3,393
Views
2
CrossRef citations to date
0
Altmetric
Review

Cancer Protective Role of Selected Dietary Polyphenols via Modulating Keap1/Nrf2/ARE and Interconnected Signaling Pathways

, , , , &
Pages 1065-1102 | Received 23 Aug 2022, Accepted 16 Feb 2023, Published online: 20 Apr 2023

Abstract

The Kelch-like ECH associated protein 1 (Keap1)/nuclear factor erythroid 2-related factor 2 (Nrf2)/antioxidant response elements (ARE) signaling pathway is considered a master regulator of the cellular response against oxidative stress. Numerous studies have investigated the role of Keap1/Nrf2/ARE in the different stages of cancer development. A comprehensive literature search using the Google Scholar, PubMed and Science Direct databases was performed to retrieve information related to the cancer protective role of 21 selected dietary polyphenols via modulation of Keap1/Nrf2/ARE and interconnected signaling pathways/proteins (MAPK/ERK1/2, PI3K/Akt, PKD, JNKs, AMPK, NF-κB). Information on the anti-inflammatory and cytoprotective effects caused by the selected dietary polyphenols following Keap1/Nrf2/ARE modulation was also collected. The majority of the studies analyzed in this review demonstrated the cancer protective role of the selected polyphenols mostly in-vitro. Limited work was performed in-vivo and only one of the selected polyphenols was subjected to a clinical trial. It is hoped that this review will encourage further in-vivo studies to confirm the cancer protective role of methyleugenol, carnosol, and catechin, as well as further clinical trials to unambiguously establish whether the consumption of dietary polyphenols impacts on the incidence and progression of cancers in humans.

Introduction

Role of Keap1/Nrf2/ARE in Carcinogenesis

The overproduction of reactive oxygen species (ROS)—associated with mitochondrial cellular respiration, phagocytosis, digestion, aging tissues, and the metabolism of xenobiotics—generates high levels of free radicals in cells and causes oxidative stress. The latter has been strongly linked to carcinogenesis as it induces DNA mutations, promotes cancer cell growth and proliferation, stimulates angiogenesis and increases resistance to cell apoptosis and autophagy. Oxidative stress interferes with the mitogen activated-protein kinase (MAPK)/extracellular-regulated kinase 1/2 (ERK1/2), phosphatidylinositol 3-kinase (PI3K)/protein kinase B (Akt), protein kinase D (PKD), c-Jun N-terminal kinases (JNKs), and nuclear factor kappa B (NF-κB), and other cell signaling proteins such as Ras, Raf, Bad, Bax, Bim and Foxo (Citation1–4). At low doses, ROS promote tumor formation by influencing the expression of genes coding for proteins such as protein kinase C (PKC), Nrf2, MAPK, activator protein-1 (AP-1), NF-κB and the hypoxia-inducible factor 1-alpha (HIF-1α) involved in the growth of mutated cells. At high doses, they cause DNA point mutations, deletions, insertions, and chromosomal translocations, which promotes tumor cell mutagenesis by oncogene activation or tumor suppressor gene inactivation (e.g., p. 53) (Citation5,Citation6). Oxidative stress also induces mutation by causing lipid peroxidation and producing malondialdehyde (MDA) (Citation7). In addition, ROS contribute to angiogenesis by regulating the vascular endothelial growth factor (VEGF), p21-activated kinase-1 (PAK-1) in Rac-associated cytoskeleton remodeling, and the ROS-generating enzymes (e.g., NADPH oxidases (NOX)) (Citation8).

The Kelch-like ECH associated protein 1 (Keap1)/nuclear factor erythroid 2-related factor 2 (Nrf2)/antioxidant response elements (ARE) signaling pathway is considered a master regulator of the cellular response against oxidative stress (Citation9,Citation10). Nrf2, encoded upon activation of the nuclear factor erythroid-derived 2 (NFE2) gene, is a transcription factor with an N-terminal conserved region that has DNA-binding specificity. Nrf2 is highly expressed in the skin, lungs, gastrointestinal tract, and in metabolic and detoxifying organs such as the liver and kidneys (Citation11,Citation12). Nrf2 is a major mediator of a variety of functions contributing to cell survival, such as drug/xenobiotic metabolism, DNA repair, mitochondrial function, iron, lipid and carbohydrate metabolism, proteostasis and cell proliferation (Citation13). Nrf2 contains 7 Nrf2-ECH homology (Neh) domains (Neh1-Neh7). The Neh1 domain recognizes ARE for the activation of gene transcription by heterodimerization with small musculoaponeurotic fibrosarcoma proteins (sMAF) K, G, and F. The Neh2 domain mediates Nrf2 ubiquitination and degradation by interacting with the Kelch domain of Keap1 (Citation14). The latter is a cytoplasmic actin-bound adaptor protein that acts as an oxidative stress sensor and negative regulator of Nrf2 (Citation15). Keap1 consists of the following domains: broad complex, tramtrack, bric-a-brac (BTB), Kelch. It also has an N-terminal region, an intervening region (IVR) and a C-terminal region (Citation16). Under basal conditions, Nrf2 exists in the cytoplasm as an inactive complex bound to its negative regulator Keap. There, it undergoes rapid proteasomal degradation triggered by Cul3-directed polyubiquitination through the Keap1/Cul3 ubiquitin ligase (Citation17). In the absence of oxidative stress (i.e., when the need for an antioxidative response is minimal), Nrf2 basal levels are low (Citation18). Under oxidative stress, the highly reactive cysteine thiol groups in the IVR region of Keap1 are oxidized. This results in conformational variations in Keap1, causing its dissociation from Nrf2 (Citation19–21). Once dissociated, the stabilized Nrf2 in the cytosol is translocated into the nucleus upon phosphorylation, along with de novo-synthesized Nrf2 proteins, forming heterodimers with nuclear small muscle aponeurosis fibromatosis (Maf) proteins (Citation22). These Nrf2/sMaf heterodimers then interact with a regulatory enhancer sequence named ARE which regulates about 250 human genes controlling the expression of antioxidant cytoprotective proteins (Citation23). These restore cellular homeostasis following an oxidative insult by regulating inflammation, apoptosis, redox metabolism, iron and heme metabolism, phase-I, -II, and -III drug/xenobiotic metabolism and proteostasis (Citation24,Citation25). The phosphorylation reaction associated with Nrf2 activation is controlled by several kinases, including JNK, PI3K, PKC and ERK (Citation26). Overall, the activation of the Keap1/Nrf2/ARE signaling pathway enhances cellular antioxidant capacity by upregulating the expression of genes coding for detoxifying and cytoprotective enzymes such as glutathione S-transferases (GSTs), NAD(P)H quinine oxidoreductase (NQO-1), glutamylcysteine ligase (GCL), glutathione peroxidases (GPx), superoxide dismutases (SODs), UDP-glucuronosyl transferases (UGTs), heme oxygenase-1 (HO-1), thioredoxin (Trx), and epoxide hydrolases (EH) (Citation27–29). Nrf2 activation also promotes the synthesis/recycling of oxidized co-factors (e.g., glutathione) and detoxifying enzymes (Citation30). As aforementioned, Keap1/Nrf2/ARE widely interacts with other signaling pathways and proteins to regulate the cell redox status. This includes MAPK/ERK1/2, PI3K/Akt, PKD, JNKs, AMPK, NF-κB and proteins such as Ras, Raf, Bad, Bax, Bim and Foxo (Citation1–4). For example, it is known that inflammation induced by cytokines released via the NF-κB pathway is implicated in carcinogenesis through an increase in the production of ROS (Citation31,Citation32). Nrf2 activation inhibits oxidative stress by downregulating NAD(P)H oxidases and genes coding for interleukin-1β (IL-1β) and IL-6, and upregulating gene encoding for the peroxisome proliferator-activated receptor gamma (PPARγ) (Citation33). Previously, we have reported the cancer chemopreventive role of dietary terpenoids via modulating the Keap1-Nrf2-ARE signaling pathway (Citation34). Here, we describe and discuss the cancer protective role of selected dietary polyphenols via the same signaling system.

Role of Keap1/Nrf2/ARE in the Hallmarks of Cancer

Numerous in-vitro and in-vivo studies have investigated the role of Keap1/Nrf2/ARE in cancer, providing evidence for its dual effects (cancer-preventive and cancer-promoting) depending on the different stages of cancer development. Whilst the controlled/transient activation of Keap1/Nrf2/ARE in normal cells can prevent cancer initiation, its uncontrolled/prolonged activation in cancer cells drives cancer promotion, progression, and metastasis. The role of Keap1/Nrf2/ARE in sustained cell proliferation, apoptosis, angiogenesis, and metastasis is described below.

Sustained Cell Proliferation

Studies have demonstrated that Nrf2 targets the expression of genes controlling cell proliferation and survival. Experiments carried out using Nrf2+/+ (wild-type) cells have revealed a significant increase in cell proliferation compared to Nrf2-/- cells. This has also been established In Vivo using Nrf2-/- mice (Citation35,Citation36).

Apoptosis

Apoptosis is generated in response to excessive ROS production, through oxidation of apoptosis signal-regulating kinase 1 (ASK) and activation of p38MAPK and JNK. In cancer cells, activation of the Keap1/Nrf2/ARE pathway leads to resistance to apoptosis via the enhanced expression of B-cell lymphoma-2 (BCL-2) and BCL-xL proteins. These suppress mitochondrial cytochrome C release and decrease caspase-3/7 activation (Citation37). Nrf2 has also recently been found to target the homeodomain-interacting protein kinase 2 (HIPK2) gene which has anti-apoptotic functions (Citation38,Citation39).

Angiogenesis

It has been demonstrated that activation of Keap1/Nrf2/ARE promotes angiogenesis following increased Nrf2-induced NQO1 expression, which allows NQO1 to bind to HIF-1α, inhibiting HIF-1α degradation. Studies have revealed that blood vessel formation is markedly suppressed in Nrf2 knockdown tumor xenograft animal models. This anti-angiogenic effect is mediated via decreased levels of HIF-1α and of VEGF, platelet derived growth factor (PDGF), angiopoietin, and angiogenin (Citation40).

Metastasis

In cancer cells, Keap1/Nrf2/ARE activation promotes the process of epithelial mesenchymal transition (EMT) important for metastasis, via downregulating the expression of the adhesion protein E-cadherin (Citation41). It has also been reported that Nrf2 downregulation correlates with reduced expression of extracellular matrix remodeling enzymes such as the matrix metalloproteinase 2 (MMP2) and matrix metalloproteinase 9 (MMP9) required for cancer cell migration (Citation42). Other conflicting reports have indicated that high NRF2 expression led to anti-metastatic effects (Citation16,Citation43,Citation44). For a comprehensive review on the role of Keap1/Nrf2/ARE in other hallmarks of cancer (e.g., avoiding of immune destruction, tumor-promoting inflammation, genomic instability and others), the reader should refer to the work by de La Vega et al. (Citation45). Two comprehensive updates provide additional details on the role of Keap1/Nrf2/ARE in metabolic reprogramming (Citation46,Citation47).

Methodology

A comprehensive literature search using the Google Scholar, PubMed and Science Direct databases was performed to retrieve information related to the cancer protective effects of 21 selected dietary polyphenols via modulation of Keap1/Nrf2/ARE and other interconnected signaling pathways/proteins (MAPK/ERK1/2, PI3K/Akt, PKD, JNKs, AMP-activated protein kinase AMPK, NF-κB). Information on the anti-inflammatory and cytoprotective effects caused by the selected dietary polyphenols following Keap1/Nrf2/ARE modulation was also collected. The polyphenols were selected based on their various roles in restricting different cancer types, their abundance in dietary foods, and the amount of published evidence for their cancer protective role compared to other compounds. The keywords used for this search included ‘dietary polyphenols’, ‘antioxidant’, ‘free radical scavenger’, and ‘Nrf2’, ‘Keap1/Nrf2’, ‘Keap1/Nrf2/ARE’, ‘MAPK’, ‘PI3K’, ‘protein kinase B or Akt’, ‘ERK’, ‘AMPK’, ‘NF-κB’ and ‘cancer’. A further search using the keywords ‘anti-inflammatory’, ‘hepatoprotective’ and ‘cytoprotective’ was performed to collect information on additional Nrf2-modulated effects caused by the selected dietary polyphenols. Only research articles detailing in-vitro and in-vivo studies as well as clinical trials (Citation419 in total) published in high-quality peer-reviewed journals between 2001–2022 were selected for the write-up of this review.

Cancer Protective Role of Plant-Based Dietary Foods via Modulation of Keap1/Nrf2/ARE and Interconnected Signaling Pathways

Multiple in-vitro and in-vivo studies have linked the effects of various cancer chemopreventive agents with activation of the Keap1/Nrf2/ARE pathway (Citation48). Nrf2 activators exert cancer chemopreventive activity as they inhibit the metabolic activation of pro-carcinogens, block their reaching of target sites, prevent their interactions with cellular macromolecules (DNA, RNA, and proteins) as well as induce detoxification and increase the production of antioxidant enzymes (Citation49). Various plant-derived natural products, including several from dietary sources, can modulate the Keap1/Nrf2/ARE pathway. This includes structurally-diverse phytoconstituents, such as isothiocyanates (Citation50), garlic-derived organosulfur compounds (Citation51), indoles from cabbage and broccoli (Citation52), terpenoids from Citrus fruits and other dietary sources (Citation53–55), and many aromatic and phenolic derivatives (Citation56). Some of these natural products, such as sulphoraphane (in broccoli, Brussels sprouts, cabbage), curcumin (in turmeric), resveratrol (in grapes) or trigonelline (in coffee and fenugreek seeds), have been used as templates for the development of future cancer chemopreventive or anticancer drugs (Citation13). The common dietary sources of the discussed polyphenols are presented in .

Table 1. Common dietary sources of the 21 selected polyphenols.

Selection of Polyphenols with Cancer Protective Activity via Modulation of the Keap1/Nrf2/ARE and Interconnected Signaling Pathways

The cancer protective role of 21 dietary polyphenols () in the modulation of Keap1/Nrf2/ARE and interconnected signaling pathways/proteins is described below and summarized in and .

Figure 1. Chemical structures of the dietary polyphenols that modulate Keap1/Nrf2/ARE and interconnected signaling pathways. Here, (1) xanthohumol, (2) punicalagin, (3) resveratrol, (4) methyleugenol, (5) 6-shogaol, (6) chlorogenic acid, (7) ferulic acid, (8) carnosic acid, (9) carnosol, (10) ellagic acid, (11) apigenin, (12) catechin, (13) epicatechin, (14) EGCG, (15) fisetin, (16) genistein, (17) isoorientin, (18) quercetin, (19) luteolin, (20) rutin, and (21) kaempferol.

Figure 1. Chemical structures of the dietary polyphenols that modulate Keap1/Nrf2/ARE and interconnected signaling pathways. Here, (1) xanthohumol, (2) punicalagin, (3) resveratrol, (4) methyleugenol, (5) 6-shogaol, (6) chlorogenic acid, (7) ferulic acid, (8) carnosic acid, (9) carnosol, (10) ellagic acid, (11) apigenin, (12) catechin, (13) epicatechin, (14) EGCG, (15) fisetin, (16) genistein, (17) isoorientin, (18) quercetin, (19) luteolin, (20) rutin, and (21) kaempferol.

Figure 2. Cancer protective role of dietary polyphenols via modulating Keap1/Nrf2/ARE and interconnected signaling pathways.

Figure 2. Cancer protective role of dietary polyphenols via modulating Keap1/Nrf2/ARE and interconnected signaling pathways.

Table 2. Cell signaling pathways and proteins targeted by the 21 selected dietary polyphenols with a cancer protective role.

Kaempferol

The prenylated chalcone xanthohumol (1) is the main flavonoid of the female inflorescences of hops (Humulus lupulus L.) and is found in beer (Citation57). This compound exhibits a range of biological effects including cardioprotective, antioxidant, anti-inflammatory, antiviral, anti-obesity, as well as cancer chemopreventive and anticancer activity (Citation221).

Xanthohumol activates the Keap1/Nrf2/ARE pathway by inducing the translocation of Nrf2 from the cytosol to the nucleus, as well as its binding to ARE (Citation84). This activation upregulates the expression of several antioxidant enzymes, leading to decreased ROS generation and reduced GSH depletion (Citation222,Citation223). Xanthohumol has been shown to reduce cisplatin-induced inflammation and oxidation via activation of the Nrf2 pathway and upregulation of HO-1 expression (Citation224), leading to a pro-apoptotic effect as apoptosis can be induced by inhibition of inflammation and oxidation (Citation225). It has demonstrated cytoprotective and cancer chemopreventive activity on normal THLE-2 hepatocytes and hepatocellular carcinoma HepG2 cells via activating the Nrf2 pathway, upregulating phase-II enzymes, such as GSTs, HO-1, and NQO1 combined with p53 induction (Citation84). In PANC-1 and Mia-Pa-Ca-2 pancreatic cancer cells, the anti-inflammatory and pro-apoptotic effects of xanthohumol have been reported to reduce cell proliferation via Nrf2 activation and to increase the expression of antioxidant and detoxifying genes (SOD, NQO1, and GSTP) (Citation85). Both in-vivo and in-vitro studies have showed that a combination of xanthohumol and phenethyl isothiocyanate induces apoptosis by modulating Nrf2 and abrogating NF-κB, signal transducer and activator of transcription 3 (STAT3), and the Akt/P70S6K signaling pathways in PSN-1 cancer cells. The same combination showed cytoprotective activity via activation of Nrf2 in non-cancerous MS1 cells (Citation86). Xanthohumol ameliorates lipopolysaccharide (LPS)-induced lung injury, increasing the expression of antioxidative enzymes via Nrf2 activation associated with AMPK and glycogen synthase kinase-3 beta (GSK3β) phosphorylation. This suppressed the LPS-activated Txnip/NLR family pyrin domain containing 3 (NLRP3) inflammasome and the NF-κB signaling pathway (Citation222,Citation226). In myeloid leukemia, the activation of PI3K/Akt, NF-κB, and other signaling pathways causes proliferation, transformation, and resistance to apoptosis (Citation227). Xanthohumol is a potent inhibitor of T cell proliferation, cytokine production, and T cell-mediated cytotoxicity through the inhibition of NF-κB (Citation228). It targets cell growth and angiogenesis, and causes impaired migration and invasion both in-vivo and in-vitro models of acute and chronic myeloid leukemia through inhibition of the PI3K/Akt and NF-κB signaling pathways (Citation87). The inactivation of NF-κB can reduce the excessive production of inflammatory mediators, such as NO, IL-1β and tumor necrosis factor-α (TNF-α) in LPS-induced microglial BV-2 cells, thus exerting an anti-inflammatory effect (Citation229). Xanthohumol also shows anticancer activity by downregulating topoisomerase-I and the expression of efflux drug transporters through induction of both caspase-dependent and caspase-independent apoptosis (Citation230,Citation231). It has been reported to significantly alleviate the proliferation of HCT116-derived human colon cancer cells upon inducing apoptosis via activation of death-receptors as well as mitochondrial-mediated programmed cell death pathways by downregulation of Bcl-2 and activation of the caspase cascades (Citation88). In human prostate cancer cells, it induces apoptosis by activating the pro-apoptotic proteins Bax and p53, decreasing the activation of anti-apoptotic NF-κB (Citation89) and inhibiting Akt, NF-κB, p-mammalian target of rapamycin (mTOR) and NF-κB-regulated anti-apoptotic proteins Bcl-2 and survivin (Citation90).

Punicalagin

The ellagitannin punicalagin (2) is the main bioactive constituent of pomegranate (Punica granatum) fruit, husk, and juice (Citation58,Citation59). This compound has antioxidant, anti-inflammatory (Citation232), antiviral (Citation233), anti-atherosclerotic (Citation234), antimicrobial (Citation235), antiquorum sensing (Citation236) and antiproliferative activity (Citation237).

Punicalagin activates the Nrf2 pathway by decreasing Keap1 levels via activation of the PI3K/Akt pathway which causes the dissociation of Nrf2 from Keap1 and increases nuclear translocation of Nrf2, thereby inducing HO-1 expression. These events, along with the reversal of the LPS-induced reduction of SOD1 mRNA expression, give punicalagin its protective effect against LPS-induced oxidative stress (Citation238). Punicalagin has been demonstrated to protect from heat stress-induced intestinal epithelial cell damage and cell death (Citation239). It also acts as an anti-mutagenic agent by inhibiting DNA adducts caused by the induction of phase-II enzymes (Citation240). The upregulated HO-1 is at the core of the Nrf2-mediated NF-κB inhibition; the pathway involved in the cytokine production (Citation241). Whilst activation of the NF-κB pathway plays a pivotal role in tumor cell progression, growth, proliferation and resistance to apoptosis, inactivation of NF-κB plays a cancer protective role as it activates the Keap1/Nrf2/ARE pathway (Citation225). The inhibition of the NF-κB pathway by punicalagin can impede cancer cell proliferation and enhance cell apoptosis in cervical cancer ME-180 cells and osteosarcoma (Citation91,Citation92). A tumor xenograft mouse model shows that punicalagin is able suppress the growth of osteosarcoma and inhibit angiogenesis via suppression of NF-κB activation (Citation92). The TNFR-induced Akt activation, required for NF-κB activity, is also abrogated by punicalagin which is responsible for inhibiting cell proliferation and induction of apoptosis in human colon cancer cells (Citation93,Citation94). Studies have also shown that punicalagin exerts its anti-inflammatory activity via inhibition of the NF-κB and MAPK pathways mediated by toll-like receptor 4 (TLR4) mRNA expression. This anti-inflammatory effect is also mediated via the inhibition of the production of pro-inflammatory cytokines and other factors such as nitric oxide (NO), prostaglandin E2 (PGE2), IL-1β, IL-6, and TNF-α (Citation242). The Keap1-Nrf2 activation induced by punicalagin has been shown to reduce palmitate-induced lipotoxicity, including attenuation of mitochondrial membrane potential loss, ATP depletion and ROS generation by the ERK/Nrf2 pathway and increase the viability of hepatocytes by blocking mitochondria-mediated caspase-dependent apoptosis (Citation58).

