504–3516. [Google Scholar] [CrossRef]
  159. Sinha, D.; Sarkar, N.; Biswas, J.; Bishayee, A. Resveratrol for breast cancer prevention and therapy: Preclinical evidence and molecular mechanisms. Semin. Cancer Biol. 2016, 40–41, 209–232. [Google Scholar] [CrossRef]
  160. Patra, S.; Pradhan, B.; Nayak, R.; Behera, C.; Rout, L.; Jena, M.; Efferth, T.; Bhutia, S.K. Chemotherapeutic efficacy of curcumin and resveratrol against cancer: Chemo-prevention, chemoprotection, drug synergism and clinical pharmacokinetics. Semin Cancer Biol. 2021, 73, 310–320. [Google Scholar] [CrossRef]
  161. Wang, X.X.; Li, Y.B.; Yao, H.J.; Ju, R.J.; Zhang, Y.; Li, R.J.; Yu, Y.; Zhang, L.; Lu, W.L. The use of mitochondrial targeting resveratrol liposomes modified with a dequalinium polyethylene glycol-distearoylphosphatidyl ethanolamine conjugate to induce apoptosis in resistant lung cancer cells. Biomaterials 2011, 32, 5673–5687. [Google Scholar] [CrossRef] [PubMed]
  162. Zhang, Y.; Yuan, F.; Li, P.; Gu, J.; Han, J.; Ni, Z.; Liu, F. Resveratrol inhibits HeLa cell proliferation by regulating mitochondrial function. Ecotoxicol. Environ. Saf. 2022, 241, 113788. [Google Scholar] [CrossRef] [PubMed]
  163. da Costa, P.S.; Ramos, P.S.; Ferreira, C.; Silva, J.L.; El-Bacha, T.; Fialho, E. Pro-Oxidant Effect of Resveratrol on Human Breast Cancer MCF-7 Cells is Associated with CK2 Inhibition. Nutr. Cancer 2022, 74, 2142–2151. [Google Scholar] [CrossRef] [PubMed]
  164. Jia, B.; Zheng, X.; Wu, M.L.; Tian, X.T.; Song, X.; Liu, Y.N.; Li, P.N.; Liu, J. Increased Reactive Oxygen Species and Distinct Oxidative Damage in Resvera-trol-suppressed Glioblastoma Cells. J. Cancer 2021, 12, 141–149. [Google Scholar] [CrossRef]
  165. Fouzder, C.; Mukhuty, A.; Kundu, R. Kaempferol inhibits Nrf2 signalling pathway via downregulation of Nrf2 mRNA and induces apoptosis in NSCLC cells. Arch. Biochem. Biophys. 2021, 697, 108700. [Google Scholar] [CrossRef]
  166. Deguchi, Y.; Ito, M. Rosmarinic acid in Perilla frutescens and perilla herb analyzed by HPLC. J. Nat. Med. 2020, 74, 341–352. [Google Scholar] [CrossRef]
  167. Chou, S.-T.; Ho, B.-Y.; Tai, Y.-T.; Huang, C.-J.; Chao, W.-W. Bidirect effects from cisplatin combine with rosmarinic acid (RA) or hot water extracts of Glechoma hederacea (HWG) on renal cancer cells. Chin. Med. 2020, 15, 77. [Google Scholar] [CrossRef]
  168. Ma, Z.; Yang, J.; Yang, Y.; Wang, X.; Chen, G.; Shi, A.; Lu, Y.; Jia, S.; Kang, X.; Lu, L. Rosmarinic acid exerts an anticancer effect on osteosarcoma cells by inhibiting DJ-1 via regulation of the PTEN-PI3K-Akt signaling pathway. Phytomedicine 2020, 68, 153186. [Google Scholar] [CrossRef]
  169. Luo, Y.; Ma, Z.; Xu, X.; Qi, H.; Cheng, Z.; Chen, L. Anticancer effects of rosmarinic acid in human oral cancer cells is mediated via endoplasmic reticulum stress, apoptosis, G2/M cell cycle arrest and inhibition of cell migration. J. Balk. Union Oncol. 2020, 25, 1245–1250. [Google Scholar]