Resveratrol

Resveratrol (3) is a stilbene that has been identified in more than 70 plant species, and predominantly occurs in grapes’ skin and seeds, as well as in red wines (Citation60). Resveratrol has antioxidant, anti-inflammatory, and anticancer activity (Citation211).

It activates the Nrf2 signaling pathway by causing dissociation of Nrf2 from Keap1 and its translocation into the nucleus, leading to the activation of ARE-driven gene promoters (Citation243). This activation of ARE downstream genes causes the scavenging of ROS that are responsible for DNA damage and the activation of phase-II enzymes (GST and HO-1) resulting in the detoxification of carcinogens (Citation99,Citation244). Resveratrol has been reported to protect porcine intestinal epithelial cells from oxidative stress through the PI3K/Akt-mediated Nrf2 signaling pathway by upregulating HO-1, SOD-1, and catalase (CAT) expression levels (Citation245). It has been shown to prevent the formation of 17,β-estradiol-induced breast tumors by upregulating antioxidant genes NQO1, SOD3, and 8-oxoguanine glycosylase (OGG1) and thereby protects cells against oxidative DNA damage (Citation95). The upregulation of NQO1 expression caused by resveratrol can protect leukemic cells from DNA adducts formation (Citation246) and to decrease estrogen metabolism in 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD)-induced breast cancer MCF-10F cells (Citation247). As an Nrf2 activator, resveratrol has been shown to decrease the growth and clonogenic potential of breast cancer MCF-7 cells exposed to doxorubicin, resulting in the inhibition of apoptosis, autophagy, and inflammation. In the ovaries, resveratrol can decrease the oxidative stress caused by chemotherapy by upregulating Nrf2 and activating the sirtuin 1 (SIRT1)/forkhead box protein O1 (FOXO1) pathway. In the presence of the cytotoxic agent sitagliptin, resveratrol was shown to ameliorate renal cell carcinoma showing antioxidant activity through restoring Nrf2/HO-1 content (Citation96). Resveratrol also attenuates diethylnitrosamine (DENA)-induced liver tumorigenesis by modulating Nrf2 signaling pathway and suppressing oxidative stress and inflammatory cytokines (Citation97). It inhibits ROS production by activating the Nrf2 pathway, reducing the expression of Keap1 and inhibiting NF-κB activation (Citation248). The latter is caused by the inhibitor of nuclear factor kappa-B kinase subunit beta (IKKβ)-mediated IκBα phosphorylation by resveratrol (Citation249). A number of reports have shown that NF-κB is an important inflammatory transcriptional regulator that can be activated by ROS (Citation250). Thus, resveratrol acts as an anti-inflammatory agent by blocking IL-1β, TNF-α and NF-κB activation. These inflammatory factors play a role in the development of carcinogenesis including tumor growth, angiogenesis, invasion, and metastasis (Citation251). Resveratrol has shown apoptotic and antiproliferative activity on human prostate cancer cells mediated by NF-κB inhibition (Citation98). Resveratrol also causes the accumulation of Nrf2 in the cytoplasm and inhibits Nrf2-dependent transcription via SIRT1 deacetylase activity, in both K562 leukemia and HepG2 hepatocellular carcinoma cells, demonstrating its effect in both Nrf2 accumulation and translocation (Citation99). In-vivo studies carried out on resveratrol showed that it suppresses pancreatic cancer (by abrogating PI3K and Akt phosphorylation), liver cancer (by enhancing Nrf2 expression), colorectal cancer (by increasing Bax expression), and breast cancer (by suppressing tumor growth factor-β1 (TGF-β1) and NF-κB expression) (Citation100). A number of clinical trials have been conducted to evaluate the effectiveness of resveratrol on patients with prostate, colorectal and breast cancer. It was concluded that resveratrol was an unlikely candidate for prostate cancer, but showed a very slight effect on colon cancer and a promising effect on breast cancer, respectively (Citation252).

Methyleugenol

Methyleugenol (4) is a phenylpropene found in the essential oils of clove, sweet basil, lemon grass, nutmeg and pimento (Citation61). This compound is commonly used as a flavoring additive (Citation253). It has anti-allergic, antinociceptive, antioxidant, anti-inflammatory (Citation62,Citation63), and anticancer (Citation254) activity. It can also ameliorate cerebral ischemic injury (Citation255).

Studies have revealed that methyleugenol can enhance the nuclear translocation of Nrf2, reduce Keap1 levels, and increase ARE activity (Citation256). This compound has demonstrated antioxidant activity on different cell lines by scavenging ROS, decreasing superoxide generation along with increasing SOD, CAT and GSH (Citation254). Its beneficial effects in ischemic cell injury both in-vivo and in-vitro have been attributed to its ability to scavenge ROS and generate an antioxidant effect via Nrf2/ARE activation (Citation254). It also upregulates the expression of antioxidant enzymes, such as glutamate cysteine ligase catalytic (GCLC), glutamate cysteine ligase modifier (GCLM), HO-1, and NQO1 by activating the Nrf2/ARE signaling pathway via activation of the AMPK/GSK-3β and ERK pathways (Citation257). Methyleugenol was found to have anticancer activity in cervical cancer individually or in combination with cisplatin by inhibiting cell growth, inducing cell apoptosis, mitochondrial membrane potential loss and caspase 3 activity (Citation101).It has also been reported to reduce cell viability, impede the expression of the PI3K/mTOR/Akt pathway and induce G2/M cell cycle arrest in human retinoblastoma cancer cells (Citation102). It has anticancer activity against leukemia cells (HL-60) and human colon carcinoma cells (Citation103). Methyleugenol also exerts an anti-inflammatory effect by suppressing the release of pro-inflammatory cytokines, such as TNF-α, IL-4, PGE2, PGD2, leukotriene B4 (LTB4), and LTC4 via suppressing the activation of spleen tyrosine kinase (Syk), ERK1/2, p38, JNK, c phospholipase A2 (PLA2), and 5-lipoxygenase (5-LO) (Citation62). To the best of our knowledge, no in-vivo studies have been carried out to investigate the cancer protective role of methyleugenol via modulation of Keap1/Nrf2/ARE and interconnected pathways.

6-Shogaol

The polyphenol 6-shogaol (5) is the main bioactive constituent of dried or cooked ginger (Zingiber officinale). This compound displays a range of biological properties, including anti-inflammatory, analgesic, antipyretic, cancer chemopreventive and antioxidant activity (Citation64).

It has been shown to decrease Keap1 levels by upregulating, phosphorylating, and translocating Nrf2 in-vitro and in-vivo, resulting in the upregulation of Nrf2 target genes including aldo-keto reductase family one member B10 (AKR1B10), Ferritin Light Chain (FTL), gamma-glutamyltransferase-like activity 4 (GGTLA4), HO-1, GCLC and GCLM. This activation of Nrf2 signaling enhances cellular antioxidant activity, GSH levels and ARE promoter activity (Citation258). 6-shogaol activates Nrf2 via the JNK-mediated pathway, causing an increase in JNK activation and in the expression of γ-glutamylcysteine synthetase (GCS) and HO-1 (Citation259,Citation260). 6-shogaol also activates Nrf2 in PC12 cells and provides cytoprotection against oxidative stress (Citation261). An analog of 6-shogaol, named 3-phenyl-3-shogaol (3-Ph-3-SG), has also been reported to provide cytoprotection by inducing the ARE-driven genes NQO1 and HO-1 and abrogate cancer cell invasion by suppressing NF-κB signaling (Citation262). The latter causes inhibition of PMA-stimulated MDA-MB-231 breast cancer cell invasion via downregulating MMP-9 expression (Citation104). Besides this, the blocking of Akt and downstream targets (including the mTOR, forkhead transcription factors (FKHR) and GSK-3β) by 6-shogaol has been reported to inhibit the survival of non-small cell lung cancer (NSCLC) cells (Citation105). Here, 6-shogaol induces apoptosis by causing cell cycle arrest in G1 or G2/M phase. Its suppression of the Akt kinase activity results in reduced STAT3 activity and decreased expression of cyclin D1/3, and Akt signaling (Citation263). The inhibition of STAT3 activity and NF-κB signaling by 6-shogaol has an effect on prostate cancer cells too, both in-vivo and in-vitro. 6-Shogaol causes a decrease in NF-κB target genes and protein levels, including cyclin D1, survivin, and cMyc, and modulates the mRNA levels of chemokines, cytokines, cell cycle, and apoptosis regulatory genes, such as IL-7, chemokine (C-C motif) ligand 5 (CCL5), Bax, Bcl2, p21, and p27 (Citation106). In human pancreatic tumors, 6-shogaol blocks the growth of tumor cells and acts as adjuvant to potentiate gemcitabine via suppression of TLR4/NF-κB-mediated inflammatory pathways linked to tumorigenesis (Citation107). 6-shogaol has an effect on pathways which is regulated for the prevention of carcinogenesis. Thus, the over expression of MAPKs (ERK1, JNK1 & p38) on UVB-induced HaCaT cells is minimized by 6-shogaol, thereby protects against UVB-induced oxidative skin damage (Citation264). This compound also causes attenuation of several pro-inflammatory mediators in response to UVB, such as inducible nitric oxide synthase (iNOS), cyclooxygenase-2 (COX-2), MMP, cell adhesion molecules, chemokines and cytokines (Citation258). As an anti-inflammatory agent, 6-shogaol has been shown to increase HO-1 levels by attenuating COX-2, iNOS, NF-κB and MAPK signaling. It also helps in reducing edema by inhibiting leukocyte infiltration into inflamed tissue (Citation265). Interestingly, one study has indicated that a novel potent Nrf2 activator molecule called (1E,4E)-1-(4-hydroxy-3-methoxyphenyl)-7-methylocta-1,4,6-trien-3-one (SA), synthesized from 6-shogaol acted as a Michael acceptor to cause dissociation of Nrf2 from Keap1, inhibit Nrf2 ubiquitination and activate the Nrf2 response (Citation266). Finally, 6-shogaol exerts activity against human endometrial cancer via mediating ROS generation both in-vitro and in-vivo (Citation108). It is also able to inhibit metastasis in endometrial carcinoma by triggering PI3K/Akt signaling both in-vitro and in-vivo (Citation109). Other in-vivo studies have established the activity of 6-shogaol in prostate and lung cancer (Citation106,Citation110).

Chlorogenic Acid

The polyphenol chlorogenic acid (6), also known as 3-caffeoylquinic acid (3-CQA), is mainly found in coffee (Coffea canephora, Coffea arabica) and mate (Ilex paraguariensis) plants as well as in some fruits and vegetables (Citation65). It exerts antioxidant, anti-inflammatory (Citation267) and anti-carcinogenic activity (Citation119).

Several chlorogenic acid derivatives are found in plants. These include 3-caffeoylquinic acid (3-CQA), 4-caffeoylquinic acid (4-CQA), 5-caffeoylquinic acid (5-CQA), 3,4-dicaffeoylquinic acid (3,4-diCQA), 3,5-dicaffeoylquinic acid (3,5-diCQA), and 4,5-dicaffeoylquinic acid (4,5-diCQA) which interact with Keap1-Nrf2 complex to activate Nrf2 signaling (Citation268). This activation leads to an increased nuclear translocation of Nrf2 followed by enhanced expression of ARE-dependent genes coding for the phase-II enzymes GST, γ − GCL, NQO1, and HO-1 in colon carcinoma HT29 cells (Citation111). Pretreatment with chlorogenic acid has been shown to protect from CCl4-induced liver injury via activation of the Nrf2 pathway and suppression of NLRP3 inflammation (Citation269). The activation of Nrf2/ARE also interacts with the phosphorylation of IĸB and suppresses the activation of NF-ĸB (Citation270). In LPS-induced inflammation, chlorogenic acid alleviates the symptoms and inhibits the LPS-induced oxidative stress through modulation of the NF-κB/MAPK pathway and nuclear translocation of Nrf2, respectively (Citation271). In breast cancer, chlorogenic acid enhances antitumor immunity, exerts antitumor and anti-metastatic effects by impairing the NF-κB/EMT (Citation120) and β-catenin of wingless-related integration site (Wnt) signaling pathways (Citation272). Chlorogenic acid is also reported to kill MDA-MB-231 and MCF-7 cells by binding and stimulating the translocation of PKC, an important molecule for malignant tumor (Citation112). Chlorogenic acid protects against cellular oxidative damage and renal cell carcinoma by activating Nrf2/ARE and modulating the PI3K/Akt pathway (Citation113). It has been shown to exert an inhibitory effect on the PI3K/Akt/mTOR pathway in A498 kidney cancer cells (Citation114). Chlorogenic acid also induces p38 MAPK and JNK gene expression, affecting apoptosis-related genes that are involved in oxidative stress in lung cancer cells (Citation115). It also has the ability to induce ROS generation and exert cytoprotective effects on human colon cancer cells (Citation116). The generation of ROS promotes anti-tumorigenic signaling and triggers oxidative stress–induced cancer cell death showing the dual characteristics of ROS (Citation273). Chlorogenic acid also acts as a chemosensitizer of 5-fluorouracil (5-FU) chemotherapy, displaying a synergistic effect in combination with 5-FU, and inactivates ERK through the overproduction of ROS in HepG2 and Hep3B hepatocellular carcinoma cells (Citation117). It has in-vitro activity against the proliferation of A549 human cancer cells and the 12-O-tetradecanoylphorbol-13-acetate (TPA)-induced neoplastic transformation of JB6 P + cells (Citation274). Both in-vitro and in-vivo studies have reported that chlorogenic acid is active against lung cancer via disruption of the binding of annexin A2 to the p50 subunit of NF-κB (Citation118). Chlorogenic acid also has in-vitro and in-vivo anti-angiogenic activity via inhibition of HIF-1α/Akt signaling (Citation275). It has been demonstrated to suppress the proliferation of HepG2 hepatocellular carcinoma cells by inactivating ERK1/2, MMP-2 and -9 expression in a xenograft model (Citation119). Studies have also reported the effect of chlorogenic acid in-vitro and in-vivo against breast cancer via inhibition of NF-κB, VEGF, EGF, IL-10, TGF-β, and CD34 (Citation120,Citation121).

Ferulic Acid

Ferulic acid (7) a caffeic acid derivative found abundantly in vegetables (aubergines, tomatoes, artichokes), fruits, grains, and some beverages. This compound displays a range of biological effects in diseases that are linked with cancer, Alzheimer’s disease, diabetes mellitus, skin and cardiovascular disorders (Citation66).

Ferulic acid modulates the Keap1-Nrf2 pathway, causing dissociation of the Keap1/Nrf2 complex, increases Nrf2 transcription (Citation276) and therein enhances expression of Nrf2-mediated phase-II enzymes, including NQO1, GSTA2 and SOD (Citation277). It has been found to reduce serum TNF-α and IL-1β, hepatic NF-κB, p65, Bax, and caspase-3, and increase Bcl-2, Nrf2, NQO1, HO-1, and PPARγ, thereby attenuate oxidative stress, inflammation, and cell death (Citation278). It causes activation of Nrf2 and enhances ARE promoter activity via phosphorylating ERK1/2 in PC-12 cells (Citation279). The activation is caused by induction of Nrf2 nuclear translocation and transcriptional activity which significantly upregulates the HO-1 mRNA and protein expression controlled by the ERK signaling pathway (Citation280). The activation of Nrf2/HO-1 increases antioxidant defenses via activating ARE-mediated genes leading to ROS scavenging and protection against γ-radiation-induced oxidative stress (Citation281). NF-κB signaling is responsible for the development and progression of several human cancers (Citation282). Blocking this pathway and activating antioxidative enzymes via Nrf2/ARE signaling abrogates the initiation of carcinogenesis. Ferulic acid can inhibit IL-6 and NF-κB promoter activity upon reduction of the nuclear translocation of Nrf2 and NF-κB through reduced expression of phosphorylated IKK (Citation283). In cervical cancer cells, ferulic acid induces MMP-9 mRNA expression and G0/G1 phase blocking by increasing p53 and p21 along with decreasing autophagy related proteins (Citation122). On the other hand, in HCT 15 colorectal cancer cells, it downregulates the human epidermal growth factor (EGF) receptor that is a vital for inducing colorectal cancer (Citation123). It protects PC-3 prostate cancer cells by inducing apoptosis and cell cycle arrest (Citation124). It exhibits antitumor activity against breast tumor cells both in-vitro and in-vivo through suppression of EMT (Citation125). Ferulic acid, when administered together with poly (ADP-ribose) polymerase (PARP) inhibitors, increases breast tumor cells sensitivity to PARP inhibitors (Citation284), exerting their protective effects through the regulation of various signaling pathways, including MAPKs, HIF-1α, Nrf2 and NF-κB (Citation285). It has also been reported to reduce the LPS-induced overexpression of pro-inflammatory enzymes, such as iNOS and the subsequent excessive production of NO and cyclic guanosine monophosphate (cGMP) in intestinal Caco-2 cell monolayers. Two ferulic acid derivatives, iso-ferulic acid and dihydroferulic acid, and their glucuronidated and sulfated metabolites can decrease the nuclear translocation of NF-κB by promoting Nrf2 expression and controlling the activation of MAPK, p38, ERK and Akt (Citation286). In LPS-activated RAW 264.7 mouse macrophages, ferulic acid inhibits the expression of inflammatory mediators, such as IL-6, TNF-α and iNOS and activates the expression of antioxidant metallothioneins (MT-1, MT-2) (Citation283). A recent study used in-vivo and in-vitro models to assess the effect of ferulic acid lipid nanocapsules in colorectal cancer. These showed potential activity via suppressing the expression of cyclin D1, insulin-like growth factor II (IGF II), and VEGF, as well as via auto-regulating the apoptotic/anti-apoptotic gene BAX/Bcl-2 (Citation126). Ferulic acid also exhibits anti-angiogenic and antitumor potential against melanoma, both in-vivo and in-vitro, via blocking PI3K-Akt signaling (Citation127).

Carnosic Acid

Carnosic acid (8) is a major polyphenol in rosemary (Rosmarinus officinalis) with various biological properties, such as antioxidant, anti-inflammatory, neuroprotective and anticarcinogenic activity (Citation67). Carnosic acid contains a catechol group that converts into electrophilic quinones upon oxidation, and this electrophilic nature helps in interacting with the cysteine residues of Keap1, activating the Keap1/Nrf2/ARE pathway and the synthesis of antioxidant enzymes (Citation287).

Along with its effect on the Keap1/Nrf2/ARE pathway, the cytoprotective effect of carnosic acid also involves inhibition of the PI3K/Akt and NF-κB pathways (Citation288,Citation289). NF-κB activation is a key factor involved in the release of pro-inflammatory cytokines and inflammation-associated cancer (Citation290). Both PI3K/Akt/mTOR and NF-κB pathways are important in chemoresistance and survival of cancer cells and are considered potential targets for cancer treatment (Citation291,Citation292). Carnosic acid has been reported to attenuate pro-inflammatory cytokine mRNA and protein levels in the colon of mice, as well as upregulate GSH and SOD activity and downregulate iNOS and MDA levels (Citation293). Carnosic acid has been found to downregulate the expression of COX-2 at both the mRNA and protein levels in Caco-2 human colorectal cancer cells (Citation128). This COX-2 inhibition is interrelated with NF-κB inactivation (Citation294). One study revealed that carnosic acid, encapsulated into albumin nanoparticles, is able to mediate apoptosis to in MCF-7 breast and Caco-2 colorectal cancer cells (Citation129). This effect is induced by upregulation of the expression of GCLC, and COX-2, and downregulation of Bcl-2. The modulation of NF-κB by carnosic acid in hepatocarcinoma is mediated by Akt signaling. This was confirmed in-vitro and in-vivo (Citation130). By modulation of the Akt/mTOR signaling pathway, carnosic acid also prevents proliferation and survival of human gastric cancer cells (Citation131). This inactivation of the Akt/mTOR pathway contributes to the induction of autophagic cell death in hepatoma cells (Citation295). Carnosic acid enhances the apoptosis of human colorectal carcinoma cells via generation of ROS, induction of p53, activation of caspases and modulation of Janus kinase 2 (JAK2)-STAT3/Src-STAT3 signaling pathway (Citation132). ROS generation also promotes anti-tumorigenic signaling and stimulates oxidative stress-induced cancer cell death (Citation273). In human cervical cancer cells, carnosic acid induces ROS production, which phosphorylates the JNK, activates endoplasmic reticulum stress, and induces apoptosis. In addition, a notable decrease in tumor formation was observed in cancer cells xenografted mice after administration of carnosic acid for five weeks (Citation133). The induction of apoptosis by carnosic acid also enhances the effect of tamoxifen in breast cancer cells via caspase-3/TNF-related apoptosis-inducing ligand (TRAIL) activation. The combination of carnosic acid and tamoxifen also led to breast tumor suppression in a mouse xenograft model (Citation134). Carnosic acid also acts in synergy with curcumin to activate the expression of antioxidants (AKR1C2, HO-1) and apoptotic genes (GDF15, PHLDA1, DDIT3) as well as inhibit the cell cycle genes (CDKN2C) (Citation296). Carnosic acid, in combination with fisetin, induces apoptosis in lung cancer cells via activation of caspase-3 (Citation297). Carnosic acid loaded into transferrin-conjugated liposomes have been reported to mediate apoptosis in liver cancer, both in-vivo and in-vitro, via upregulating the expression of cleaved PARP, caspase-3 and -9, and downregulating the expression of Bcl-2 (Citation135).

Carnosol

Carnosol (9) is a phenolic diterpene which is present in rosemary (Rosmarinus officinalis) (Citation68). This molecule has been found to increase the nuclear translocation and accumulation of Nrf2 as well as ARE activity resulting in the induction of phase-II enzymes and the increased expression of the GSH synthesis enzyme subunit GCLC/GCLM (Citation298). Carnosol also enhances HO-1 activity in both mRNA and protein levels by causing nuclear translocation of Nrf2, increasing its binding to ARE and inducing Nrf2-dependent activation of the HO-1 promoter region (Citation299). Being a strong Nrf2 activator, it also induces NQO1 and can modulate other intra and extracellular signaling in addition to activation of Nrf2/HO-1 signaling (Citation300). Carnosol activates the ERK, p38 and JNK pathways with PI3K driven survival pathway in PC-12 cells (Citation301). One study revealed that a carnosol-containing rosemary extract was able to suppress the growth of HCT116 colon cancer cells by upregulating Nrf2 through the protein kinase RNA-like endoplasmic reticulum kinase (PERK)/Nrf2/Sestrin-2 mediated pathway (Citation136). Carnosol can also inhibit EGF-induced epithelial to mesenchymal transition which enables cancer cells to become invasive and undergo metastasis via inhibiting the phosphorylation of ERK (Citation302). Carnosol also protects prostate cancer PC3 cells via modulation of the PI3K/Akt/mTOR and AMPK signaling pathways (Citation137). It has antitumorigenic activity against human colon cancer cells, reducing STAT3 signaling and ROS generation (Citation303). The latter promotes anti-tumorigenic signaling, oxidative stress-induced apoptosis (Citation273), and blocks the growth, invasion and migration of MDA-MB-231, Hs578T, MCF-7, and T-47D breast cancer cells via targeting STAT3 signaling (Citation138). Carnosol has demonstrated a cytoprotective effect against H2O2 in HepG2 cells (Citation298). It has also been demonstrated that the activation of Nrf2 in HepG2 cells is caused by increased expression of sestrin2 and MRP2 (Citation139). In stress situations, the sestrin (1–3) family can interact directly with AMPK to maintain genomic integrity and suppress tumorigenesis (Citation304). Carnosol can enhance the effect of curcumin on cancer cells via reducing the phosphorylation of ERK1/2, Akt and STAT3 (Citation302). Although one recent review included in-vitro and in-vivo studies on the effects of carnosol against cancer (Citation305), this did not include any in-vivo work on its role in the Keap1/Nrf2/ARE and interconnected pathways.

Ellagic Acid

Ellagic acid (10) is a polyphenol abundant in fruits and nuts, particularly cranberries, raspberries, walnuts and pecans (Citation69).

Ellagic acid increases Nrf2 translocation and HO-1 activity in cells by downregulation of mRNA and Keap1 protein levels (Citation306). This Nrf2/HO-1 activation by ellagic acid modulates Akt and ERK phosphorylation, thereby preventing oxidative stress in HepG2 cells (Citation307). The activation of Nrf2 also suppresses ROS and MDA levels while enhancing GSH and Mn-SOD levels (Citation308). In UVA stimulated HaCaT keratinocytes, ellagic acid induces autophagy by enhancing the expression of γ-GCLC, HO-1, and NQO1 proteins via Nrf2 activation and other signaling pathways, such as ERK, JNK, and PI3K/Akt/mTOR (Citation309). Its antitumor effect against endometrial cancer, via inhibition of PI3K and MMP9 expression, was reported in-vitro and in-vivo (Citation140). In MDA-MB-231 breast cancer cells, it acts as an anti-angiogenetic agent, limiting the development and migration of cells via inhibition of the PI3K/Akt and MAPK pathway of VEGFR 2. In MDA-MB-231 xenografted animal models, ellagic acid reduced tumor growth via suppressing P-VEGFR2 expression (Citation141). Ellagic acid decreases the proliferation and development of prostate cancer PC3 cells via downregulating the phosphorylated STAT3, ERK and Akt signaling proteins (Citation142). In Mia PACA-2 and PANC-1 human pancreatic carcinoma cells, ellagic acid has an antiproliferative effect, causing apoptosis via inhibition of the NF-κB pathway (Citation143,Citation310). In PANC-1 tumor-bearing mice, it was found to reduce tumor growth via downregulation of COX-2 and NF-κB expression (Citation143). Ellagic acid exerts a hepatoprotective effect against methotrexate-induced toxicity in-vivo by upregulating Nrf2 and HO-1 expression and inhibiting the NF-κB signaling pathway and the overproduction and expression of inflammatory factors (Citation311). In T24 human bladder cancer cells, it induces apoptosis via increasing G0/G1 phase cell cycle arrest and upregulating p53 and p21 expression (Citation144). In human bladder cancer xenografted animal models, ellagic acid demonstrates notable decrease in tumor growth rate, infiltrative behavior and tumor-associated angiogenesis, with upregulation and downregulation of VEGF-A and VEGFR-2 expression, respectively (Citation312). In lung cancer, ellagic acid abolishes mitochondrial respiration and tumor growth both in-vivo and in-vitro via upregulating AMPK and downregulating HIF-1α expression (Citation145).

Apigenin

Apigenin (11) is a flavonoid that constitutes with the aglycone moiety of many naturally-occurring glycosides, including apigetrin, vitexin, and isovitexin. It is widely distributed in many plants and vegetables, such as parsley, celeriac, celery, and in chamomile tea (Citation70). It possesses different biological effects including anti-inflammatory, antioxidative, antitumor properties (Citation313).

It protects against oxidative stress through the upregulation of the antioxidant HO-1 enzyme mediated via Nrf2 signaling (Citation146,Citation314). The α, β unsaturated ketone moiety of apigenin interacts with the cysteine residues in Keap1, leading to activation of the Keap1-Nrf2-ARE system (Citation134). Apigenin induces ARE activity, enhances Nrf2 mRNA and protein levels and the expression of HO-1 in HepG2-C8 cells. This ARE activation is mediated through different signaling pathways including MAPK, p38, ERK 1/2 and JNK (Citation315). In human melanocytes, it exerts antioxidative activity via increasing levels of SOD, CAT and GPx enzymes. In another study using melanocytes under hydrogen peroxide-induced oxidative stress, apigenin enhances cell viability, SOD, CAT, and GPx activities and inhibits MDA levels (Citation316). In hepatocellular carcinoma cells, apigenin inhibits cell proliferation and autophagy via suppressing the PI3K/Akt/mTOR pathway (Citation70). Skin carcinogenesis is also prevented by apigenin via restoration of the activation of Nrf2 and enhancing its downstream gene NQO1 as well as decreasing epigenetic proteins, such as DNA methyltransferase (DNMT) and histone deacetylases (HDAC) expression (Citation146). It can also protect hepatocytes from tBHP-induced stress by upregulating HO-1, GCLC, and GCLM gene transcription via the ERK2/Nrf2/ARE signaling pathway (Citation314). Overexpression of Nrf2 can cause chemoresistance and in this regard, apigenin shows Nrf2 inhibiting activity which is enhanced by hyaluronic acid-based nanostructured lipid carriers (NLCs). HAAPG-NLCs significantly decrease Nrf2, multidrug resistance-associated protein 2 (MRP2), HO-1 and Bcl-2 with an increase in Bid mRNA levels, thereby improving the efficacy of docetaxel in lung cancer (Citation147). Apigenin is also effective in doxorubicin-resistant hepatocellular carcinoma cells (Citation317). Several pharmaceutical preparations containing apigenin in combination with other natural products have been reported as useful for preventing or treating cancers (Citation318,Citation319). It can also upregulate the mRNA and protein expression of Nrf2 and its downstream genes via activation of the PI3K/Nrf2/ARE pathway. It exerts its anti-inflammatory effect by suppressing LPS-induced NO, iNOS, and cPLA2 (Citation315). Two different studies conducted on the activity of apigenin against prostate cancer report that it exerts this effect both in-vivo and in-vitro via inhibiting Akt signaling and inducing apoptosis (Citation148,Citation149). Apigenin also shows anti-colon cancer effects. This has been demonstrated in-vitro and on xenografted mice model via inhibition of the m-TOR/PI3K/Akt pathway (Citation150). Furthermore, apigenin has anti-leukemic activity. This has been observed in-vitro and in U937 xenografts via inactivation of Akt and activation of JNK expression (Citation151). In cervical cancer, xenograft models suggest evidence for the antitumor activity of apigenin which is able to reverse the abnormal estrogen receptor signal in tumor tissue. The underlying molecular mechanism of this effect is via suppression of the PI3K/Akt/mTOR signaling (Citation152).

Catechin

The polyphenol catechin (12) is found in green, white, and black tea (Camellia sinensis) (Citation71,Citation72). Green tea contains the highest amount of catechin making up to 25% of its leaf composition (Citation320).

Catechin interacts with the Nrf2 binding site of Keap1 thus suppressing the Keap1-Nrf2 interaction. It increases the ERK1/2 expression, promotes the phosphorylation of ERK1/2, enhances Nrf2 nuclear translocation and increases the expression of Nrf2-dependent genes including GCLC/GCLM, HO-1 and NQO1 (Citation321). In addition, catechin upregulates Nrf2 expression by inactivating the NF-κB signaling pathway (Citation322). This inactivation blocks anti-apoptotic gene Bcl-XL expression and increases apoptosis which contributes to the anticancer activity of catechin (Citation290). Catechin has also been demonstrated to synergize the effects of the anticancer drug 5-FU in-vitro (Citation323). The activation of the Keap1/Nrf2/ARE pathway by catechin leads to an antioxidative effect via the upregulation of Gpx, glutathione reductase (GR), and an increase in total sulfhydryl groups associated with high Nrf2 and HO-1 expression (Citation324). Interestingly, one study showed that catechin-derived metabolites (produced by the intestinal microbiota) when combined with curcumin could abrogate VEGF expression and suppress miR-210 and miR-21 oncogenic microRNAs, protecting against cervical cancer (Citation153). In addition, catechin and curcumin exert synergistic activity against human colon adenocarcinoma HCT 15, HCT 116, and human larynx carcinoma HepG-2 cell lines (Citation154). To the best of our knowledge, no In Vivo studies have been carried out to investigate the cancer protective role of catechin via modulation of Keap1/Nrf2/ARE and interconnected pathways.

Epicatechin

Among four diastereoisomers of catechin, (-)-epicatechin (13) is found in green tea, grapes and cocoa (Citation73,Citation74). It has antioxidant and anti-inflammatory activity and is effective against various diseases such as cancer, diabetes, cardiovascular disease, stroke and neurodegenerative disorders (Citation325).

Studies demonstrated that epicatechin increases the nuclear accumulation of Nrf2 by interacting with Keap1, therein upregulating the expression of phase II enzymes, and providing protection against oxidative injury (Citation326). Epicatechin increases GSH levels via activating the Nrf2 pathway and upregulating ARE-mediated HO-1 and NQO1 gene expressions (Citation327). Some studies have demonstrated the anticancer effect of epicatechin. For instance, it induces apoptosis via generating ROS and modulating pro-apoptotic proteins resulting in decreased viability in MCF-7 and MDA-MB-231 breast cancer cells (Citation155). Epicatechin can synergize the anticancer activity of panaxadiol on human colorectal cancer cells (Citation156). The synergistic effect has also been observed on human gastric carcinoma cells where epicatechin, along with (-)-epigallocatechin-3-gallate, exerts cytoprotective activity (Citation157). The exact mechanism behind this activity is not clear, but this might be exerted through modulation of various cell cycle regulatory pathways, including activation of the Keap1/Nrf2/ARE and modulation of the PI3K/Akt and ERK pathways (Citation328). In addition, epicatechin has been shown suppress the PI3K/Akt/mTOR signaling pathway to inhibit the prostate cancer cell migration (Citation158). It can also upregulate ERK1/2 and suppress MAPKs (JNK1/2 and p38) (Citation326), thus impacting on cell proliferation, differentiation, migration, senescence, apoptosis, and inflammation (Citation315). All of the pathways aforementioned are linked with cellular transformation, tumorigenesis, cancer promotion and progression (Citation329). Epicatechin has been demonstrated to reduce liver inflammatory injury by inhibiting the NF-κB activation (Citation73). Epicatechin can also activate this NF-κB pathway in HepG2 cells and this activation is related to ERK, PI3K/Akt and Nrf2 signaling as well (Citation328). One study reported that an epicatechin-rich extract had in-vitro and in-vivo anti-leukemic activity as it arrested the cell-cycle at the G0/G1 phase, activated caspase-3 and -8, increased levels of the anti-inflammatory cytokines IL-10 and IL-4, and suppressed of NF-κB activation (Citation159).

EGCG

(-)-Epigallocatechin-3-gallate or EGCG (14) is the predominant polyphenol in green tea. It has antioxidant potential and has been widely studied for its anticancer potential in-vitro and in-vivo (Citation75).

A study showed that EGCG exerts significant cytoprotection against H2O2 by upregulating HO-1, Nrf2 levels in nuclear extracts and ARE-luciferase activity via modulating the Akt and ERK1/2 pathways (Citation330). The Akt and MAPK pathways contribute to anti-apoptotic and growth stimulatory signaling (Citation331). EGCG also regulates ARE-driven antioxidant gene expression through induction of HO-1 and Nrf2 nuclear translocation mostly mediated by modulation of Akt, p38 and MAPK signaling in B lymphoblasts (Citation332). EGCG inhibits the MAPK and PI3K/Akt signaling pathways and modulates the expression of target genes which are associated with induction of apoptosis and cell cycle arrest in cancer cells (Citation333). Moreover, it upregulates the expression of Nrf2 and related antioxidant enzymes (GST, NQO1) as well as reduces DNA binding of NF-κB that inhibits the expression of inflammatory markers, such as monocyte chemotactic protein-1 (MCP-1) and vascular cell adhesion protein-1 (VCAM-1) producing PCB 126-induced inflammatory responses in endothelial cells (Citation334). Another report suggests the same notion regarding the anti-oxidative role of EGCG where they show that prevention of the activation of carcinogens by EGCG is mainly induced by the phase II detoxifying enzymes (Citation335) and thus contributes to cancer chemoprevention (Citation75). It also stimulates caspase-3 activity and induces apoptosis (Citation336). The activation of Nrf2 by EGCG increases the sensitivity of colorectal cancer cells toward radiation therapy, causing autophagy and inhibiting cell proliferation (Citation160). EGCG can increase the activity of chemotherapeutic agents by activating the AMPK pathway and inhibiting COX-2 expression (Citation337). The activation of Nrf2 by EGCG suppresses ROS production and has been reported to exert a cytoprotective effect on pancreatic cells (Citation338). EGCG also reduces ERK activation and activates p38 and JNK, and thus reduces the growth, invasion and angiogenesis in pancreatic cancer cells (Citation339). EGCG is able to suppress multi drug resistance (MDR). It has been reported to reduce Nrf2-mediated etoposide resistance in lung adenocarcinoma cells (Citation161). It reverses MDR by downregulating the Akt/mTOR pathway and sensitizing MDR cancer cells to chemotherapeutic agents (Citation340). The downregulation of PI3K/Akt/mTOR/HIF-1α by EGCG can also inhibit endometrial tumor angiogenesis (Citation162). EGCG has also been demonstrated to stop tumor growth, proliferation, migration, and angiogenesis in breast cancer cells via modulating the aforementioned pathways (Citation341). EGCG induces apoptosis in estrogen receptor negative MDA-MB-468 breast adenocarcinoma cells (Citation163). The downregulating effect of EGCG on the PI3K/ERK/NF-κB and PI3K/Akt pathways suppresses cancer cell invasion and induces apoptosis, respectively, while modulation of Nrf2 by EGCG causes antioxidant activity and induces apoptosis in breast cancer cells (Citation342). In-vivo studies have revealed that EGCG is most active against breast, colorectal, esophageal, gastric, lung, neural, oral and prostate tumors. These effects are induced via upregulation of Nrf2-UDP glucuronosyltransferase one family, polypeptide A (UGT1A) and VEGF/VEGFR, and downregulation of p-ERK1/2, COX-2, IL-6, and HIF-1α expression (Citation164).

Fisetin

Fisetin (15) is a flavonoid found in various vegetables and fruits, including strawberry, persimmon, grape, onion, apple and cucumber (Citation76). Fisetin has anticancer, antioxidant, and anti-inflammatory activity (Citation343).

The mechanism behind its antioxidative activity can be explained by its effect on Nrf2 activity. By dissociating Nrf2 from Keap1, fisetin enhances the accumulation of Nrf2 in the nucleus and increases the upregulation of ARE-regulated downstream genes coding for HO-1, GCLC, GCLM and NQO1 (Citation344). Fisetin has been demonstrated to increase GSH and SOD and reduce inflammatory cytokines release in-vivo via activating Nrf2/HO-1 and inhibiting the TLR4/NF-κB pathway, respectively (Citation345). Fisetin also regulates antioxidative mechanisms, such as SIRT1/Nrf2 signaling, and suppresses the activated p-JNK/NF-kB pathway to protect against oxidative stress (Citation346). The Nrf2 mediated upregulation of HO-1 provides cytoprotection from H2O2 induced cell injury in human umbilical vein endothelial cells (HUVECs) (Citation347). This Nrf2/HO-1 activation is also responsible for induction of apoptosis and attenuation of liver damage (Citation120). The cytoprotective effect of fisetin is mediated by phosphorylation of ERK, JNK and p38/MAPK pathways (Citation348). These pathways play a central role in cell proliferation, differentiation, transformation, migration and apoptosis (Citation349). The cytoprotection through ERK modulation helps in protecting from cholangiocarcinoma (Citation165). Fisetin induces apoptosis in oral squamous cell carcinoma by inhibition of autophagy (Citation166). This is also seen in MCF-7 breast cancer cells where fisetin induces apoptosis by inhibiting autophagy (Citation350). Fisetin inhibits autophagy by activating PI3K/Akt/mTOR and modulating the AMPK signaling pathway (Citation351). Another report suggests that it ameliorates mammary carcinoma both in-vitro and in-vivo via inhibiting PI3K/Akt/mTOR (Citation167). It can suppress the phosphorylation of MAPK/ERK/JNK and mRNA levels of pro-inflammatory factors (Citation352). It can eliminate damaged mitochondria in a p62-dependent manner, inhibiting the TLR4/MD2-mediated activation of the NLRP3 inflammasome (Citation353). Fisetin also inhibits inflammation and has a cytoprotective role in nephropathy by blocking of iRhom2/NF-κB signaling (Citation354). In hypoxia/re-oxygenation-treated RAW264.7 cells, fisetin exerts anti-inflammatory activity through modulation of the GSK-3β/AMPK signaling (Citation355). In LPS-stimulated human pulmonary artery endothelial cells, fisetin suppresses iNOS and TNF-α via downregulating p-STAT-1 and NF-κB pathway (Citation356). Through inhibiting this pathway and inhibiting the phosphorylation of ERK1/2 proteins, fisetin exerts anti-inflammatory activity (Citation357). Along with the inhibitory activity toward NF-κB, fisetin upregulates p53 to cause the induction of apoptosis in bladder cancer cells (Citation168). In colon cancer cells, fisetin also induces apoptosis upon modulation of the COX-2 and Wnt/EGFR/NF-κB signaling pathways where it inhibits COX-2 and EGF production (Citation169). Fisetin inhibits proliferation and migration of colorectal cancer cells and induces apoptosis, partially via interfering with signaling pathways related to the cell cycle regulators p21, p27, cyclin D1 and NF-κB p65. Suppression of tumor growth is also observed in mice inoculated with human HCT 116 colorectal cancer cells (Citation170). In prostate cancer cells, fisetin inhibits the PI3K/Akt pathway and induces apoptosis (Citation171). In-vitro and in-vivo evidence indicate that fisetin is able to suppress angiogenesis by mediating cell cycle arrest at G1 and G2/M phases, reducing cyclin D1 and survivin, enhancing the levels of p53, cleaved caspases-3 and -7 and PARP, and the ratio of Bax to Bcl-2 (Citation358). Fisetin downregulates eNOS, VEGF, iNOS, MMP-2 and -9 expression in A549 lung and DU145 prostate cancer cells. The effect of fisetin in pancreatic cancer via suppression of the PI3K/Akt/mTOR pathway was confirmed in-vivo (Citation172). Fisetin causes apoptosis in HeLa cells by ERK1/2-induced activation of caspase-8 and -3. This effect was also confirmed with decreased tumor growth in xenografted mice (Citation173). A recent study suggests evidence in favor of the anti-colorectal cancer effect of fisetin in-vivo and in-vitro via downregulation of cyclin D1 and NF-κB and induction of apoptosis (Citation359).