  170. Majumder, D.; Das, A.; Saha, C. Catalase inhibition an anti cancer property of flavonoids: A kinetic and structural evaluation. Int. J. Biol. Macromol. 2017, 104, 929–935. [Google Scholar] [CrossRef]
  171. Ma, Y.; Yao, C.; Liu, H.; Yu, F.; Lin, J.; Lu, K.; Liao, C.; Chueh, F.; Chung, J. Quercetin induced apoptosis of human oral cancer SAS cells through mitochondria and endoplasmic reticulum mediated signaling pathways. Oncol. Lett. 2018, 15, 9663–9672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Raja, S.B.; Rajendiran, V.; Kasinathan, N.K.; P, A.; Venkatabalasubramanian, S.; Murali, M.R.; Devaraj, H.; Devaraj, S.N. Differential cytotoxic activity of Quercetin on colonic cancer cells depends on ROS generation through COX-2 expression. Food Chem. Toxicol. 2017, 106, 92–106. [Google Scholar] [CrossRef]
  173. Wu, Q.; Needs, P.W.; Lu, Y.; Kroon, P.A.; Ren, D.; Yang, X. Different antitumor effects of quercetin, quercetin-3’-sulfate and quercetin-3-glucuronide in human breast cancer MCF-7 cells. Food Funct. 2018, 9, 1736–1746. [Google Scholar] [CrossRef]
  174. Wang, Z.; Ma, J.; Li, X.; Wu, Y.; Shi, H.; Chen, Y.; Lu, G.; Shen, H.; Lu, G.; Zhou, J. Quercetin induces p53-independent cancer cell death through lysosome activation by the transcription factor EB and Reactive Oxygen Species-dependent ferroptosis. Br. J. Pharmacol. 2021, 178, 1133–1148. [Google Scholar] [CrossRef] [PubMed]
  175. Oliveira, A.L.; Valente, D.; Moreira, H.R.; Pintado, M.; Costa, P. Effect of squalane-based emulsion on polyphenols skin penetration: Ex vivo skin study. Colloids Surfaces B Biointerfaces 2022, 218, 112779. [Google Scholar] [CrossRef] [PubMed]
  176. Pietta, P.G. Flavonoids as antioxidants. J. Nat. Prod. 2000, 63, 1035–1042. [Google Scholar] [CrossRef] [PubMed]
  177. Ouyang, J.; Zhu, K.; Liu, Z.; Huang, J. Prooxidant Effects of Epigallocatechin-3-Gallate in Health Benefits and Potential Adverse Effect. Oxidative Med. Cell. Longev. 2020, 2020, 9723686. [Google Scholar] [CrossRef]
  178. Costa, C.; Tsatsakis, A.; Mamoulakis, C.; Teodoro, M.; Briguglio, G.; Caruso, E.; Tsoukalas, D.; Margina, D.; Dardiotis, E.; Kouretas, D.; et al. Current evidence on the effect of dietary polyphenols intake on chronic diseases. Food Chem. Toxicol. 2017, 110, 286–299. [Google Scholar] [CrossRef]
  179. Fraga, C.G.; Galleano, M.; Verstraeten, S.V.; Oteiza, P.I. Basic biochemical mechanisms behind the health benefits of poly-phenols. Mol. Asp. Med. 2010, 31, 435–445. [Google Scholar] [CrossRef]
  180. Saunders, N.A.; Simpson, F.; Thompson, E.W.; Hill, M.M.; Endo-Munoz, L.; Leggatt, G.; Minchin, R.; Guminski, A. Role of intratumoural heterogeneity in cancer drug resistance: Molecular and clinical perspectives. EMBO Mol. Med. 2012, 4, 675–684. [Google Scholar] [CrossRef] [Green Version]
  181. Zhong, Y.; Shahidi, F. Lipophilized Epigallocatechin Gallate (EGCG) Derivatives as Novel Antioxidants. J. Agric. Food Chem. 2011, 59, 6526–6533. [Google Scholar] [CrossRef] [PubMed]