Genistein

Genistein (16) is a flavonoid found in soybean seeds and known for its estrogen-like biological activity (Citation77). Genistein has demonstrated activity in various cancer cells, including leukemia, lymphoma, prostate, breast, lung and head and neck cancer cells (Citation360).

Its anticancer effect, along with its antioxidative and anti-inflammatory activity can be attributed to its modulation of Nrf2 signaling. Genistein causes an increase in Keap1 S-nitrosylation with nuclear accumulation and DNA binding of Nrf2, resulting in an elevated level of antioxidant gene HO-1 expression. Genistein has demonstrated its cytoprotective activity against oxidative stress-induced epithelial cells through the activation of Nrf2 pathway and consequent upregulation of HO-1, SOD, CAT, GSH and NQO1 expressions (Citation361). A study reported that the activation of HO-1 and GCLC mRNA and protein expression is mediated via activation of ERK1/2 and PKC/Nrf2 signaling (Citation362). Genistein also activates Nrf2 via modulating PI3K activity, contributing to cytoprotective activity in cerebrovascular endothelial cells (Citation363). The modulation of PI3K/Akt and Nrf2/ARE activity by genistein contributes to the cancer chemopreventive activity of this molecule as these pathways regulate cell cycle progression, transformation, migration and apoptosis (Citation364). Genistein helps prevent prostate cancer by inhibiting the nuclear translocation of NF-κB, its binding to DNA, as well as blocking NF-κB activation by DNA-damaging agents (Citation174). Genistein, in combination with gemcitabine, displays anti-pancreatic cancer activity. This has been evidenced in-vitro and in-vivo via abolishing NF-κB and Akt expressions (Citation175). Genistein shows anti-inflammatory activity via inhibiting the production of pro-inflammatory mediators, including NOS2, COX-2 and MMPs that are produced following NF-κB activation (Citation365). It can modulate the genes associated with cell cycle and apoptosis by inactivating NF-κB and Akt pathways that affect the cell cycle and apoptosis (Citation360). By inducing apoptosis and inhibiting uncontrolled cell proliferation, genistein prevents the growth of MCF-7 breast cancer cells as well (Citation176). It activates the phase-II UDP-glucuronosyltransferases (UGTs) via SIRT1 activation (Citation366). Active SIRT1 has a biological effect on growth regulation and tumorigenesis and therefore, SIRT1 modulation provides anticancer activity (Citation367,Citation368). Genistein has anti-leukemic activity in-vitro and in-vivo, inducing ROS and Ca2+ generation (Citation177). It mediates apoptosis by G2/M phase cell cycle arrest, enhances the expressions of pro-apoptotic proteins, including Bax, PARP-cleavage, caspase-9, and -3, and reduces the expression of the anti-apoptotic protein Bcl-2. It sensitizes bladder cancer cells to hydroxycamptothecin in-vitro and in-vivo via apoptosis and the suppression of NF-κB (Citation178). Genistein has also been reported to enhance the activity of cisplatin on NSCLC in-vitro and in-vivo, via reduction of Akt and PI3K phosphorylation (Citation179).

Isoorientin

Isoorientin (17) is a flavonoid found in different plant species used for edible purposes, like Phyllostachys pubescens, Sasamorpha borealis, Eremurus spectabilis, and Fagopyrum esculentum (buckwheat) (Citation78,Citation79). Isoorientin has anti-nociceptive, anticancer and anti-inflammatory activity (Citation369).

It activates Nrf2 signaling and upregulates the expression of Nrf2-mediated antioxidative proteins (GCLC, GCLM, HO-1, NQO1 and Trx-1) and decreases the expression of Keap1 causing its dissociation from Nrf2 (Citation370). This activity helps to exert its antioxidative effect against cell injury. The antioxidative activity of isoorientin has been shown to ameliorate the cisplatin-induced side effects via activating the SIRT1/SIRT6/Nrf2 pathway. Apart from increasing the phase-II detoxifying enzymes, isoorientin also increases GSH levels in the liver and acts as a hepatoprotective agent (Citation371). The activation of phase-II enzymes, specially NQO1 by isoorientin fights against oxidative damage in liver carcinoma and provides cytoprotective activity that is dependent on the PI3K/Akt pathway (Citation79). Isoorientin induces apoptosis in HepG2 cells by modulating the PI3K/Akt pathway as well upon inhibition of Akt phosphorylation (Citation180). Apoptosis in HepG2 cells is also caused via modulation of the MAPK/ERK pathway (Citation372). It also exerts an inhibitory effect on p53 and the PI3K/Akt dependent NF-κB pathway (Citation78). Isoorientin inhibits the expression of inflammatory mediators, like COX-2, TNF-α, IL-6, 5-LO, and IL-1β via NF-κB inhibition (Citation373). Furthermore, isoorientin increases the expression of p-GSK-3β, thereby causing inhibition of GSK-3β and suppression of inflammation (Citation374). Isoorientin has also demonstrated antitumorigenic activity on pancreatic cancer cells by activating the AMPK pathway and decreasing the secretion of VEGF by AMPK (Citation181). In UVB-induced skin carcinogenesis, isoorientin provides protection by its autophagic action and suppression of JNK pathway activation (Citation182). Isoorientin induces apoptosis in lung cancer cells via modulating the MAPK/STAT3/NF-κB signaling pathways (Citation183), which are all linked with cellular transformation, tumorigenesis, cancer promotion and progression (Citation329). A recent study revealed that isoorientin is effective against oral squamous cell carcinoma in-vitro and in-vivo. The inhibition of EMT potential via suppression of JAK/STAT3 and Wnt/β-catenin signaling is involved behind this effect (Citation184).

Quercetin

Quercetin (18) is a flavonoid found in various fruits, vegetables, and in tea, red wine and medicinal herbs (Citation80). This compound has numerous biological properties including an ability to protect different organs (Citation375).

Quercetin increases Nrf2 mediated transcription and binding activity with ARE, stabilizes Nrf2 and enhances the mRNA and protein expression of Nrf2. It also reduces the level of Keap1 and enhances Nrf2 translocation in the nucleus, causing the activation of Nrf2 regulated antioxidant genes and phase-II detoxifying enzymes (Citation376). The activation of Nrf2 by quercetin plays a vital role for various disease prevention. In liver carcinoma, quercetin provides protection from hepatotoxicity upon activation of Keap1-Nrf2 signaling, causing the dissociation and translocation of Nrf2 as well as induction of the JNK pathway (Citation80). It modulates Nrf2 and induces p38-MAPK signaling and cell death in HepG2 cells by increasing glutathione related enzymes, such as GSH, glutamyl cysteine-synthetase (GCS), Gpx, and GR (Citation377). Quercetin restores SOD and MDA levels via upregulating Nrf2 (Citation378). It activates the phosphorylation of JNK, p38 and PI3K/Akt as well as enhances Nrf2 DNA binding activity (Citation379). In LPS-induced oxidative stress, quercetin attenuates the LPS-mediated inhibition of JNK, ERK and p38 phosphorylation in the MAPK/Nrf2 signaling pathway (Citation380). It also suppresses NF-κB nuclear translocation and expression, causing the downregulation of COX-2 and thus exerting anti-inflammatory activity via Nrf2 activation (Citation381). Quercetin protects from inflammatory liver damage by reducing PI3K/Nrf2-mediated oxidative stress, activating mTOR in autophagy, inhibiting the expression of apoptotic factors and suppressing the NF-κB/TLR/NLRP3 pathway (Citation382). It modulates Nrf2/HO-1 and p38/STAT1/NF-κB signaling pathway by upregulating Nrf2 and inducing HO-1 activity, inhibiting p38 and STAT1 activation and inactivating NF-κB (Citation383). This NF-κB inactivation not only provides quercetin with anti-inflammatory activity but also with a cancer protective role as NF-κB signaling affects cell survival and proliferation, and is linked with carcinogenesis and the response of cancer cells to therapy (Citation290). It shows antigenotoxic effect and prevents DNA damage in human hepatoma cells by blocking the NF-κB pathway as well (Citation185). This effect of attenuating DNA damage also protects colon cancer cells from 1,2-dimethylhydrazine-induced colon cancer (Citation186). Along with vitamin C, quercetin exhibits cytotoxicity against MDA-MB 231 breast cancer cells via reducing the overexpression of Nrf2 and balancing ROS levels (Citation187). This cytotoxic effect of quercetin and vitamin C has also been observed in DU145 and PC3 prostate cancer cells upon decrease in Nrf2 gene expression (Citation188). It has been noted that in some cases it is important to decrease Nrf2 levels because overexpression of Nrf2 causes a marked increase in chemoresistance (Citation384). In combination with kaempferol and pterostilbene, it exerts a synergistic effect on ROS scavenging by activating Nrf2/ARE signaling and increasing the expression of mRNA and protein Nrf2 levels (Citation385). Quercetin also induces apoptosis by modulating the Nrf2 pathway and providing cytoprotection in malignant mesothelioma MSTO-211H and H2452 cells (Citation189). Quercetin has anti-prostate cancer activity in-vitro and in-vivo. It significantly increases antioxidant enzymes (SOD, CAT, Gpx, GR, and GST), reduces the expression of the anti-apoptotic protein Bcl-2, and enhances the expression of caspase-8. It is able to suppress Akt and ERK levels in prostate cancer-induced rats (Citation190). Studies on different animal models have reported that its cancer protective mechanism is mediated via downregulation of PI3K/Akt, Akt/mTOR and upregulation of JNK/ERK MAPK signaling (Citation386). Another study showed that quercetin could, in addition to its apoptotic effect, enhance the levels of SOD, GSH, and reduce the level of MDA in lung cancer cells (Citation191).

Luteolin

Luteolin (19) is a flavonoid present in high concentrations in celery, green pepper, parsley, perilla leaf, chamomile tea and various other fruits, vegetables and medicinal herbs (Citation81). Luteolin has anti-inflammatory, antioxidative, anti-allergic and anticancer activity (Citation387).

It exerts its biological effects by modulating various pathways, such as the Nrf2/ARE, PI3K/Akt and NF-κB signaling (Citation387). Luteolin shows cytoprotective effect by increasing the binding of Nrf2 with ARE followed by an upregulation of its downstream HO-1 mRNA and protein expression via activating the ERK1/2 signaling (Citation388). This cytoprotective effect of luteolin enables it to protect hepatocytes from tBHP-induced oxidative injury by upregulating GCLC and GCLM, intracellular GSH and HO-1 expression via increasing Nrf2 activity (Citation314). Luteolin has shown synergistic activity with metformin on carbon tetrachloride-induced hepatotoxicity by activating Nrf2/ARE signaling and decreasing the release of inflammatory cytokines IL-1β, TNF-α, and IL-6 (Citation389). Luteolin protects against pyroptosis-linked inflammation by suppressing ROS production via Nrf2 activation and inactivation of the NF-κB signaling that is associated with carcinogenesis via the production of inflammatory cytokines (Citation390). Luteolin also shows activity against oxidative cell injury via the activation of Nrf2/ARE that causes the upregulation of Nrf2 downstream genes HO-1, NQO1, SOD, GPx, CAT and enhancement of the eNOS mediated S-nitrosylation of Keap1 (Citation391). Luteolin induces apoptosis and exerts its anticancer effect on colon cancer cells by upregulating Nrf2 via the suppression of DNA methylation, followed by initiation of the interaction between Nrf2 and p53 (Citation192). Luteolin attenuates aflatoxin B1-induced apoptosis in mice by decreasing Bax, Cyt-c, caspase-3 and caspase-9 transcription and upregulating Nrf2 and its downstream protein expressions (HO-1, NQO1, GCLC, SOD1) (Citation392). Another study indicated that luteolin attenuates the proliferation and transformation of HCT116 and HT29 cells upon demethylation of the Nrf2 promoter region, thereby upregulating Nrf2 and its downstream antioxidative products (Citation193). Luteolin has also been reported to induce apoptosis in HT29 cells through Nrf2 modulation (Citation393). Apart from its anticancer potential, luteolin also prevents chemoresistance. It is known that overexpression of Nrf2 causes cancer cell promotion and growth that confers therapeutic resistance toward anticancer drugs (Citation394). Luteolin can inhibit Nrf2 when it overexpresses in-vivo and downregulate the Nrf2-regulated NQO1 gene expression (Citation395). For instance, A549 human NSCLC cells show resistance toward anticancer drugs as these cells possess constitutively active Nrf2. Luteolin is able to sensitize these cells toward therapeutic drugs by repressing Nrf2 activation (Citation81). Luteolin inhibits the expression of Nrf2, HO-1 and Cripto-1 proteins which cause breast cancer stemness and thereby enhances chemosensitivity (Citation396). It also inhibits the chemoresistance in human colorectal cells toward oxaliplatin by inhibiting the overexpressed Nrf2 and downregulating NQO1, HO-1 and GSTα1/2 expressions (Citation397). It prevents chemoresistance in breast cancer cells as well, by significantly increasing cancer cell death. Here, luteolin downregulates the expression of HO-1 and MDR1 via blocking the Nrf2 activation in MDA-MB 231 breast cancer cells resistant to doxorubicin (Citation194). Luteolin has activity against human bladder cancer in-vitro and in-vivo by enhancing TRX1 and reducing ROS levels. The inhibition of mTOR signaling is the major pathway by which luteolin exerts this effect (Citation195). Luteolin has an effect on ER‑negative breast tumors and melanoma in-vitro and in-vivo via induction of apoptosis, as evidenced by a marked reduction in MMP‑2 and ‑9 expression, and inhibition of PI3K and Akt phosphorylation (Citation196,Citation197). In a mouse model of pancreatic cancer, luteolin and gemcitabine synergistically induce apoptosis via abrogating K-ras/GSK-3β/NF-κB signaling, reducing the Bcl-2/Bax ratio, releasing cytochrome C, and activating caspase 3 (Citation198). Luteolin, in combination with paclitaxel, acts synergistically against esophageal cancer in-vitro and in-vivo by abrogating cell migration and EMT that is associated with suppression of SIRT1 expression, and the ROS/JNK-induced activation of the mitochondrial apoptotic pathway (Citation199). Another study shows that luteolin alone is active against esophageal cancer by inhibiting the PI3K/Akt pathway (Citation200).

Rutin

Rutin (20), also known as rutoside or vitamin P is present in various plants, including asparagus, buckwheat, cherries, plums, oranges, grapes, grapefruit, apricots, apples and tea (Citation82). It possesses antioxidant, anti-inflammatory, antiangiogenic, pro-apoptotic, and antiproliferative effects, all of which may participate in the prevention and treatment of cancer (Citation398).

Studies demonstrated that rutin upregulates the expression of Nrf2 by enhancing the activity of downstream HO-1, NQO-1, GST and Mn-SOD phase-II detoxifying enzymes which repair the oxidative imbalance in cells (Citation399). The activation of Nrf2 in association with the degradation of Keap1 via modulating phosphorylation of PI3K/Akt is a major antioxidative mechanism of rutin (Citation399). This demonstrates the anticancer potential of rutin as ROS imbalance is a key factor affecting apoptosis and autophagy (Citation400). Rutin has been reported to hinder tumor growth in colon cancer cells via inhibition of NF-κB signaling and modulation of MAPK and MAPK activated protein kinase 2 (Citation201). This role of rutin in inactivating NF-κB via p38 along with cell cycle arrest also helps prevent lung (Citation202) and gastric cancer (Citation203). Lung cancer can also be blocked by rutin through modulating TNF-α and GSK-3β expressions, which play a vital role in cell cycle, cell proliferation and apoptosis (Citation204). In case of prostate cancer, rutin upregulates p53 expression synergistically with 5-FU (Citation205). It also provides protection against breast cancer by modulating Akt/mTOR signaling as well as inducing cell cycle arrest at G2/M phase via p53 signaling (Citation206). Rutin has been reported to protect liver cells against inflammation via downregulating CCl4-induced activation of NF-κB, TNF-α and COX-2 (Citation401). It suppresses JNK-mediated autophagy in brain cancer cells as well (Citation207). On the other hand, in leukemia THP-1 cells, rutin enhances autophagy and suppresses inflammation by inactivating NF-κB and reducing TNF-α levels (Citation208). Autophagy is an important process that acts both as tumor generator and tumor inhibitor (Citation402). Rutin has also been demonstrated to protect against oxidative stress and inflammation following bisphenol and dibutyl phthalate exposure through upregulation of Nrf2, SOD, GSH and inhibition of NF-κB activation (Citation403). Rutin, along with ascorbic acid, abrogates UVA-and UVB-induced damage in skin keratinocytes, providing cytoprotective activity via the activation of Nrf2 signaling (Citation404). The in-vivo activity of rutin against various cancers has been associated with the suppression of STAT3/NF-κB, Bcl-2, AP-1, p38 MAPK and the activation of the Wnt/β-catenin pathway. Its anticancer effect has also been linked to its inhibition of COX-2, iNOS, TNF-α, and ROS (Citation209,Citation210).

Kaempferol

Kaempferol (21) is a flavonoid found in many plant-derived foods like Aloe vera, Coccinia grandis, Moringa oleifera, broccoli, tea, beans and strawberries (Citation83). Kaempferol exerts diverse biological effects, such as antioxidative, anti-inflammatory, and anticancer with potential uses in diseases, such as diabetes, allergy, osteoporosis, cardiovascular, neurodegenerative and infectious diseases (Citation405).

It increases Nrf2 protein expression leading to the upregulation of its downstream HO-1 gene and an increase in SOD and GSH levels (Citation406). This activation of Nrf2/HO-1 signaling reduces ROS levels and is regarded as the major mechanism behind antioxidative activity of kaempferol (Citation385). Kaempferol also elevates nuclear levels of HO-1 and Nrf2 through attenuation of the cisplatin-mediated phosphorylation of p38, ERK1/2 and JNK (Citation407). In case of pancreatic cancer, kaempferol promotes apoptosis by elevating ROS generation and decreasing transglutaminase-2 (TGM2) mRNA and protein levels (Citation212). Kaempferol has also been reported to exert cytoprotective activity in liver and lung cells via upregulating Nrf2 and increasing CAT, SOD and p38 levels (Citation408). In HepG2 cells, t-BHQ induced phase-II enzymes are dependent on Nrf2 stability and kaempferol can influence this stability (Citation409). Besides activating the Nrf2 pathway, kaempferol also suppresses the activation of NF-κB and reduces the levels of TNF-α and IL-6 which helps to exert its protective effect against cellular damage (Citation410). The reduction in intracellular ROS and inflammatory cytokines can help prevent tumorigenesis and carcinogenesis. It induces apoptosis in human MCF-7, SGC-7901, Hela and A549 cells (Citation378). Kaempferol is a potent inducer of Nrf2 and its downstream NQO1 gene and plays a pivotal role in preventing carcinogenesis in breast cancer cells (Citation411). It increases the activity of phosphatase and tensin homolog (PTEN) and AMPK while decreasing the activation of Akt/mTOR signaling (Citation412). Akt/mTOR is a key signaling pathway involved in tumorigenesis (Citation413) and an imbalance in this pathway is responsible for apoptosis resistance (Citation414). Studies have reported that kaempferol had anti-cervical and colorectal cancer activity via inhibition of PI3K/Akt signaling (Citation213,Citation214). Furthermore, kaempferol inhibits the proliferation of human hepatocellular carcinoma cells via inducing autophagy through the activation of AMPK signaling (Citation211). Kaempferol induces G1 and G2/M cell cycle arrest by inhibiting the activity of cyclin dependent kinase 2 (CDK2), CDK4, and Cdc2 in HT-29 human colon cancer cells (Citation415). To combat oxidative injury, kaempferol activates and increases the accumulation of the SIRT1 linked with cellular growth regulation and tumorigenesis (Citation367), as well as inhibits PARP1 that leads to increased Nrf2 expression (Citation416). Kaempferol also shows anticancer activity by inhibiting GSK-3β/Nrf2 signaling as GSK-3β plays a vital role in regulating cell cycle, cell proliferation and apoptosis (Citation405). In ovarian cancer, kaempferol upregulates p53 expression and induces cell cycle arrest at G2/M phase (Citation417). It also prevents gastric cancer tumor growth in-vivo and in-vitro via modulating Akt, ERK and COX-2 expressions (Citation215). By modulating these pathways that control cell cycle, growth, and apoptosis, kaempferol affects tumorigenesis. It has been reported to alleviate endothelial cell injury and oxidative stress and induce apoptosis in HUVEC cells by activating AMPK/Nrf2/HO-1 signaling (Citation418). It can also modulate the NF-κB/MAPK and AMPK/Nrf2 pathways to decrease inflammation and oxidative stress (Citation419). Besides inducing activation of Nrf2, kaempferol can also downregulate the excess Nrf2 activation through reduction of Nrf2 mRNA and protein levels along with Nrf2 target genes (NQO1, HO-1, AKR1C1 and GST). In this way, kaempferol increases ROS accumulation and makes NSCLC cells sensitive to apoptosis which causes abrogation of chemoresistance (Citation216). Kaempferol exerts anti-angiogenic effect in-vitro in ovarian cancer via abolishing VEGF secretion, suppressing ERK phosphorylation and NF-κB and cMyc expression, but facilitating p21 expression (Citation217). In NSCLC, kaempferol in combination with radiotherapy was reported to show promising antitumor activity in-vitro and in-vivo via inhibiting the Akt/PI3K and ERK pathways and activating mitochondrial apoptosis (Citation218). In MCF-7 breast cancer cell xenografted mice, kaempferol inhibits the phosphorylation of insulin receptor substrate 1 (IRS-1), Akt, MAPK/extracellular signal-regulated kinase 1/2, and ERK signaling proteins (Citation219). Additional In Vitro and In Vivo studies indicated that kaempferol had anti-breast cancer activity via inhibiting the PKCδ/MAPK/AP-1 pathway and downregulating MMP-9 expression (Citation220).