  182. Chen, J.; Zhang, L.; Li, C.; Chen, R.; Liu, C.; Chen, M. Lipophilized Epigallocatechin Gallate Derivative Exerts Anti-Proliferation Efficacy through Induction of Cell Cycle Arrest and Apoptosis on DU145 Human Prostate Cancer Cells. Nutrients 2019, 12, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Forester, S.C.; Lambert, J.D. The catechol-O-methyltransferase inhibitor, tolcapone, increases the bioavailability of un-methylated (−)-epigallocatechin-3-gallate in mice. J. Funct Foods 2015, 17, 183–188. [Google Scholar] [CrossRef]
  184. Zhang, M.; Zhang, X.; Ho, C.T.; Huang, Q. Chemistry and Health Effect of Tea Polyphenol (-)-Epigallocatechin 3- O-(3- O-Methyl)gallate. J. Agric. Food Chem. 2019, 67, 5374–5378. [Google Scholar] [CrossRef] [PubMed]
  185. Kim, J.; Nguyen, T.T.H.; Kim, N.M.; Moon, Y.-H.; Ha, J.-M.; Park, N.; Lee, D.-G.; Hwang, K.-H.; Park, J.-S.; Kim, D. Functional Properties of Novel Epigallocatechin Gallate Glucosides Synthesized by Using Dextransucrase from Leuconostoc mesenteroides B-1299CB4. J. Agric. Food Chem. 2016, 64, 9203–9213. [Google Scholar] [CrossRef] [PubMed]
  186. Méndez-Líter, J.A.; Pozo-Rodríguez, A.; Madruga, E.; Rubert, M.; Santana, A.G.; de Eugenio, L.I.; Sánchez, C.; Martínez, A.; Prieto, A.; Martínez, M.J. Glycosylation of Epigallocatechin Gallate by Engineered Glyco-side Hydrolases from Talaromyces amestolkiae: Potential Antiproliferative and Neuroprotective Effect of These Molecules. Antioxidants 2022, 11, 1325. [Google Scholar] [CrossRef] [PubMed]
  187. Gonzalez-Alfonso, J.L.; Leemans, L.; Poveda, A.; Jimenez-Barbero, J.; Ballesteros, A.O.; Plou, F.J. Efficient α-Glucosylation of Epigallocatechin Gallate Catalyzed by Cyclodextrin Glucanotransferase from Thermoanaerobacter Species. J. Agric. Food Chem. 2018, 66, 7402–7408. [Google Scholar] [CrossRef]
  188. Boateng, I.D. Recent processing of fruits and vegetables using emerging thermal and non-thermal technologies. A critical review of their potentialities and limitations on bioactives, structure, and drying performance. Crit. Rev. 2022, 1–35. [Google Scholar] [CrossRef]
  189. Rambaran, T.F. A patent review of polyphenol nano-formulations and their commercialization. Trends Food Sci Tech. 2022, 120, 111–122. [Google Scholar] [CrossRef]
  190. Cao, H.; Saroglu, O.; Karadag, A.; Diaconeasa, Z.; Zoccatelli, G.; Conte-Junior, C.A.; Gonzalez-Aguilar, G.A.; Ou, J.; Bai, W.; Zamarioli, C.M.; et al. Available technologies on improving the stability of polyphenols in food processing. Food Front. 2021, 2, 109–139. [Google Scholar] [CrossRef]
  191. De Oliveira, T.V.; Stein, R.; de Andrade, D.F.; Beck, R.C.R. Preclinical studies of the antitumor effect of curcumin-loaded polymeric nanocapsules: A systematic review and meta-analysis. Phytother. Res. 2022, 36, 3202–3214. [Google Scholar] [CrossRef]
  192. Gugleva, V.; Ivanova, N.; Sotirova, Y.; Andonova, V. Dermal Drug Delivery of Phytochemicals with Phenolic Structure via Lipid-Based Nanotechnologies. Pharmaceuticals 2021, 14, 837. [Google Scholar] [CrossRef]
  193. Bohara, R.A.; Tabassum, N.; Singh, M.P.; Gigli, G.; Ragusa, A.; Leporatti, S. Recent Overview of Resveratrol’s Beneficial Effects and Its Nano-Delivery Systems. Molecules 2022, 27, 5154. [Google Scholar] [CrossRef] [PubMed]
  194. Lu, S.; Tian, H.; Li, L.; Li, B.; Yang, M.; Zhou, L.; Jiang, H.; Li, Q.; Wang, W.; Nice, E.C.; et al. Nanoengineering a Zeolitic Imidazolate Framework-8 Capable of Manipulating Energy Me-tabolism against Cancer Chemo-Phototherapy Resistance. Small 2022, 18, e2204926. [Google Scholar] [CrossRef] [PubMed]
  195. Yu, H.; Palazzolo, J.S.; Ju, Y.; Niego, B.; Pan, S.; Hagemeyer, C.E.; Caruso, F. Polyphenol-Functionalized Cubosomes as Thrombolytic Drug Carriers. Adv. Health Mater. 2022, 11, e2201151. [Google Scholar] [CrossRef] [PubMed]
  196. Hafez Ghoran, S.; Calcaterra, A.; Abbasi, M.; Taktaz, F.; Nieselt, K.; Babaei, E. Curcumin-Based Nanoformulations: A Promising Adjuvant towards Cancer Treatment. Molecules 2022, 27, 5236. [Google Scholar] [CrossRef] [PubMed]
  197. Soleti, R.; Andriantsitohaina, R.; Martinez, M.C. Impact of polyphenols on extracellular vesicle levels and effects and their properties as tools for drug delivery for nutrition and health. Arch. Biochem. Biophys. 2018, 644, 57–63. [Google Scholar] [CrossRef]
  198. Ma, Y.; Liu, J.; Cui, X.; Hou, J.; Yu, F.; Wang, J.; Wang, X.; Chen, C.; Tong, L. Hyaluronic Acid Modified Nanostructured Lipid Carrier for Targeting Delivery of Kaempferol to NSCLC: Preparation, Optimization, Characterization, and Performance Evaluation In Vitro. Molecules 2022, 27, 4553. [Google Scholar] [CrossRef]
  199. Shrivastava, N.; Parikh, A.; Dewangan, R.P.; Biswas, L.; Verma, A.K.; Mittal, S.; Ali, J.; Garg, S.; Baboota, S. Solid Self-Nano Emulsifying Nanoplatform Loaded with Tamoxifen and Resveratrol for Treatment of Breast Cancer. Pharmaceutics 2022, 14, 1486. [Google Scholar] [CrossRef]
  200. Azadikhah, F.; Karimi, A.R.; Yousefi, G.H.; Hadizadeh, M. Dual antioxidant-photosensitizing hydrogel system: Cross-linking of chitosan with tannic acid for enhanced photodynamic efficacy. Int. J. Biol. Macromol. 2021, 188, 114–125. [Google Scholar] [CrossRef]
  201. Zu, M.; Xie, D.; Canup, B.S.; Chen, N.; Wang, Y.; Sun, R.; Zhang, Z.; Fu, Y.; Dai, F.; Xiao, B. ‘Green’ nanotherapeutics from tea leaves for orally targeted prevention and alleviation of colon diseases. Biomaterials 2021, 279, 121178. [Google Scholar] [CrossRef] [PubMed]
  202. Chen, Q.; Li, Q.; Liang, Y.; Zu, M.; Chen, N.; Canup, B.S.; Luo, L.; Wang, C.; Zeng, L.; Xiao, B. Natural exosome-like nanovesicles from edible tea flowers suppress metastatic breast cancer via ROS generation and microbiota modulation. Acta Pharm. Sin. B 2022, 12, 907–923. [Google Scholar] [CrossRef] [PubMed]
  203. Mereles, D.; Hunstein, W. Epigallocatechin-3-gallate (EGCG) for clinical trials: More pitfalls than promises? Int. J. Mol. Sci. 2011, 12, 5592–5603. [Google Scholar] [CrossRef] [PubMed]