Strengths/Limitations of Previous Studies and Suggestions for Future Work

The results of the above studies indicate that the 21 selected dietary polyphenols have a promising cancer protective potential via modulation of Keap1/Nrf2/ARE and other interconnected signaling pathways. Studies, carried out using in-vitro and/or in-vivo models, showed that these compounds exerted their effects (antiproliferative, antitumorigenic, pro-apoptotic, anti-inflammatory, and antioxidative) in a variety of different cancers. A limited number of in-vivo studies were performed to confirm the in-vitro findings. Only one clinical trial was conducted to evaluate the effectiveness of resveratrol on patients with prostate, colorectal and breast cancer.

Further studies are required to confirm the cancer protective role of the selected dietary polyphenols. In particular, the cancer protective role of polyphenols, such as methyleugenol, carnosol, and catechin, has yet to be studied in-vivo. In addition, further work should be focused on designing rigorous controlled clinical trials to establish whether the consumption of such polyphenols has an impact on the incidence and progression of cancers in humans. These compounds may prove to be particularly useful as alternative options for patients with pre-neoplastic lesions, early-stage cancers, as well as in end stage disease where there is enhanced drug resistance.

Authors’ Contributions

M.A.I. and M.M.M. were responsible for literature analysis and curation, drafting the final manuscript and drawing the figures. A.U.N. and M.A.A.F were responsible for literature curation, initial drafting of the manuscript and drawing the figures. M.A.S. was responsible for conceptualization, literature searching, revision and supervision. V.S. critically revised the manuscript and was responsible for drafting the final manuscript. All authors have read and approved the submitted version of the manuscript.

Abbreviations
ROS=

Reactive oxygen species;

Keap1=

Kelch-like ECH associated protein 1;

Nrf2=

Nuclear factor erythroid 2-related factor 2;

ARE=

Antioxidant response elements;

GSH=

Glutathione;

MAPK=

Mitogen activated-protein kinase;

ERK1/2=

Extracellular-regulated kinase 1/2;

PI3K=

Phosphatidylinositol 3-kinase;

Akt=

Protein kinase B;

PKD=

Protein kinase D;

JNKs=

c-Jun N-terminal kinases;

NF-κB=

Nuclear factor kappa B;

PKC=

Protein kinase C;

AP-1=

Activator protein-1;

HIF 1α=

hypoxia-inducible factor 1-alpha;

MDA=

Malondialdehyde;

VEGF=

Vascular endothelial growth factor;

PAK-1=

p21-activated kinase-1;

NOX=

NADPH oxidases;

NFE2=

Nuclear factor erythroid-derived 2;

Neh=

NRF2-ECH homology;

sMAF=

Small musculo-aponeurotic fibrosarcoma;

BTB=

Broad complex, tramtrack, bric-a-brac;

IVR=

Intervening region;

Maf=

Muscle aponeurosis fibromatosis;

GST=

Glutathione S-transferase;

NQO1=

NADPH quinine oxidoreductase;

GCL=

Glutamylcysteine ligase;

GPx=

Glutathione peroxidases;

SOD=

Superoxide dismutase;

UGT=

UDP-glucuronosyl transferase;

HO-1=

Heme oxygenase-1;

Trx=

thioredoxin;

EH=

Epoxide hydrolases;

IL-1β=

Interleukin-1β;

PPARγ=

peroxisome proliferator-activated receptor gamma;

AMPK=

AMP-activated protein kinase;

ASK=

Apoptosis signal-regulating kinase 1;

BCL-2=

B-cell lymphoma 2;

HIPK2=

Homeodomain-interacting protein kinase 2;

PDGF=

Platelet derived growth factor;

EMT=

Epithelial to mesenchymal transition;

MMP-2=

Matrix metalloproteinase-2;

STAT3=

Signal transducer and activator of transcription 3;

LPS=

Lipopolysaccharide;

GSK3β=

Glycogen synthase kinase-3 beta;

NLRP3=

NLR family pyrin domain containing 3;

TNF-α=

Tumor necrosis factor-α;

mTOR=

Mammalian target of rapamycin;

TLR4=

Toll-like receptor 4;

NO=

Nitric oxide;

PGE2=

Prostaglandin E2;

CAT=

Catalase;

OGG1=

8-oxoguanine glycosylase;

TCDD=

2,3,7,8-Tetrachlorodibenzo-p-dioxin;

SIRT1=

Sirtuin 1;

FOXO1=

Forkhead box protein O 1;

DENA=

Diethylnitrosamine;

TGF-β1=

Tumor growth factor-β1;

IKKβ=

Inhibitor of nuclear factor kappa-B kinase subunit beta;

GCLC=

Glutamate cysteine ligase catalytic;

GCLM=

Glutamate cysteine ligase modifier;

LTB4=

Leukotriene B4;

Syk=

Spleen tyrosine kinase;

PLA2=

Phospholipase A2;

5-LO=

5-lipoxygenase;

AKR1B10=

Aldo-keto reductase family 1 member B10;

FTL=

Ferritin Light Chain;

GGTLA4=

Gamma-glutamyltransferase-like activity 4;

GCS=

γ-glutamylcysteine synthetase;

FKHR=

Forkhead transcription factors;

NSCLC=

Non-small cell lung cancer cell;

CCL5=

Chemokine (C-C motif) ligand 5;

iNOS=

Inducible nitric oxide synthase;

COX-2=

Cyclooxygenase-2;

Wnt=

Wingless-related integration site;

5-FU=

5-Fluorouracil;

TPA=

12-O-tetradecanoylphorbol-13-acetate;

EGF=

Epidermal growth factor;

PARP=

Poly (ADP-ribose) polymerase;

cGMP=

Cyclic guanosine monophosphate;

MT=

Metallothionein;

IGF II=

Insulin-like growth factor II;

JAK2=

Janus kinase 2;

TRAIL=

TNF-related apoptosis-inducing ligand;

PERK=

Protein kinase RNA-like endoplasmic reticulum kinase;

DNMT=

DNA methyltransferase;

HDAC=

Histone deacetylases;

NLCs=

Nanostructured lipid carriers;

MRP2=

Multidrug resistance-associated protein 2;

MCP-1=

Monocyte chemotactic protein-1;

VCAM-1=

Vascular cell adhesion protein-1;

MDR=

Multi drug resistance;

UGT1A=

UDP glucuronosyltransferase 1 family, polypeptide A;

HUVECs=

Human umbilical vein endothelial cells;

UGTs=

UDP-glucuronosyltransferases;

GCS=

Glutamyl cysteine-synthetase;

GR=

Glutathione-reductase;

TGM2=

Transglutaminase-2;

PTEN=

Phosphatase and tensin homolog;

CDK2=

Cyclin dependent kinase 2: IRS-1: insulin receptor substrate 1

Disclosure Statement

The authors declare no conflicts of interest.

Funding

This review received no specific grant from any funding agency, commercial, or not-for-profit sectors.