  204. Yu, C.; Jiao, Y.; Xue, J.; Zhang, Q.; Yang, H.; Xing, L.; Chen, G.; Wu, J.; Zhang, S.; Zhu, W.; et al. Metformin Sensitizes Non-small Cell Lung Cancer Cells to an Epigallocatechin-3-Gallate (EGCG) Treatment by Suppressing the Nrf2/HO-1 Signaling Pathway. Int. J. Biol. Sci. 2017, 13, 1560–1569. [Google Scholar] [CrossRef]
  205. Du, D.; Tang, X.; Li, Y.; Gao, Y.; Chen, R.; Chen, Q.; Wen, J.; Wu, T.; Zhang, Y.; Lu, H.; et al. Senotherapy Protects against Cisplatin-Induced Ovarian Injury by Removing Senescent Cells and Alleviating DNA Damage. Oxidative Med. Cell. Longev. 2022, 2022, 9144644. [Google Scholar] [CrossRef]
  206. Schwingel, T.E.; Klein, C.P.; Nicoletti, N.F.; Dora, C.L.; Hädrich, G.; Bica, C.G.; Lopes, T.G.; Da Silva, V.D.; Morrone, F.B. Effects of the compounds resveratrol, rutin, quercetin, and quercetin nanoemulsion on oxaliplatin-induced hepatotoxicity and neurotoxicity in mice. Naunyn-Schmiedebergs Arch. Exp. Pathol. Pharmakol. 2014, 387, 837–848. [Google Scholar] [CrossRef]
  207. Zhang, X.; Huang, J.; Yu, C.; Xiang, L.; Li, L.; Shi, D.; Lin, F. Quercetin Enhanced Paclitaxel Therapeutic Effects Towards PC-3 Prostate Cancer Through ER Stress Induction and ROS Production. OncoTargets Ther. 2020, 13, 513–523. [Google Scholar] [CrossRef] [Green Version]
  208. Jalili-Nik, M.; Soltani, A.; Moussavi, S.; Ghayour-Mobarhan, M.; Ferns, G.A.; Hassanian, S.M.; Avan, A. Current status and future prospective of Curcumin as a potential therapeutic agent in the treatment of colorectal cancer. J. Cell. Physiol. 2017, 233, 6337–6345. [Google Scholar] [CrossRef]
  209. Zangui, M.; Atkin, S.L.; Majeed, M.; Sahebkar, A. Current evidence and future perspectives for curcumin and its analogues as promising adjuncts to oxaliplatin: State-of-the-art. Pharmacol. Res. 2019, 141, 343–356. [Google Scholar] [CrossRef]
  210. Zhang, X.; Guan, Z.; Wang, X.; Sun, D.; Wang, D.; Li, Y.; Pei, B.; Ye, M.; Xu, J.; Yue, X. Curcumin Alleviates Oxaliplatin-Induced Peripheral Neuropathic Pain through In-hibiting Oxidative Stress-Mediated Activation of NF-κB and Mitigating Inflammation. Biol. Pharm. Bull. 2020, 43, 348–355. [Google Scholar] [CrossRef] [Green Version]
  211. Santandreu, F.M.; Valle, A.; Oliver, J.; Roca, P. Resveratrol potentiates the cytotoxic oxidative stress induced by chemo-therapy in human colon cancer cells. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2011, 28, 219–228. [Google Scholar] [CrossRef] [PubMed]
  212. Le Gal, K.; Ibrahim, M.X.; Wiel, C.; Sayin, V.I.; Akula, M.K.; Karlsson, C.; Dalin, M.G.; Akyürek, L.M.; Lindahl, P.; Nilsson, J.; et al. Antioxidants can increase melanoma metastasis in mice. Sci. Transl. Med. 2015, 7, 308re308. [Google Scholar] [CrossRef] [PubMed]
  213. Sayin, V.I.; Ibrahim, M.X.; Larsson, E.; Nilsson, J.A.; Lindahl, P.; Bergo, M.O. Antioxidants accelerate lung cancer progression in mice. Sci. Transl. Med. 2014, 6, 221ra215. [Google Scholar] [CrossRef] [PubMed]
  214. Piskounova, E.; Agathocleous, M.; Murphy, M.M.; Hu, Z.; Huddlestun, S.E.; Zhao, Z.; Leitch, A.M.; Johnson, T.M.; DeBerardinis, R.J.; Morrison, S.J. Oxidative stress inhibits distant metastasis by human melanoma cells. Nature 2015, 527, 186–191. [Google Scholar] [CrossRef] [Green Version]
  215. Brunner, M.; Ziegler, S.; Di Stefano, A.F.D.; Dehghanyar, P.; Kletter, K.; Tschurlovits, M.; Villa, R.; Bozzella, R.; Celasco, G.; Moro, L.; et al. Gastrointestinal transit, release and plasma pharmacokinetics of a new oral budesonide formulation. Br. J. Clin. Pharmacol. 2006, 61, 31–38. [Google Scholar] [CrossRef]
Figure 1. The basic structure of flavonoids and their six main subgroups. This figure was created using Kingdraw.
Figure 1. The basic structure of flavonoids and their six main subgroups. This figure was created using Kingdraw.
Pharmaceuticals 15 01540 g001
Figure 2. Redox homeostasis involved in cancer evolution. Reactive oxygen species (ROS) almost participate in all aspects of cancer, including initiation, proliferation, metastasis, and drug resistance. Cancer cells evolved a set of robust antioxidant system, including reducing power (e.g., NADPH), non-enzymatic and enzymatic antioxidants, and transcription factors, for adapting to oxidative stress.
Figure 2. Redox homeostasis involved in cancer evolution. Reactive oxygen species (ROS) almost participate in all aspects of cancer, including initiation, proliferation, metastasis, and drug resistance. Cancer cells evolved a set of robust antioxidant system, including reducing power (e.g., NADPH), non-enzymatic and enzymatic antioxidants, and transcription factors, for adapting to oxidative stress.
Pharmaceuticals 15 01540 g002
Figure 3. Outlines of antioxidant properties of polyphenols in cancer therapy. Polyphenols can reinforce the transcriptional activation of the classical antioxidant transcription factor NRF2 to defend against ROS-induced DNA damage, thus preventing oxidative stress-mediated carcinogenesis. Meanwhile, polyphenols could also upregulate/activate antioxidant enzymes or activate antioxidant signaling pathways to restore intracellular homeostasis for enhancing P21-mediated cell cycle arrest and blocking the epithelial mesenchymal transition (EMT) process and matrix metalloproteinase (MMP)-mediated metastasis. In addition, polyphenols are able to manipulate the activity of antioxidant or prooxidant transcription factors, thus overcoming ROS-mediated drug resistance.
Figure 3. Outlines of antioxidant properties of polyphenols in cancer therapy. Polyphenols can reinforce the transcriptional activation of the classical antioxidant transcription factor NRF2 to defend against ROS-induced DNA damage, thus preventing oxidative stress-mediated carcinogenesis. Meanwhile, polyphenols could also upregulate/activate antioxidant enzymes or activate antioxidant signaling pathways to restore intracellular homeostasis for enhancing P21-mediated cell cycle arrest and blocking the epithelial mesenchymal transition (EMT) process and matrix metalloproteinase (MMP)-mediated metastasis. In addition, polyphenols are able to manipulate the activity of antioxidant or prooxidant transcription factors, thus overcoming ROS-mediated drug resistance.