References

  • Chio IIC, Tuveson DA. ROS in cancer: the burning question. Trends Mol Med. 2017;23(5):411–29. doi:10.1016/j.molmed.2017.03.004.
  • Chun K-S, Raut PK, Kim D-H, Surh Y-J. Role of chemopreventive phytochemicals in NRF2-mediated redox homeostasis in humans. Free Radic Biol Med. 2021;172:699–715. doi:10.1016/j.freeradbiomed.2021.06.031.
  • Mandelker L. Introduction to oxidative stress and mitochondrial dysfunction. Vet Clin North Am Small Anim Pract. 2008;38(1):1–30, v. doi:10.1016/j.cvsm.2007.10.005.
  • Moloney JN, Cotter TG, ROS signalling in the biology of cancer. Semin Cell Dev Biol. 2018;80:50–64. 10.1016/j.semcdb.2017.05.023
  • Klaunig JE, Wang Z. Oxidative stress in carcinogenesis. Curr Opin Toxicol. 2018;7:116–21. doi:10.1016/j.cotox.2017.11.014.
  • Klaunig JE, Kamendulis LM, Hocevar BA. Oxidative stress and oxidative damage in carcinogenesis. Toxicol Pathol. 2010;38(1):96–109. doi:10.1177/0192623309356453.
  • Noda N, Wakasugi H. Cancer and oxidative stress. Japan Med Assoc J. 2001;44(12):535–9.
  • Sosa V, Moliné T, Somoza R, Paciucci R, Kondoh H, LLeonart ME. Oxidative stress and cancer: an overview. Ageing Res Rev. 2013;12(1):376–90. doi:10.1016/j.arr.2012.10.004.
  • Shaw P, Chattopadhyay A. Nrf2–ARE signaling in cellular protection: mechanism of action and the regulatory mechanisms. J Cell Physiol. 2020;235(4):3119–30. doi:10.1002/jcp.29219.
  • Liu X-F, Hao J-L, Xie T, Malik TH, Lu C-B, Liu C, Shu C, Lu C-W, Zhou D-D. Nrf2 as a target for prevention of age‐related and diabetic cataracts by against oxidative stress. Aging Cell. 2017;16(5):934–42. doi:10.1111/acel.12645.
  • Kobayashi M, Yamamoto M. Nrf2-Keap1 regulation of cellular defense mechanisms against electrophiles and reactive oxygen species. Adv Enzyme Regul. 2006;46(1):113–40. 10.1016/j.advenzreg.2006.01.007
  • Kim HM, Han JW, Chan JY. Nuclear factor erythroid-2 like 1 (NFE2L1): structure, function and regulation. Gene. 2016;584(1):17–25. doi:10.1016/j.gene.2016.03.002.
  • Dodson M, De La Vega MR, Cholanians AB, Schmidlin CJ, Chapman E, Zhang DD. Modulating NRF2 in disease: timing is everything. Annu Rev Pharmacol Toxicol. 2019;59:555–75. doi:10.1146/annurev-pharmtox-010818-021856.
  • He F, Ru X, Wen T. NRF2, a transcription factor for stress response and beyond. IJMS. 2020;21(13):4777. doi:10.3390/ijms21134777.
  • Ojo OA, Ajiboye B, Fadaka A, Taro P, Shariati MA. Nrf2-Keap1 activation, a promising strategy in the prevention of cancer. FRA. 2017;7(1):1–7. doi:10.5530/fra.2017.1.1.
  • Tebay LE, Robertson H, Durant ST, Vitale SR, Penning TM, Dinkova-Kostova AT, Hayes JD. Mechanisms of activation of the transcription factor Nrf2 by redox stressors, nutrient cues, and energy status and the pathways through which it attenuates degenerative disease. Free Radic Biol Med. 2015;88(Pt B):108–46. 10.1016/j.freeradbiomed.2015.06.021
  • Zhou S, Ye W, Zhang M, Liang J. The effects of nrf2 on tumor angiogenesis: a review of the possible mechanisms of action. Crit Rev Eukaryot Gene Expr. 2012;22(2):149–60. doi:10.1615/critreveukargeneexpr.v22.i2.60.
  • Lin T-Y, Cantley LC, DeNicola GM. NRF2 rewires cellular metabolism to support the antioxidant response. In: Morales-Gonzalez JA, Morales-Gonzalez A and Madrigal-Santillan EO, editors. A master regulator of oxidative stress - the transcription factor Nrf2. London: InTechOpen; 2016. p. 107–8.
  • Stewart JD, Hengstler JG, Bolt HM. Control of oxidative stress by the Keap1-Nrf2 pathway. Arch Toxicol. 2011;85(4):239. doi:10.1007/s00204-011-0694-1.
  • Giudice A, Montella M. Activation of the Nrf2–ARE signaling pathway: a promising strategy in cancer prevention. Bioessays. 2006;28(2):169–81. doi:10.1002/bies.20359.
  • Qin S, Deng F, Wu W, Jiang L, Yamashiro T, Yano S, Hou D-X. Baicalein modulates Nrf2/Keap1 system in both Keap1-dependent and Keap1-independent mechanisms. Arch Biochem Biophys. 2014;559:53–61. doi:10.1016/j.abb.2014.03.011.
  • Xiang M, Namani A, Wu S, Wang X. Nrf2: bane or blessing in cancer? J Cancer Res Clin Oncol. 2014;140(8):1251–9. doi:10.1007/s00432-014-1627-1.
  • Robledinos-Antón N, Fernández-Ginés R, Manda G, Cuadrado A. Activators and inhibitors of NRF2: a review of their potential for clinical development. Oxid Med Cell Longev. 2019;2019:1–20. doi:10.1155/2019/9372182.
  • Schmidlin CJ, Dodson MB, Madhavan L, Zhang DD. Redox regulation by NRF2 in aging and disease. Free Radic Biol Med. 2019;134:702–7. doi:10.1016/j.freeradbiomed.2019.01.016.
  • Kensler TW, Wakabayashi N. Nrf2: friend or foe for chemoprevention? Carcinogenesis. 2010;31(1):90–9. doi:10.1093/carcin/bgp231.
  • Smith RE, Tran K, Smith CC, McDonald M, Shejwalkar P, Hara K. The role of the Nrf2/ARE antioxidant system in preventing cardiovascular diseases. Diseases. 2016;4(4):34. doi:10.3390/diseases4040034.
  • McWalter GK, Higgins LG, McLellan LI, Henderson CJ, Song L, Thornalley PJ, Itoh K, Yamamoto M, Hayes JD. Transcription factor Nrf2 is essential for induction of NAD (P) H: quinone oxidoreductase 1, glutathione S-transferases, and glutamate cysteine ligase by broccoli seeds and isothiocyanates. J Nutr. 2004;134(12 Suppl):3499S–506S. doi:10.1093/jn/134.12.3499S.
  • Marampon F, Codenotti S, Megiorni F, Del Fattore A, Camero S, Gravina GL, Festuccia C, Musio D, De Felice F, Nardone V, et al. NRF2 orchestrates the redox regulation induced by radiation therapy, sustaining embryonal and alveolar rhabdomyosarcoma cells radioresistance. J Cancer Res Clin Oncol. 2019;145(4):881–93. doi:10.1007/s00432-019-02851-0.
  • Su Z-Y, Shu L, Khor TO, Lee JH, Fuentes F, Kong A-NT. A perspective on dietary phytochemicals and cancer chemoprevention: oxidative stress, nrf2, and epigenomics. Top Curr Chem. 2012;329:133–62. doi:10.1007/128_2012_340.
  • Ma Q. Role of nrf2 in oxidative stress and toxicity. Annu Rev Pharmacol Toxicol. 2013;53:401–26. doi:10.1146/annurev-pharmtox-011112-140320.
  • de Freitas Silva M, Pruccoli L, Morroni F, Sita G, Seghetti F, Viegas C, Tarozzi A. The Keap1/Nrf2-ARE pathway as a pharmacological target for chalcones. Molecules. 2018;23(7):1803. 10.3390/molecules23071803
  • Roque AT, Gambeloni RZ, Felitti S, Ribeiro ML, Santos JC. Inflammation-induced oxidative stress in breast cancer patients. Med Oncol. 2015;32(12):1–4. doi:10.1007/s12032-015-0709-5.
  • Hayes JD, Dinkova-Kostova AT, Tew KD. Oxidative stress in cancer. Cancer Cell. 2020;38(2):167–97. doi:10.1016/j.ccell.2020.06.001.
  • Siraj MA, Islam MA, Al Fahad MA, Kheya HR, Xiao J, Simal-Gandara J. Cancer chemopreventive role of dietary terpenoids by modulating keap1-Nrf2-ARE signaling system—A comprehensive update. Appl Sci. 2021;11(22):10806. 10.3390/app112210806
  • Wakabayashi N, Shin S, Slocum SL, Agoston ES, Wakabayashi J, Kwak M-K, Misra V, Biswal S, Yamamoto M, Kensler TW. Regulation of notch1 signaling by nrf2: implications for tissue regeneration. Sci Signal. 2010;3(130):ra52–ra. 10.1126/scisignal.2000762
  • Malhotra D, Portales-Casamar E, Singh A, Srivastava S, Arenillas D, Happel C, Shyr C, Wakabayashi N, Kensler TW, Wasserman WW, et al. Global mapping of binding sites for Nrf2 identifies novel targets in cell survival response through ChIP-Seq profiling and network analysis. Nucleic Acids Res. 2010;38(17):5718–34. 10.1093/nar/gkq212
  • Niture SK, Jaiswal AK. Nrf2-induced antiapoptotic Bcl-xL protein enhances cell survival and drug resistance. Free Radic Biol Med. 2013;57:119–31. 10.1016/j.freeradbiomed.2012.12.014
  • Li J, Lee J-M, Johnson JA. Microarray analysis reveals an antioxidant responsive element-driven gene set involved in conferring protection from an oxidative stress-induced apoptosis in IMR-32 cells. J Biol Chem. 2002;277(1):388–94. 10.1074/jbc.M109380200
  • Niso-Santano M, González-Polo RA, Bravo-San Pedro JM, Gómez-Sánchez R, Lastres-Becker I, Ortiz-Ortiz MA, Soler G, Morán JM, Cuadrado A, Fuentes JM, Centro de Investigación Biomédica en Red sobre Enfermedades Neurodegenerativas (CIBERNED), et al. Activation of apoptosis signal-regulating kinase 1 is a key factor in paraquat-induced cell death: modulation by the Nrf2/Trx axis. Free Radic Biol Med. 2010;48(10):1370–81. 10.1016/j.freeradbiomed.2010.02.024
  • Oh E-T, Kim J-w, Kim JM, Kim SJ, Lee J-S, Hong S-S, Goodwin J, Ruthenborg RJ, Jung MG, Lee H-J, et al. NQO1 inhibits proteasome-mediated degradation of HIF-1α. Nat Commun. 2016;7(1):1–14. 10.1038/ncomms13593
  • Arfmann-Knübel S, Struck B, Genrich G, Helm O, Sipos B, Sebens S, Schäfer H. The crosstalk between Nrf2 and TGF-β1 in the epithelial-mesenchymal transition of pancreatic duct epithelial cells. PLoS One. 2015;10(7):e0132978. 10.1371/journal.pone.0132978
  • Zhao Q, Mao A, Guo R, Zhang L, Yan J, Sun C, Tang J, Ye Y, Zhang Y, Zhang H, et al. Suppression of radiation-induced migration of non-small cell lung cancer through inhibition of Nrf2-Notch Axis. Oncotarget. 2017;8(22):36603–13. 10.18632/oncotarget.16622
  • Zhou W, Mo X, Cui W, Zhang Z, Li D, Li L, Xu L, Yao H, Gao J. Nrf2 inhibits epithelial-mesenchymal transition by suppressing snail expression during pulmonary fibrosis. Sci Rep. 2016;6(1):1–10. 10.1038/srep38646
  • Hayes JD, Dinkova-Kostova AT. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem Sci. 2014;39(4):199–218. 10.1016/j.tibs.2014.02.002
  • de la Vega MR, Chapman E, Zhang DD. NRF2 and the hallmarks of cancer. Cancer Cell. 2018;34(1):21–43. 10.1016/j.ccell.2018.03.022
  • Song M-Y, Lee D-Y, Chun K-S, Kim E-H. The role of NRF2/KEAP1 signaling pathway in cancer metabolism. IJMS. 2021;22(9):4376. doi:10.3390/ijms22094376.
  • Wang Y-Y, Chen J, Liu X-M, Zhao R, Zhe H. Nrf2-mediated metabolic reprogramming in cancer. Oxid Med Cell Longev. 2018;2018:1–7. doi:10.1155/2018/9304091.
  • Giudice A, Arra C, Turco MC. Review of molecular mechanisms involved in the activation of the Nrf2-ARE signaling pathway by chemopreventive agents. Methods Mol Biol. 2010;647:37–74. doi:10.1007/978-1-60761-738-9_3.
  • Kaspar JW, Niture SK, Jaiswal AK. Nrf2: INrf2 (Keap1) signaling in oxidative stress. Free Radic Biol Med. 2009;47(9):1304–9. doi:10.1016/j.freeradbiomed.2009.07.035.
  • Kensler TW, Curphey TJ, Maxiutenko Y, Roebuck BD. Chemoprotection by organosulfur inducers of phase 2 enzymes: dithiolethiones and dithiins. Drug Metabol Drug Interact. 2000;17(1–4):3–22. doi:10.1515/dmdi.2000.17.1-4.3.
  • Chen C, Pung D, Leong V, Hebbar V, Shen G, Nair S, Li W, Kong A-NT. Induction of detoxifying enzymes by garlic organosulfur compounds through transcription factor Nrf2: effect of chemical structure and stress signals. Free Radic Biol Med. 2004;37(10):1578–90. doi:10.1016/j.freeradbiomed.2004.07.021.
  • Melkamu T, Zhang X, Tan J, Zeng Y, Kassie F. Alteration of microRNA expression in vinyl carbamate-induced mouse lung tumors and modulation by the chemopreventive agent indole-3-carbinol. Carcinogenesis. 2010;31(2):252–8. 10.1093/carcin/bgp208
  • Higgins LG, Cavin C, Itoh K, Yamamoto M, Hayes JD. Induction of cancer chemopreventive enzymes by coffee is mediated by transcription factor Nrf2. Evidence that the coffee-specific diterpenes cafestol and kahweol confer protection against acrolein. Toxicol Appl Pharmacol. 2008;226(3):328–37. doi:10.1016/j.taap.2007.09.018.
  • Cavin C, Holzhaeuser D, Scharf G, Constable A, Huber WW, Schilter B. Cafestol and kahweol, two coffee specific diterpenes with anticarcinogenic activity. Food Chem Toxicol. 2002;40(8):1155–63. doi:10.1016/s0278-6915(02)00029-7.
  • Nakamura Y, Yoshida C, Murakami A, Ohigashi H, Osawa T, Uchida K. Zerumbone, a tropical ginger sesquiterpene, activates phase II drug metabolizing enzymes. FEBS Lett. 2004;572(1–3):245–50.
  • Jeong W-S, Jun M, Kong A-NT. Nrf2: a potential molecular target for cancer chemoprevention by natural compounds. Antioxid Redox Signal. 2006;8(1–2):99–106. doi:10.1089/ars.2006.8.99.
  • Stevens JF, Page JE. Xanthohumol and related prenylflavonoids from hops and beer: to your good health!. Phytochemistry. 2004;65(10):1317–30.
  • Yan C, Sun W, Wang X, Long J, Liu X, Feng Z, Liu J. Punicalagin attenuates palmitate‐induced lipotoxicity in HepG2 cells by activating the Keap1‐Nrf2 antioxidant defense system. Mol Nutr Food Res. 2016;60(5):1139–49. doi:10.1002/mnfr.201500490.
  • Seeram N, Lee R, Hardy M, Heber D. Rapid large scale purification of ellagitannins from pomegranate husk, a by-product of the commercial juice industry. Purif Technol. 2005;41(1):49–55. doi:10.1016/j.seppur.2004.04.003.
  • Salehi B, Mishra AP, Nigam M, Sener B, Kilic M, Sharifi-Rad M, et al. Resveratrol: a double-edged sword in health benefits. Biomedicines. 2018;6(3):91.
  • De Vincenzi M, Silano M, Stacchini P, Scazzocchio B. Constituents of aromatic plants: I. Methyleugenol. Fitoterapia. 2000;71(2):216–21. doi:10.1016/s0367-326x(99)00150-1.
  • Tang F, Chen F, Ling X, Huang Y, Zheng X, Tang Q, et al. Inhibitory effect of methyleugenol on IgE-mediated allergic inflammation in RBL-2H3 cells. Mediators Inflamm. 2015;2015:463530. doi:10.1155/2015/463530.
  • Yano S, Suzuki Y, Yuzurihara M, Kase Y, Takeda S, Watanabe S, Aburada M, Miyamoto K-I. Antinociceptive effect of methyleugenol on formalin-induced hyperalgesia in mice. Eur J Pharmacol. 2006;553(1–3):99–103. doi:10.1016/j.ejphar.2006.09.020.
  • Suekawa M, Ishige A, Yuasa K, Sudo K, Aburada M, Hosoya E. Pharmacological studies on ginger. I. Pharmacological actions of pungent constituents,(6)-gingerol and (6)-shogaol. J Pharmacobiodyn. 1984;7(11):836–48. doi:10.1248/bpb1978.7.836.
  • Liu D, Wang H, Zhang Y, Zhang Z. Protective effects of chlorogenic acid on cerebral ischemia/reperfusion injury rats by regulating oxidative stress-related Nrf2 pathway. Drug Des Devel Ther. 2020;14:51–60. doi:10.2147/dddt.S228751.
  • Mancuso C, Santangelo R. Ferulic acid: pharmacological and toxicological aspects. Food Chem Toxicol. 2014;65:185–95. doi:10.1016/j.fct.2013.12.024.
  • Hadad N, Levy R. The synergistic anti-inflammatory effects of lycopene, lutein, β-carotene, and carnosic acid combinations via redox-based inhibition of NF-κB signaling. Free Radic Biol Med. 2012;53(7):1381–91. doi:10.1016/j.freeradbiomed.2012.07.078.
  • Frankel EN, Huang S-W, Aeschbach R, Prior E. Antioxidant activity of a rosemary extract and its constituents, carnosic acid, carnosol, and rosmarinic acid, in bulk oil and oil-in-water emulsion. J Agric Food Chem. 1996;44(1):131–5. doi:10.1021/jf950374p.
  • Larrosa M, García-Conesa MT, Espín JC, Tomás-Barberán FA. Ellagitannins, ellagic acid and vascular health. Mol Aspects Med. 2010;31(6):513–39. doi:10.1016/j.mam.2010.09.005.
  • Yang J, Pi C, Wang G. Inhibition of PI3K/Akt/mTOR pathway by apigenin induces apoptosis and autophagy in hepatocellular carcinoma cells. Biomed Pharmacother. 2018;103:699–707. doi:10.1016/j.biopha.2018.04.072.
  • Cabrera C, Artacho R, Giménez R. Beneficial effects of green tea—a review. J Am Coll Nutr. 2006;25(2):79–99. doi:10.1080/07315724.2006.10719518.
  • Singh BN, Rawat AKS, Bhagat RM, Singh BR. Black tea: phytochemicals, cancer chemoprevention, and clinical studies. Crit Rev Food Sci Nutr. 2017;57(7):1394–410. doi:10.1080/10408398.2014.994700.
  • Huang Z, Jing X, Sheng Y, Zhang J, Hao Z, Wang Z, Ji L. (-)-Epicatechin attenuates hepatic sinusoidal obstruction syndrome by inhibiting liver oxidative and inflammatory injury. Redox Biol. 2019;22:101117. doi:10.1016/j.redox.2019.101117.
  • Abdulkhaleq LA, Assi MA, Noor MHM, Abdullah R, Saad MZ, Taufiq-Yap YH. Therapeutic uses of epicatechin in diabetes and cancer. Vet World. 2017;10(8):869–72. doi:10.14202/vetworld.2017.869-872.
  • Stuart EC, Scandlyn MJ, Rosengren RJ. Role of epigallocatechin gallate (EGCG) in the treatment of breast and prostate cancer. Life Sci. 2006;79(25):2329–36. doi:10.1016/j.lfs.2006.07.036.
  • Arai Y, Watanabe S, Kimira M, Shimoi K, Mochizuki R, Kinae N. Dietary intakes of flavonols, flavones and isoflavones by Japanese women and the inverse correlation between quercetin intake and plasma LDL cholesterol concentration. J Nutr. 2000;130(9):2243–50. doi:10.1093/jn/130.9.2243.
  • Guo J, Yang G, He Y, Xu H, Fan H, An J, Zhang L, Zhang R, Cao G, Hao D, et al. Involvement of α7nAChR in the protective effects of genistein against β-amyloid-induced oxidative stress in neurons via a PI3K/Akt/Nrf2 pathway-related mechanism. Cell Mol Neurobiol. 2021;41(2):377–93. doi:10.1007/s10571-020-01009-8.
  • Yuan L, Wei S, Wang J, Liu X. Isoorientin induces apoptosis and autophagy simultaneously by reactive oxygen species (ROS)-related p53, PI3K/Akt, JNK, and p38 signaling pathways in HepG2 cancer cells. J Agric Food Chem. 2014;62(23):5390–400. doi:10.1021/jf500903g.
  • Lim JH, Park H-S, Choi J-K, Lee I-S, Choi HJ. Isoorientin induces Nrf2 pathway-driven antioxidant response through phosphatidylinositol 3-kinase signaling. Arch Pharm Res. 2007;30(12):1590–8. doi:10.1007/BF02977329.
  • Ji L-L, Sheng Y-C, Zheng Z-Y, Shi L, Wang Z-T. The involvement of p62–Keap1–Nrf2 antioxidative signaling pathway and JNK in the protection of natural flavonoid quercetin against hepatotoxicity. Free Radic Biol Med. 2015;85:12–23. doi:10.1016/j.freeradbiomed.2015.03.035.
  • Tang X, Wang H, Fan L, Wu X, Xin A, Ren H, Wang XJ. Luteolin inhibits Nrf2 leading to negative regulation of the Nrf2/ARE pathway and sensitization of human lung carcinoma A549 cells to therapeutic drugs. Free Radic Biol Med. 2011;50(11):1599–609. doi:10.1016/j.freeradbiomed.2011.03.008.
  • Patel K, Patel DK. The beneficial role of rutin, a naturally occurring flavonoid in health promotion and disease prevention: a systematic review and update. In: Watson RR, Preedy VR, editors. Bioactive food as dietary interventions for arthritis and related inflammatory diseases. 2nd ed. Academic Press; 2019. p. 457–79. doi:10.1016/B978-0-12-813820-5.00026-X.
  • Zhang L, Guo Z, Wang Y, Geng J, Han S. The protective effect of kaempferol on heart via the regulation of Nrf2, NF‐κβ, and PI3K/Akt/GSK‐3β signaling pathways in isoproterenol‐induced heart failure in diabetic rats. Drug Dev Res. 2019;80(3):294–309. doi:10.1002/ddr.21495.
  • Krajka-Kuźniak V, Paluszczak J, Baer-Dubowska W. Xanthohumol induces phase II enzymes via Nrf2 in human hepatocytes in vitro. Toxicol In Vitro. 2013;27(1):149–56. doi:10.1016/j.tiv.2012.10.008.
  • Cykowiak M, Krajka-Kuźniak V, Baer-Dubowska W. Combinations of phytochemicals more efficiently than single components activate Nrf2 and induce the expression of antioxidant enzymes in pancreatic cancer cells. Nutr Cancer. 2022;74(3):996–1011. doi:10.1080/01635581.2021.1933097.
  • Cykowiak M, Kleszcz R, Kucińska M, Paluszczak J, Szaefer H, Plewiński A, Piotrowska-Kempisty H, Murias M, Krajka-Kuźniak V. Attenuation of pancreatic cancer in vitro and in vivo via modulation of Nrf2 and NF-κB signaling pathways by natural compounds. Cells. 2021;10(12):3556. doi:10.3390/cells10123556.
  • Benelli R, Venè R, Ciarlo M, Carlone S, Barbieri O, Ferrari N. The AKT/NF-κB inhibitor xanthohumol is a potent anti-lymphocytic leukemia drug overcoming chemoresistance and cell infiltration. Biochem Pharmacol. 2012;83(12):1634–42. 10.1016/j.bcp.2012.03.006
  • Nishimura R, Tabata K, Arakawa M, Ito Y, Kimura Y, Akihisa T, Nagai H, Sakuma A, Kohno H, Suzuki T, et al. Isobavachalcone, a chalcone constituent of Angelica keiskei, induces apoptosis in neuroblastoma. Biol Pharm Bull. 2007;30(10):1878–83. doi:10.1248/bpb.30.1878.
  • Colgate EC, Miranda CL, Stevens JF, Bray TM, Ho E. Xanthohumol, a prenylflavonoid derived from hops induces apoptosis and inhibits NF-kappaB activation in prostate epithelial cells. Cancer Lett. 2007;246(1–2):201–9.
  • Deeb D, Gao X, Jiang H, Arbab AS, Dulchavsky S, Gautam SC. Growth inhibitory and apoptosis-inducing effects of xanthohumol, a prenylated chalone present in hops, in human prostate cancer cells. Anticancer Res. 2010;30(9):3333–9.
  • Zhang L, Chinnathambi A, Alharbi SA, Veeraraghavan VP, Mohan SK, Zhang G. Punicalagin promotes the apoptosis in human cervical cancer (ME-180) cells through mitochondrial pathway and by inhibiting the NF-kB signaling pathway. Saudi J Biol Sci. 2020;27(4):1100–6. doi:10.1016/j.sjbs.2020.02.015.
  • Huang T, Zhang X, Wang H. Punicalagin inhibited proliferation, invasion and angiogenesis of osteosarcoma through suppression of NF‑κB signaling. Mol Med Rep. 2020;22(3):2386–94. doi:10.3892/mmr.2020.11304.
  • Adams LS, Seeram NP, Aggarwal BB, Takada Y, Sand D, Heber D. Pomegranate juice, total pomegranate ellagitannins, and punicalagin suppress inflammatory cell signaling in colon cancer cells. J Agric Food Chem. 2006;54(3):980–5. doi:10.1021/jf052005r.
  • Berdowska I, Matusiewicz M, Fecka I. Punicalagin in cancer prevention—Via signaling pathways targeting. Nutrients. 2021;13(8):2733. doi:10.3390/nu13082733.
  • Singh B, Shoulson R, Chatterjee A, Ronghe A, Bhat NK, Dim DC, et al. Resveratrol inhibits estrogen-induced breast carcinogenesis through induction of NRF2-mediated protective pathways. Carcinogenesis. 2014;35(8):1872–80.
  • Farkhondeh T, Folgado SL, Pourbagher-Shahri AM, Ashrafizadeh M, Samarghandian S. The therapeutic effect of resveratrol: focusing on the Nrf2 signaling pathway. Biomed Pharmacother. 2020;127:110234. doi:10.1016/j.biopha.2020.110234.
  • Bishayee A, Barnes KF, Bhatia D, Darvesh AS, Carroll RT. Resveratrol suppresses oxidative stress and inflammatory response in diethylnitrosamine-initiated rat hepatocarcinogenesis. Cancer Prev Res (Phila). 2010;3(6):753–63. doi:10.1158/1940-6207.CAPR-09-0171.
  • Benitez DA, Hermoso MA, Pozo‐Guisado E, Fernández‐Salguero PM, Castellón EA. Regulation of cell survival by resveratrol involves inhibition of NFκB‐regulated gene expression in prostate cancer cells. Prostate. 2009;69(10):1045–54.
  • Whitlock NC, Baek SJ. The anticancer effects of resveratrol: modulation of transcription factors. Nutr Cancer. 2012;64(4):493–502. doi:10.1080/01635581.2012.667862.
  • Carter LG, D’Orazio JA, Pearson KJ. Resveratrol and cancer: focus on in vivo evidence. Endocr Relat Cancer. 2014;21(3):R209–25.
  • Yi J-L, Shi S, Shen Y-L, Wang L, Chen H-Y, Zhu J, et al. Myricetin and methyl eugenol combination enhances the anticancer activity, cell cycle arrest and apoptosis induction of cis-platin against HeLa cervical cancer cell lines. Int J Clin Exp Path. 2015;8(2):1116.
  • Yin L, Sun Z, Ren Q, Su X, Zhang D. Methyl eugenol induces potent anticancer effects in RB355 human retinoblastoma cells by inducing autophagy, cell cycle arrest and inhibition of PI3K/mTOR/Akt signalling pathway. J Buon. 2018;23:1174–8.
  • Groh IAM, Chen C, Lüske C, Cartus AT, Esselen M. Plant polyphenols and oxidative metabolites of the herbal alkenylbenzene methyleugenol suppress histone deacetylase activity in human colon carcinoma cells. J Nutr Metab. 2013;2013:821082. doi:10.1155/2013/821082.
  • Ling H, Yang H, Tan SH, Chui WK, Chew EH. 6‐Shogaol, an active constituent of ginger, inhibits breast cancer cell invasion by reducing matrix metalloproteinase‐9 expression via blockade of nuclear factor‐κB activation. Br J Pharmacol. 2010;161(8):1763–77. doi:10.1111/j.1476-5381.2010.00991.x.
  • Hung J-Y, Hsu Y-L, Li C-T, Ko Y-C, Ni W-C, Huang M-S, Kuo P-L. 6-Shogaol, an active constituent of dietary ginger, induces autophagy by inhibiting the AKT/mTOR pathway in human non-small cell lung cancer A549 cells. J Agric Food Chem. 2009;57(20):9809–16. doi:10.1021/jf902315e.
  • Saha A, Blando J, Silver E, Beltran L, Sessler J, DiGiovanni J. 6-Shogaol from dried ginger inhibits growth of prostate cancer cells both in vitro and in vivo through inhibition of STAT3 and NF-κB signaling. Cancer Prev Res (Phila). 2014;7(6):627–38. doi:10.1158/1940-6207.CAPR-13-0420.
  • Zhou L, Qi L, Jiang L, Zhou P, Ma J, Xu X, Li P. Antitumor activity of gemcitabine can be potentiated in pancreatic cancer through modulation of TLR4/NF-κB signaling by 6-shogaol. Aaps J. 2014;16(2):246–57. doi:10.1208/s12248-013-9558-3.
  • Ma R-H, Ni Z-J, Zhang F, Zhang Y-Y, Liu M-M, Thakur K, et al. 6-Shogaol mediated ROS production and apoptosis via endoplasmic reticulum and mitochondrial pathways in human endometrial carcinoma Ishikawa cells. J Funct Foods. 2020;74:104178.
  • Ma R-H, Ni Z-J, Thakur K, Cespedes-Acuña CL, Zhang J-G, Wei Z-J. Transcriptome and proteomics conjoint analysis reveal metastasis inhibitory effect of 6-shogaol as ferroptosis activator through the PI3K/AKT pathway in human endometrial carcinoma in vitro and in vivo. Food Chem Toxicol. 2022;170:113499. doi:10.1016/j.fct.2022.113499.
  • Warin RF, Chen H, Soroka DN, Zhu Y, Sang S. Induction of lung cancer cell apoptosis through a p53 pathway by [6]-shogaol and its cysteine-conjugated metabolite M2. J Agric Food Chem. 2014;62(6):1352–62.
  • Boettler U, Volz N, Pahlke G, Teller N, Kotyczka C, Somoza V, Stiebitz H, Bytof G, Lantz I, Lang R, et al. Coffees rich in chlorogenic acid or N‐methylpyridinium induce chemopreventive phase II‐enzymes via the Nrf2/ARE pathway in vitro and in vivo. Mol Nutr Food Res. 2011;55(5):798–802. doi:10.1002/mnfr.201100115.
  • Deka SJ, Gorai S, Manna D, Trivedi V. Evidence of PKC binding and translocation to explain the anticancer mechanism of chlorogenic acid in breast cancer cells. Curr Mol Med. 2017;17(1):79–89. doi:10.2174/1566524017666170209160619.
  • Han D, Chen W, Gu X, Shan R, Zou J, Liu G, Shahid M, Gao J, Han B. Cytoprotective effect of chlorogenic acid against hydrogen peroxide-induced oxidative stress in MC3T3-E1 cells through PI3K/Akt-mediated Nrf2/HO-1 signaling pathway. Oncotarget. 2017;8(9):14680–92. doi:10.18632/oncotarget.14747.
  • Wang X, Liu J, Xie Z, Rao J, Xu G, Huang K, Li W, Yin Z. Chlorogenic acid inhibits proliferation and induces apoptosis in A498 human kidney cancer cells via inactivating PI3K/Akt/mTOR signalling pathway. J Pharm Pharmacol. 2019;71(7):1100–9. doi:10.1111/jphp.13095.
  • Yamagata K, Izawa Y, Onodera D, Tagami M. Chlorogenic acid regulates apoptosis and stem cell marker-related gene expression in A549 human lung cancer cells. Mol Cell Biochem. 2018;441(1–2):9–19. doi:10.1007/s11010-017-3171-1.
  • Hou N, Liu N, Han J, Yan Y, Li J. Chlorogenic acid induces reactive oxygen species generation and inhibits the viability of human colon cancer cells. Anticancer Drugs. 2017;28(1):59–65. doi:10.1097/cad.0000000000000430.
  • Yan Y, Li J, Han J, Hou N, Song Y, Dong L. Chlorogenic acid enhances the effects of 5-fluorouracil in human hepatocellular carcinoma cells through the inhibition of extracellular signal-regulated kinases. Anticancer Drugs. 2015;26(5):540–6. doi:10.1097/cad.0000000000000218.
  • Wang L, Du H, Chen P. Chlorogenic acid inhibits the proliferation of human lung cancer A549 cell lines by targeting annexin A2 in vitro and in vivo. Biomed Pharmacother. 2020;131:110673.
  • Yan Y, Liu N, Hou N, Dong L, Li J. Chlorogenic acid inhibits hepatocellular carcinoma in vitro and in vivo. J Nutr Biochem. 2017;46:68–73. doi:10.1016/j.jnutbio.2017.04.007.
  • Zeng A, Liang X, Zhu S, Liu C, Wang S, Zhang Q, Zhao J, Song L. Chlorogenic acid induces apoptosis, inhibits metastasis and improves antitumor immunity in breast cancer via the NF‑κB signaling pathway. Oncol Rep. 2021;45(2):717–27. doi:10.3892/or.2020.7891.
  • Li Y, Li X, Cuiping C, Pu R, Weihua Y. Study on the anticancer effect of an astragaloside-and chlorogenic acid-containing herbal medicine (RLT-03) In breast cancer. Evid-based Complement Altern Med. 2020;2020:1515081. doi:10.1155/2020/1515081.
  • Gao J, Yu H, Guo W, Kong Y, Gu L, Li Q, Yang S, Zhang Y, Wang Y. The anticancer effects of ferulic acid is associated with induction of cell cycle arrest and autophagy in cervical cancer cells. Cancer Cell Int. 2018;18:102. doi:10.1186/s12935-018-0595-y.
  • Roy N, Narayanankutty A, Nazeem PA, Valsalan R, Babu TD, Mathew D. Plant phenolics ferulic acid and P-coumaric acid inhibit colorectal cancer cell proliferation through EGFR down-regulation. Asian Pac J Cancer Prev: APJCP. 2016;17(8):4019–23.
  • Eroğlu C, Seçme M, Bağcı G, Dodurga Y. Assessment of the anticancer mechanism of ferulic acid via cell cycle and apoptotic pathways in human prostate cancer cell lines. Tumour Biol. 2015;36(12):9437–46. doi:10.1007/s13277-015-3689-3.
  • Zhang X, Lin D, Jiang R, Li H, Wan J, Li H. Ferulic acid exerts antitumor activity and inhibits metastasis in breast cancer cells by regulating epithelial to mesenchymal transition. Oncol Rep. 2016;36(1):271–8. doi:10.3892/or.2016.4804.
  • El-Gogary RI, Nasr M, Rahsed LA, Hamzawy MA. Ferulic acid nanocapsules as a promising treatment modality for colorectal cancer: preparation and in vitro/in vivo appraisal. Life Sci. 2022;298:120500. 10.1016/j.lfs.2022.120500
  • Yang G-W, Jiang J-S, Lu W-Q. Ferulic acid exerts anti-angiogenic and anti-tumor activity by targeting fibroblast growth factor receptor 1-mediated angiogenesis. Int J Mol Sci. 2015;16(10):24011–31.
  • Barni MV, Carlini MJ, Cafferata EG, Puricelli L, Moreno S. Carnosic acid inhibits the proliferation and migration capacity of human colorectal cancer cells. Oncol Rep. 2012;27(4):1041–8. doi:10.3892/or.2012.1630.
  • Khella KF, Abd E, Maksoud AI, Hassan A, Abdel-Ghany SE, Elsanhoty RM, Aladhadh MA, et al. Carnosic acid encapsulated in albumin nanoparticles induces apoptosis in breast and colorectal cancer cells. Molecules. 2022;27(13):4102.
  • Tang B, Tang F, Wang Z, Qi G, Liang X, Li B, Yuan S, Liu J, Yu S, He S, et al. Upregulation of Akt/NF-κB-regulated inflammation and Akt/Bad-related apoptosis signaling pathway involved in hepatic carcinoma process: suppression by carnosic acid nanoparticle. Int J Nanomedicine. 2016;11:6401–20. doi:10.2147/ijn.S101285.
  • El-Huneidi W, Bajbouj K, Muhammad JS, Vinod A, Shafarin J, Khoder G, et al. Carnosic acid induces apoptosis and inhibits Akt/mTOR signaling in human gastric cancer cell lines. Pharmaceuticals (Basel, Switzerland). 2021;14(3):230–239. doi:10.3390/ph14030230.
  • Kim D-H, Park K-W, Chae IG, Kundu J, Kim E-H, Kundu JK, Chun K-S. Carnosic acid inhibits STAT3 signaling and induces apoptosis through generation of ROS in human colon cancer HCT116 cells. Mol Carcinog. 2016;55(6):1096–110. doi:10.1002/mc.22353.
  • Su K, Wang CF, Zhang Y, Cai YJ, Zhang YY, Zhao Q. The inhibitory effects of carnosic acid on cervical cancer cells growth by promoting apoptosis via ROS-regulated signaling pathway. Biomed Pharmacother. 2016;82:180–91. doi:10.1016/j.biopha.2016.04.056.
  • Sang Y, Zhang F, Wang H, Yao J, Chen R, Zhou Z, Yang K, Xie Y, Wan T, Ding H, et al. Apigenin exhibits protective effects in a mouse model of d-galactose-induced aging via activating the Nrf2 pathway. Food Funct. 2017;8(6):2331–40. doi:10.1039/c7fo00037e.
  • Liu X, Dong S, Dong M, Li Y, Sun Z, Zhang X, et al. Transferrin-conjugated liposomes loaded with carnosic acid inhibit liver cancer growth by inducing mitochondria-mediated apoptosis. Int J Pharm. 2021;607:121034.
  • Yan M, Vemu B, Veenstra J, Petiwala SM, Johnson JJ. Carnosol, a dietary diterpene from rosemary (Rosmarinus officinalis) activates Nrf2 leading to sestrin 2 induction in colon cells. Integr Mol Med. 2018;5(4). doi:10.15761/imm.1000335.
  • Johnson JJ, Syed DN, Heren CR, Suh Y, Adhami VM, Mukhtar H. Carnosol, a dietary diterpene, displays growth inhibitory effects in human prostate cancer PC3 cells leading to G2-phase cell cycle arrest and targets the 5’-AMP-activated protein kinase (AMPK) pathway. Pharm Res. 2008;25(9):2125–34. doi:10.1007/s11095-008-9552-0.
  • Alsamri H, Hasasna E, Al H, Dhaheri Y, Eid AH, Attoub S, Iratni R. Carnosol, a natural polyphenol, inhibits migration, metastasis, and tumor growth of breast cancer via a ROS-dependent proteasome degradation of STAT3. Front Oncol. 2019;9:743. doi:10.3389/fonc.2019.00743.
  • Tong XP, Ma YX, Quan DN, Zhang L, Yan M, Fan XR. Rosemary extracts upregulate Nrf2, Sestrin2, and MRP2 protein level in human hepatoma HepG2 cells. Evid Based Complement Alternat Med. 2017;2017:7359806. doi:10.1155/2017/7359806.
  • Wang Y, Ren F, Li B, Song Z, Chen P, Ouyang L. Ellagic acid exerts antitumor effects via the PI3K signaling pathway in endometrial cancer. J Cancer. 2019;10(15):3303.
  • Wang N, Wang Z-Y, Mo S-L, Loo TY, Wang D-M, Luo H-B, Yang D-P, Chen Y-L, Shen J-G, Chen J-P, et al. Ellagic acid, a phenolic compound, exerts anti-angiogenesis effects via VEGFR-2 signaling pathway in breast cancer. Breast Cancer Res Treat. 2012;134(3):943–55. doi:10.1007/s10549-012-1977-9.
  • Eskandari E, Heidarian E, Amini SA, Saffari-Chaleshtori J. Evaluating the effects of ellagic acid on pSTAT3, pAKT, and pERK1/2 signaling pathways in prostate cancer PC3 cells. J Cancer Res Ther. 2016;12(4):1266–71. doi:10.4103/0973-1482.165873.
  • Cheng H, Lu C, Tang R, Pan Y, Bao S, Qiu Y, et al. Ellagic acid inhibits the proliferation of human pancreatic carcinoma PANC-1 cells in vitro and in vivo. Oncotarget. 2017;8(7):12301.
  • Li TM, Chen GW, Su CC, Lin JG, Yeh CC, Cheng KC, et al. Ellagic acid induced p53/p21 expression, G1 arrest and apoptosis in human bladder cancer T24 cells. Anticancer Res. 2005;25(2a):971–9.
  • Duan J, Li Y, Gao H, Yang D, He X, Fang Y, Zhou G. Phenolic compound ellagic acid inhibits mitochondrial respiration and tumor growth in lung cancer. Food Funct. 2020;11(7):6332–9.
  • Paredes-Gonzalez X, Fuentes F, Su Z-Y, Kong A-NT. Apigenin reactivates Nrf2 anti-oxidative stress signaling in mouse skin epidermal JB6 P + cells through epigenetics modifications. AAPS J. 2014;16(4):727–35.
  • Mahmoudi S, Ghorbani M, Sabzichi M, Ramezani F, Hamishehkar H, Samadi N. Targeted hyaluronic acid-based lipid nanoparticle for apigenin delivery to induce Nrf2-dependent apoptosis in lung cancer cells. J Drug Deliv Sci Technol. 2019;49:268–76. doi:10.1016/j.jddst.2018.11.013.
  • Kaur P, Shukla S, Gupta S. Plant flavonoid apigenin inactivates Akt to trigger apoptosis in human prostate cancer: an in vitro and in vivo study. Carcinogenesis. 2008;29(11):2210–7.
  • Shukla S, Gupta S. Apigenin suppresses insulin‐like growth factor I receptor signaling in human prostate cancer: an in vitro and in vivo study. Mol Carcinog. 2009;48(3):243–52.
  • Chen X, Xu H, Yu X, Wang X, Zhu X, Xu X. Apigenin inhibits in vitro and in vivo tumorigenesis in cisplatin-resistant colon cancer cells by inducing autophagy, programmed cell death and targeting m-TOR/PI3K/Akt signalling pathway. J Buon. 2019;24(2):488–93.
  • Budhraja A, Gao N, Zhang Z, Son Y-O, Cheng S, Wang X, et al. Apigenin induces apoptosis in human leukemia cells and exhibits anti-leukemic activity in vivo. Mol Cancer Ther. 2012;11(1):132–42.
  • Zhang E, Zhang Y, Fan Z, Cheng L, Han S, Che H. Apigenin inhibits histamine-induced cervical cancer tumor growth by regulating estrogen receptor expression. Molecules. 2020;25(8):1960.
  • Khojaste E, Ahmadizadeh C. Catechin metabolites along with curcumin inhibit proliferation and induce apoptosis in cervical cancer cells by regulating VEGF expression in-vitro. Nutr Cancer. 2022;74(3):1048–57.
  • Manikandan R, Beulaja M, Arulvasu C, Sellamuthu S, Dinesh D, Prabhu D. Synergistic anticancer activity of curcumin and catechin: an in vitro study using human cancer cell lines. Microsc Res Tech. 2012;75(2):112–6.
  • Pereyra-Vergara F, Olivares-Corichi IM, Perez-Ruiz AG, Luna-Arias JP, García-Sánchez JR. Apoptosis induced by (−)-epicatechin in human breast cancer cells is mediated by reactive oxygen species. Molecules. 2020;25(5):1020.
  • Rodriguez M, Du G-J, Wang C-Z, Yuan C-S. Panaxadiol’s anticancer activity is enhanced by epicatechin. Am J Chin Med. 2010;38(6):1233–5. doi:10.1142/S0192415X10008597.
  • Horie N, Hirabayashi N, Takahashi Y, Miyauchi Y, Taguchi H, Takeishi K. Synergistic effect of green tea catechins on cell growth and apoptosis induction in gastric carcinoma cells. Biol Pharm Bull. 2005;28(4):574–9. doi:10.1248/bpb.28.574.
  • Chen L, Guo Y, Wu Z, Zhao S, Zhang Z, Zheng F, Sun L, Hao Z, Xu C, Wang T, et al. Epicatechin gallate prevents the de novo synthesis of fatty acid and the migration of prostate cancer cells. Acta Biochim Biophys Sin (Shanghai). 2021;53(12):1662–9. doi:10.1093/abbs/gmab144.
  • Ahmadi N, Mohamed S, Rahman S, Rosli H. R. Epicatechin and scopoletin‐rich Morinda citrifolia leaf ameliorated leukemia via anti‐inflammatory, anti‐angiogenesis, and apoptosis pathways in vitro and in vivo. J Food Biochem. 2019;43(7):e12868.
  • Enkhbat T, Nishi M, Yoshikawa K, Jun H, Tokunaga T, Takasu C, Kashihara H, Ishikawa D, Tominaga M, Shimada M, et al. Epigallocatechin-3-gallate enhances radiation sensitivity in colorectal cancer cells through Nrf2 activation and autophagy. Anticancer Res. 2018;38(11):6247–52. doi:10.21873/anticanres.12980.
  • Datta S, Sinha D. EGCG maintained Nrf2-mediated redox homeostasis and minimized etoposide resistance in lung cancer cells. J Func Foods. 2019;62:103553. doi:10.1016/j.jff.2019.103553.
  • Wang J, Man GCW, Chan TH, Kwong J, Wang CC. A prodrug of green tea polyphenol (–)-epigallocatechin-3-gallate (Pro-EGCG) serves as a novel angiogenesis inhibitor in endometrial cancer. Cancer Lett. 2018;412:10–20. doi:10.1016/j.canlet.2017.09.054.
  • Roy AM, Baliga MS, Katiyar SK. Epigallocatechin-3-gallate induces apoptosis in estrogen receptor–negative human breast carcinoma cells via modulation in protein expression of p53 and Bax and caspase-3 activation. Mol Cancer Ther. 2005;4(1):81–90.
  • Gan R-Y, Li H-B, Sui Z-Q, Corke H. Absorption, metabolism, anti-cancer effect and molecular targets of epigallocatechin gallate (EGCG): an updated review. Crit Rev Food Sci Nutr. 2018;58(6):924–41.
  • Kim N, Lee SH, Son JH, Lee JM, Kang M-J, Kim BH, Lee J-S, Ryu JK, Kim Y-T. Fisetin reduces cell viability through up-regulation of phosphorylation of ERK1/2 in cholangiocarcinoma cells. Anticancer Res. 2016;36(11):6109–16. doi:10.21873/anticanres.11201.
  • Park B-S, Choi N-E, Lee JH, Kang H-M, Yu S-B, Kim H-J, Kang H-K, Kim I-R. Crosstalk between fisetin-induced apoptosis and autophagy in human oral squamous cell carcinoma. J Cancer. 2019;10(1):138–46. doi:10.7150/jca.28500.
  • Sun X, Ma X, Li Q, Yang Y, Xu X, Sun J, et al. Anti‑cancer effects of fisetin on mammary carcinoma cells via regulation of the PI3K/Akt/mTOR pathway: in vitro and in vivo studies. Int J Mol Med. 2018;42(2):811–20.
  • Li J, Cheng Y, Qu W, Sun Y, Wang Z, Wang H, Tian B. Fisetin, a dietary flavonoid, induces cell cycle arrest and apoptosis through activation of p53 and inhibition of NF‐kappa B pathways in bladder cancer cells. Basic Clin Pharmacol Toxicol. 2011;108(2):84–93. doi:10.1111/j.1742-7843.2010.00613.x.
  • Suh Y, Afaq F, Johnson JJ, Mukhtar H. A plant flavonoid fisetin induces apoptosis in colon cancer cells by inhibition of COX2 and Wnt/EGFR/NF-κB-signaling pathways. Carcinogenesis. 2009;30(2):300–7.
  • Youns M, Abdel Halim Hegazy W. The natural flavonoid fisetin inhibits cellular proliferation of hepatic, colorectal, and pancreatic cancer cells through modulation of multiple signaling pathways. PLoS One. 2017;12(1):e0169335. doi:10.1371/journal.pone.0169335.
  • Khan N, Afaq F, Syed DN, Mukhtar H. Fisetin, a novel dietary flavonoid, causes apoptosis and cell cycle arrest in human prostate cancer LNCaP cells. Carcinogenesis. 2008;29(5):1049–56.
  • Xiao Y, Liu Y, Gao Z, Li X, Weng M, Shi C, et al. Fisetin inhibits the proliferation, migration and invasion of pancreatic cancer by targeting PI3K/AKT/mTOR signaling. Aging (Albany NY). 2021;13(22):24753.
  • Ying T-H, Yang S-F, Tsai S-J, Hsieh S-C, Huang Y-C, Bau D-T, Hsieh Y-H. Fisetin induces apoptosis in human cervical cancer HeLa cells through ERK1/2-mediated activation of caspase-8-/caspase-3-dependent pathway. Arch Toxicol. 2012;86(2):263–73.