Pharmaceuticals 15 01540 g003
Figure 4. Prooxidant roles of polyphenols in cancer therapy. Typically, polyphenols induce glutathione (GSH) depletion, iron overload, lipid peroxidation, autophagosome augmentation, and caspase3 cleavage to trigger programmed death of cancer cells, including ferroptosis, autophagy, apoptosis, and pyroptosis. Moreover, polyphenols evoke endoplasmic reticulum (ER) stress and decrease in mitochondrial membrane potential, which could also induce cell death. On the other hand, polyphenol-induced ROS are capable of cleaving caspase3/8 9 and impairing the activities of MMP2/9, thus suppressing metastasis of cancer cell. In addition, polyphenols could inhibit the transcriptional activity of NRF2 to improve the sensitivity of cancer cells to chemotherapy.
Figure 4. Prooxidant roles of polyphenols in cancer therapy. Typically, polyphenols induce glutathione (GSH) depletion, iron overload, lipid peroxidation, autophagosome augmentation, and caspase3 cleavage to trigger programmed death of cancer cells, including ferroptosis, autophagy, apoptosis, and pyroptosis. Moreover, polyphenols evoke endoplasmic reticulum (ER) stress and decrease in mitochondrial membrane potential, which could also induce cell death. On the other hand, polyphenol-induced ROS are capable of cleaving caspase3/8 9 and impairing the activities of MMP2/9, thus suppressing metastasis of cancer cell. In addition, polyphenols could inhibit the transcriptional activity of NRF2 to improve the sensitivity of cancer cells to chemotherapy.
Pharmaceuticals 15 01540 g004
Figure 5. Novel strategies promote the application of polyphenols in cancer therapy. Structural modifications, such as esterification, methylation, and glycosylation, protect polyphenols from rapid degradation and promote their bioactivities. Nanotechnologies can make up the inherent limitations of polyphenols and support targeting delivery of them to specific tumor lesions with minor side effects. Combination therapies achieve efficient tumor eliminate via enhanced ROS-mediated mechanisms.
Figure 5. Novel strategies promote the application of polyphenols in cancer therapy. Structural modifications, such as esterification, methylation, and glycosylation, protect polyphenols from rapid degradation and promote their bioactivities. Nanotechnologies can make up the inherent limitations of polyphenols and support targeting delivery of them to specific tumor lesions with minor side effects. Combination therapies achieve efficient tumor eliminate via enhanced ROS-mediated mechanisms.
Pharmaceuticals 15 01540 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, L.; Jin, P.; Guan, Y.; Luo, M.; Wang, Y.; He, B.; Li, B.; He, K.; Cao, J.; Huang, C.; et al. Exploiting Polyphenol-Mediated Redox Reorientation in Cancer Therapy. Pharmaceuticals 2022, 15, 1540. https://doi.org/10.3390/ph15121540

AMA Style

Li L, Jin P, Guan Y, Luo M, Wang Y, He B, Li B, He K, Cao J, Huang C, et al. Exploiting Polyphenol-Mediated Redox Reorientation in Cancer Therapy. Pharmaceuticals. 2022; 15(12):1540. https://doi.org/10.3390/ph15121540

Chicago/Turabian Style

Li, Lei, Ping Jin, Yueyue Guan, Maochao Luo, Yu Wang, Bo He, Bowen Li, Kai He, Jiangjun Cao, Canhua Huang, and et al. 2022. "Exploiting Polyphenol-Mediated Redox Reorientation in Cancer Therapy" Pharmaceuticals 15, no. 12: 1540. https://doi.org/10.3390/ph15121540

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