  • Davis JN, Kucuk O, Sarkar FH. Genistein inhibits NF-kB activation in prostate cancer cells. Nutr Cancer. 1999;35(2):167–74. doi:10.1207/S15327914NC352_11.
  • Banerjee S, Zhang Y, Ali S, Bhuiyan M, Wang Z, Chiao PJ, Philip PA, Abbruzzese J, Sarkar FH. Molecular evidence for increased antitumor activity of gemcitabine by genistein in vitro and in vivo using an orthotopic model of pancreatic cancer. Cancer Res. 2005;65(19):9064–72.
  • Pagliacci M, Smacchia M, Migliorati G, Grignani F, Riccardi C, Nicoletti I. Growth-inhibitory effects of the natural phyto-oestrogen genistein in MCF-7 human breast cancer cells. Eur J Cancer. 1994;30(11):1675–82. doi:10.1016/0959-8049(94)00262-4.
  • Hsiao YC, Peng SF, Lai KC, Liao CL, Huang YP, Lin CC, et al. Genistein induces apoptosis in vitro and has antitumor activity against human leukemia HL‐60 cancer cell xenograft growth in vivo. Environ Toxicol. 2019;34(4):443–56.
  • Wang Y, Wang H, Zhang W, Shao C, Xu P, Shi CH, et al. Genistein sensitizes bladder cancer cells to HCPT treatment in vitro and in vivo via ATM/NF-κB/IKK pathway-induced apoptosis. PLoS One. 2013;8(1):e50175.
  • Liu D, Yan L, Wang L, Tai W, Wang W, Yang C. Genistein enhances the effect of cisplatin on the inhibition of non‑small cell lung cancer A549 cell growth in vitro and in vivo. Oncol Lett. 2014;8(6):2806–10.
  • Yuan L, Wang J, Xiao H, Xiao C, Wang Y, Liu X. Isoorientin induces apoptosis through mitochondrial dysfunction and inhibition of PI3K/Akt signaling pathway in HepG2 cancer cells. Toxicol Appl Pharmacol. 2012;265(1):83–92. doi:10.1016/j.taap.2012.09.022.
  • Ye T, Su J, Huang C, Yu D, Dai S, Huang X, Chen B, Zhou M. Isoorientin induces apoptosis, decreases invasiveness, and downregulates VEGF secretion by activating AMPK signaling in pancreatic cancer cells. Onco Targets Ther. 2016;9:7481–92. doi:10.2147/OTT.S122653.
  • Zheng H, Zhang M, Luo H, Li H. Isoorientin alleviates UVB-induced skin injury by regulating mitochondrial ROS and cellular autophagy. Biochem Biophys Res Commun. 2019;514(4):1133–9. doi:10.1016/j.bbrc.2019.04.195.
  • Xu W-T, Shen G-N, Li T-Z, Zhang Y, Zhang T, Xue H, Zuo W-B, Li Y-N, Zhang D-J, Jin C-H, et al. Isoorientin induces the apoptosis and cell cycle arrest of A549 human lung cancer cells via the ROS‑regulated MAPK, STAT3 and NF‑κB signaling pathways. Int J Oncol. 2020;57(2):550–61. doi:10.3892/ijo.2020.5079.
  • Liu S-C, Huang C-S, Huang C-M, Hsieh M-S, Huang M-S, Fong I-H, Yeh C-T, Lin C-C. Isoorientin inhibits epithelial-to-mesenchymal properties and cancer stem-cell-like features in oral squamous cell carcinoma by blocking Wnt/β-catenin/STAT3 axis. Toxicol Appl Pharmacol. 2021;424:115581. 10.1016/j.taap.2021.115581
  • Ramos AA, Lima CF, Pereira M, Fernandes-Ferreira M, Pereira-Wilson C. Antigenotoxic effects of quercetin, rutin and ursolic acid on HepG2 cells: evaluation by the comet assay. Toxicol Lett. 2008;177(1):66–73. doi:10.1016/j.toxlet.2008.01.001.
  • Darband SG, Sadighparvar S, Yousefi B, Kaviani M, Ghaderi-Pakdel F, Mihanfar A, Rahimi Y, Mobaraki K, Majidinia M. Quercetin attenuated oxidative DNA damage through NRF2 signaling pathway in rats with DMH induced colon carcinogenesis. Life Sci. 2020;253:117584. doi:10.1016/j.lfs.2020.117584.
  • Mostafavi Pour Z, Ramezani F, Keshavarzi F, Samadi N. ‑The role of quercetin and vitamin C in Nrf2‑dependent oxidative stress production in breast cancer cells. Oncol Lett. 2017;13(3):1965–73. doi:10.3892/ol.2017.5619.
  • Abbasi A, Mostafavi-Pour Z, Amiri A, Keshavarzi F, Nejabat N, Ramezani F, Sardarian A, Zal F. Chemoprevention of prostate cancer cells by vitamin C plus quercetin: role of Nrf2 in inducing oxidative stress. Nutr Cancer. 2021;73(10):2003–13. doi:10.1080/01635581.2020.1819346.
  • Lee Y-J, Lee DM, Lee S-H. Nrf2 expression and apoptosis in quercetin-treated malignant mesothelioma cells. Mol Cells. 2015;38(5):416–25. doi:10.14348/molcells.2015.2268.
  • Sharmila G, Bhat FA, Arunkumar R, Elumalai P, Raja Singh P, Senthilkumar K, Arunakaran J. Chemopreventive effect of quercetin, a natural dietary flavonoid on prostate cancer in in vivo model. Clin Nutr. 2014;33(4):718–26. 10.1016/j.clnu.2013.08.011
  • Zhaorigetu, Farrag IM, Belal A, Badawi MHA, Abdelhady AA, Galala FMAA, El-Sharkawy A, El-Dahshan AA, Mehany, ABM. Antiproliferative, apoptotic effects and suppression of oxidative stress of quercetin against induced toxicity in lung cancer cells of rats: in vitro and in vivo study. J Cancer. 2021;12(17):5249–59. 10.7150/jca.52088
  • Kang KA, Piao MJ, Hyun YJ, Zhen AX, Cho SJ, Ahn MJ, Yi JM, Hyun JW. Luteolin promotes apoptotic cell death via upregulation of Nrf2 expression by DNA demethylase and the interaction of Nrf2 with p53 in human colon cancer cells. Exp Mol Med. 2019;51(4):1–14. doi:10.1038/s12276-019-0238-y.
  • Zuo Q, Wu R, Xiao X, Yang C, Yang Y, Wang C, Lin L, Kong A-N. The dietary flavone luteolin epigenetically activates the Nrf2 pathway and blocks cell transformation in human colorectal cancer HCT116 cells. J Cell Biochem. 2018;119(11):9573–82. doi:10.1002/jcb.27275.
  • Sabzichi M, Hamishehkar H, Ramezani F, Sharifi S, Tabasinezhad M, Pirouzpanah M, Ghanbari P, Samadi N. Luteolin-loaded phytosomes sensitize human breast carcinoma MDA-MB 231 cells to doxorubicin by suppressing Nrf2 mediated signalling. Asian Pac J Cancer Prev. 2014;15(13):5311–6. doi:10.7314/apjcp.2014.15.13.5311.
  • Iida K, Naiki T, Naiki-Ito A, Suzuki S, Kato H, Nozaki S, Nagai T, Etani T, Nagayasu Y, Ando R, et al. Luteolin suppresses bladder cancer growth via regulation of mechanistic target of rapamycin pathway. Cancer Sci. 2020;111(4):1165–79. 10.1111/cas.14334
  • Feng J, Zheng T, Hou Z, Lv C, Xue A, Han T, Han B, Sun X, Wei Y. Luteolin, an aryl hydrocarbon receptor ligand, suppresses tumor metastasis in vitro and in vivo. Oncol Rep. 2020;44(5):2231–40. 10.3892/or.2020.7781
  • Yao X, Jiang W, Yu D, Yan Z. Luteolin inhibits proliferation and induces apoptosis of human melanoma cells in vivo and in vitro by suppressing MMP-2 and MMP-9 through the PI3K/AKT pathway. Food Funct. 2019;10(2):703–12. 10.1039/c8fo02013b
  • Johnson JL, Dia VP, Wallig M, De Mejia EG. Luteolin and gemcitabine protect against pancreatic cancer in an orthotopic mouse model. Pancreas. 2015;44(1):144–51. 10.1097/MPA.0000000000000215
  • Qin T, Zhao J, Liu X, Li L, Zhang X, Shi X, Ke Y, Liu W, Huo J, Dong Y, et al. Luteolin combined with low‐dose paclitaxel synergistically inhibits epithelial–mesenchymal transition and induces cell apoptosis on esophageal carcinoma in vitro and in vivo. Phytother Res. 2021;35(11):6228–40. 10.1002/ptr.7267
  • Zhao J, Li L, Wang Z, Li L, He M, Han S, Dong Y, Liu X, Zhao W, Ke Y, et al. Luteolin attenuates cancer cell stemness in PTX-resistant oesophageal cancer cells through mediating SOX2 protein stability. Pharmacol Res. 2021;174:105939. 10.1016/j.phrs.2021.105939
  • Nafees S, Mehdi SH, Zafaryab M, Zeya B, Sarwar T, Rizvi MA. Synergistic interaction of rutin and silibinin on human colon cancer cell line. Arch Med Res. 2018;49(4):226–34. doi:10.1016/j.arcmed.2018.09.008.
  • Xie X, Feng J, Kang Z, Zhang S, Zhang L, Zhang Y, et al. Taxifolin protects RPE cells against oxidative stress-induced apoptosis. Mol Vis. 2017;23:520.
  • Li Q, Ren L, Zhang Y, Gu Z, Tan Q, Zhang T, et al. P38 signal transduction pathway has more cofactors on apoptosis of SGC-7901 gastric cancer cells induced by combination of rutin and oxaliplatin. BioMed Res Int. 2019;2019:6407210. doi:10.1155/2019/6407210.
  • Gao Y, Liu Z, Zhang X, He J, Pan Y, Hao F, Xie L, Li Q, Qiu X, Wang E, et al. Inhibition of cytoplasmic GSK-3β increases cisplatin resistance through activation of Wnt/β-catenin signaling in A549/DDP cells. Cancer Lett. 2013;336(1):231–9. doi:10.1016/j.canlet.2013.05.005.
  • Satari A, Amini SA, Raeisi E, Lemoigne Y, Heidarian E. Synergetic impact of combined 5-fluorouracil and rutin on apoptosis in pc3 cancer cells through the modulation of p53 gene expression. Adv Pharm Bull. 2019;9(3):462–9. doi:10.15171/apb.2019.055.
  • Elsayed HE, Ebrahim HY, Mohyeldin MM, Siddique AB, Kamal AM, Haggag EG, El Sayed KA. Rutin as a novel c-Met inhibitory lead for the control of triple negative breast malignancies. Nutr Cancer. 2017;69(8):1256–71. doi:10.1080/01635581.2017.1367936.
  • Zhang P, Sun S, Li N, Ho ASW, Kiang KMY, Zhang X, Cheng YS, Poon MW, Lee D, Pu JKS, et al. Rutin increases the cytotoxicity of temozolomide in glioblastoma via autophagy inhibition. J Neurooncol. 2017;132(3):393–400. doi:10.1007/s11060-017-2387-y.
  • Park MH, Kim S, Song Y-r, Kim S, Kim H-J, Na HS, et al. Rutin induces autophagy in cancer cells. Int J Oral Biol. 2016;41(1):45–51. doi:10.11620/IJOB.2016.41.1.045.
  • Imani A, Maleki N, Bohlouli S, Kouhsoltani M, Sharifi S, Maleki Dizaj S. Molecular mechanisms of anticancer effect of rutin. Phytother Res. 2021;35(5):2500–13.
  • Farha AK, Gan R-Y, Li H-B, Wu D-T, Atanasov AG, Gul K, Zhang J-R, Yang Q-Q, Corke H. The anticancer potential of the dietary polyphenol rutin: current status, challenges, and perspectives. Crit Rev Food Sci Nutr. 2022;62(3):832–59.
  • Ko JH, Sethi G, Um JY, Shanmugam MK, Arfuso F, Kumar AP, et al. The role of resveratrol in cancer therapy. Int J Mol Sci. 2017;18(12):2589. doi:10.3390/ijms18122589.
  • Wang F, Wang L, Qu C, Chen L, Geng Y, Cheng C, Yu S, Wang D, Yang L, Meng Z, et al. Kaempferol induces ROS-dependent apoptosis in pancreatic cancer cells via TGM2-mediated Akt/mTOR signaling. BMC Cancer. 2021;21(1):396. doi:10.1186/s12885-021-08158-z.
  • Kashafi E, Moradzadeh M, Mohamadkhani A, Erfanian S. Kaempferol increases apoptosis in human cervical cancer HeLa cells via PI3K/AKT and telomerase pathways. Biomed Pharmacother. 2017;89:573–7.
  • Li Q, Wei L, Lin S, Chen Y, Lin J, Peng J. Synergistic effect of kaempferol and 5‑fluorouracil on the growth of colorectal cancer cells by regulating the PI3K/Akt signaling pathway. Mol Med Rep. 2019;20(1):728–34.
  • Song H, Bao J, Wei Y, Chen Y, Mao X, Li J, Yang Z, Xue Y. Kaempferol inhibits gastric cancer tumor growth: an in vitro and in vivo study. Oncol Rep. 2015;33(2):868–74. doi:10.3892/or.2014.3662.
  • 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. doi:10.1016/j.abb.2020.108700.
  • Luo H, Rankin GO, Juliano N, Jiang B-H, Chen YC. Kaempferol inhibits VEGF expression and in vitro angiogenesis through a novel ERK-NFκB-cMyc-p21 pathway. Food Chem. 2012;130(2):321–8.
  • Kuo W-T, Tsai Y-C, Wu H-C, Ho Y-J, Chen Y-S, Yao C-H, et al. Radiosensitization of non-small cell lung cancer by kaempferol. Oncol Rep. 2015;34(5):2351–6.
  • Kim S-H, Hwang K-A, Choi K-C. Treatment with kaempferol suppresses breast cancer cell growth caused by estrogen and triclosan in cellular and xenograft breast cancer models. J Nutr Biochem. 2016;28:70–82.
  • Li C, Zhao Y, Yang D, Yu Y, Guo H, Zhao Z, Zhang B, Yin X. Inhibitory effects of kaempferol on the invasion of human breast carcinoma cells by downregulating the expression and activity of matrix metalloproteinase-9. Biochem Cell Biol. 2015;93(1):16–27.
  • Yao J, Zhang B, Ge C, Peng S, Fang J. Xanthohumol, a polyphenol chalcone present in hops, activating Nrf2 enzymes to confer protection against oxidative damage in PC12 cells. J Agric Food Chem. 2015;63(5):1521–31. doi:10.1021/jf505075n.
  • Lv H, Liu Q, Wen Z, Feng H, Deng X, Ci X. Xanthohumol ameliorates lipopolysaccharide (LPS)-induced acute lung injury via induction of AMPK/GSK3β-Nrf2 signal axis. Redox Biol. 2017;12:311–24. doi:10.1016/j.redox.2017.03.001.
  • Liu X, Song Z, Bai J, Nauwynck H, Zhao Y, Jiang P. Xanthohumol inhibits PRRSV proliferation and alleviates oxidative stress induced by PRRSV via the Nrf2–HMOX1 axis. Vet Res. 2019;50(1):61. doi:10.1186/s13567-019-0679-2.
  • Li F, Yao Y, Huang H, Hao H, Ying M. Xanthohumol attenuates cisplatin-induced nephrotoxicity through inhibiting NF-κB and activating Nrf2 signaling pathways. Int Immunopharmacol. 2018;61:277–82. doi:10.1016/j.intimp.2018.05.017.
  • Bellezza I, Mierla AL, Minelli A. Nrf2 and NF-κB and their concerted modulation in cancer pathogenesis and progression. Cancers. 2010;2(2):483–97. 10.3390/cancers2020483
  • Zhou Y, Jiang Z, Lu H, Xu Z, Tong R, Shi J, Jia G. Recent advances of natural polyphenols activators for Keap1‐Nrf2 signaling pathway. Chem Biodivers. 2019;16(11):e1900400. doi:10.1002/cbdv.201900400.
  • Vakana E, Platanias LC. AMPK in BCR-ABL expressing leukemias. Regulatory effects and therapeutic implications. Oncotarget. 2011;2(12):1322.
  • Gao X, Deeb D, Liu Y, Gautam S, Dulchavsky SA, Gautam SC. Immunomodulatory activity of xanthohumol: inhibition of T cell proliferation, cell-mediated cytotoxicity and Th1 cytokine production through suppression of NF-κB. Immunopharmacol Immunotoxicol. 2009;31(3):477–84. doi:10.1080/08923970902798132.
  • Lee I-S, Lim J, Gal J, Kang JC, Kim HJ, Kang BY, Choi HJ. Anti-inflammatory activity of xanthohumol involves heme oxygenase-1 induction via NRF2-ARE signaling in microglial BV2 cells. Neurochem Int. 2011;58(2):153–60. doi:10.1016/j.neuint.2010.11.008.
  • Zajc I, Filipič M, Lah T. Xanthohumol induces different cytotoxicity and apoptotic pathways in malignant and normal astrocytes. Phytother Res. 2012;26(11):1709–13. doi:10.1002/ptr.4636.
  • Lee SH, Kim HJ, Lee JS, Lee I-S, Kang BY. Inhibition of topoisomerase I activity and efflux drug transporters’ expression by xanthohumol from hops. Arch Pharm Res. 2007;30(11):1435–9. doi:10.1007/BF02977368.
  • Aloqbi A, Omar U, Yousr M, Grace M, Lila MA, Howell N. Antioxidant activity of pomegranate juice and punicalagin. Nat Sci. 2016;8(06):235. doi:10.4236/ns.2016.86028.
  • Lin C-C, Hsu Y-F, Lin T-C. Effects of punicalagin and punicalin on carrageenan-induced inflammation in rats. Am J Chin Med. 1999;27(3–4):371–6. doi:10.1142/S0192415X99000422.
  • Quirós-Fernández R, López-Plaza B, Bermejo LM, Palma-Milla S, Gómez-Candela C. Supplementation with hydroxytyrosol and punicalagin improves early atherosclerosis markers involved in the asymptomatic phase of atherosclerosis in the adult population: a randomized, placebo-controlled, crossover trial. Nutrients. 2019;11(3):640. 10.3390/nu11030640
  • Xu Y, Shi C, Wu Q, Zheng Z, Liu P, Li G, Peng X, Xia X. Antimicrobial activity of punicalagin against Staphylococcus aureus and its effect on biofilm formation. Foodborne Pathog Dis. 2017;14(5):282–7. doi:10.1089/fpd.2016.2226.
  • Li G, Yan C, Xu Y, Feng Y, Wu Q, Lv X, Yang B, Wang X, Xia X. Punicalagin inhibits Salmonella virulence factors and has anti-quorum-sensing potential. Appl Environ Microbiol. 2014;80(19):6204–11. doi:10.1128/AEM.01458-14.
  • Aqil F, Munagala R, Vadhanam MV, Kausar H, Jeyabalan J, Schultz DJ, Gupta RC. Anti-proliferative activity and protection against oxidative DNA damage by punicalagin isolated from pomegranate husk. Food Res Int. 2012;49(1):345–53. doi:10.1016/j.foodres.2012.07.059.
  • Xu X, Li H, Hou X, Li D, He S, Wan C, et al. Punicalagin induces Nrf2/HO-1 expression via upregulation of PI3K/AKT pathway and inhibits LPS-induced oxidative stress in RAW264. 7 macrophages. Mediat Inflamm. 2015;2015:380218. doi:10.1155/2015/380218.
  • Xu L, He S, Yin P, Li D, Mei C, Yu X, Shi Y, Jiang L, Liu F. Punicalagin induces Nrf2 translocation and HO-1 expression via PI3K/Akt, protecting rat intestinal epithelial cells from oxidative stress. Int J Hyperthermia. 2016;32(5):465–73. doi:10.3109/02656736.2016.1155762.
  • Zahin M, Ahmad I, Gupta RC, Aqil F. Punicalagin and ellagic acid demonstrate antimutagenic activity and inhibition of benzo [a] pyrene induced DNA adducts. Biomed Res Int. 2014;2014:467465. doi:10.1155/2014/467465.
  • Wardyn JD, Ponsford AH, Sanderson CM. Dissecting molecular cross-talk between Nrf2 and NF-κB response pathways. Biochem Soc Trans. 2015;43(4):621–6. doi:10.1042/BST20150014.
  • Xu X, Yin P, Wan C, Chong X, Liu M, Cheng P, et al. Punicalagin inhibits inflammation in LPS-induced RAW264. 7 macrophages via the suppression of TLR4-mediated MAPKs and NF-κB activation. Inflammation. 2014;37(3):956–65.
  • Rubiolo JA, Mithieux G, Vega FV. Resveratrol protects primary rat hepatocytes against oxidative stress damage:: activation of the Nrf2 transcription factor and augmented activities of antioxidant enzymes. Eur J Pharmacol. 2008;591(1–3):66–72. doi:10.1016/j.ejphar.2008.06.067.
  • García-García J, Micol V, de Godos A, Gómez-Fernández JC. The cancer chemopreventive agent resveratrol is incorporated into model membranes and inhibits protein kinase C α activity. Arch Biochem Biophys. 1999;372(2):382–8. doi:10.1006/abbi.1999.1507.
  • Zhuang Y, Wu H, Wang X, He J, He S, Yin Y. Resveratrol attenuates oxidative stress-induced intestinal barrier injury through PI3K/Akt-mediated Nrf2 signaling pathway. Oxid Med Cell Longev. 2019;2019:7591840. doi:10.1155/2019/7591840.
  • Hsieh T-c, Lu X, Wang Z, Wu JM. Induction of quinone reductase NQO1 by resveratrol in human K562 cells involves the antioxidant response element ARE and is accompanied by nuclear translocation of transcription factor Nrf2. Med Chem. 2006;2(3):275–85. doi:10.2174/157340606776930709.
  • Lu F, Zahid M, Wang C, Saeed M, Cavalieri EL, Rogan EG. Resveratrol prevents estrogen-DNA adduct formation and neoplastic transformation in MCF-10F cells. Cancer Prev Res (Phila). 2008;1(2):135–45. doi:10.1158/1940-6207.CAPR-08-0037.
  • Zhang Y, Wang G, Wang T, Cao W, Zhang L, Chen X. Nrf2–Keap1 pathway–mediated effects of resveratrol on oxidative stress and apoptosis in hydrogen peroxide–treated rheumatoid arthritis fibroblast‐like synoviocytes. Ann N Y Acad Sci. 2019;1457(1):166–78. doi:10.1111/nyas.14196.
  • Roy P, Madan E, Kalra N, Nigam N, George J, Ray RS, Hans RK, Prasad S, Shukla Y. Resveratrol enhances ultraviolet B-induced cell death through nuclear factor-κB pathway in human epidermoid carcinoma A431 cells. Biochem Biophys Res Commun. 2009;384(2):215–20. doi:10.1016/j.bbrc.2009.04.100.
  • Chang J, Zhang Y, Li Y, Lu K, Shen Y, Guo Y, Qi Q, Wang M, Zhang S. NrF2/ARE and NF-κB pathway regulation may be the mechanism for lutein inhibition of human breast cancer cell. Future Oncol. 2018;14(8):719–26. doi:10.2217/fon-2017-0584.
  • Wu Y, Zhou BP. Inflammation: a driving force speeds cancer metastasis. Cell Cycle. 2009;8(20):3267–73. doi:10.4161/cc.8.20.9699.
  • Berman AY, Motechin RA, Wiesenfeld MY, Holz MK. The therapeutic potential of resveratrol: a review of clinical trials. npj Precis Oncol. 2017;1(1):1–9.
  • Tan KH, Nishida R. Methyl eugenol: its occurrence, distribution, and role in nature, especially in relation to insect behavior and pollination. J Insect Sci. 2012;12(1):56. doi:10.1673/031.012.5601.
  • Choi YK, Cho G-S, Hwang S, Kim BW, Lim JH, Lee J-C, Kim HC, Kim W-K, Kim YS. Methyleugenol reduces cerebral ischemic injury by suppression of oxidative injury and inflammation. Free Radic Res. 2010;44(8):925–35. doi:10.3109/10715762.2010.490837.
  • Shin B-K, Lee E-H, Kim H-M. Suppression ofL-histidine decarboxylase mRNA expression by methyleugenol. Biochem Biophys Res Commun. 1997;232(1):188–91. doi:10.1006/bbrc.1997.6260.
  • Zhou J, Ma X, Cui Y, Song Y, Yao L, Liu Y, Li S. Methyleugenol protects against t-BHP-triggered oxidative injury by induction of Nrf2 dependent on AMPK/GSK3β and ERK activation. J Pharmacol Sci. 2017;135(2):55–63. doi:10.1016/j.jphs.2017.09.003.
  • Ma L, Liu J, Lin Q, Gu Y, Yu W. Eugenol protects cells against oxidative stress via Nrf2. Exp Ther Med. 2021;21(2):1. doi:10.3892/etm.2020.9539.
  • Chen H, Fu J, Chen H, Hu Y, Soroka DN, Prigge JR, Schmidt EE, Yan F, Major MB, Chen X, et al. Ginger compound [6]-shogaol and its cysteine-conjugated metabolite (M2) activate Nrf2 in colon epithelial cells in vitro and in vivo. Chem Res Toxicol. 2014;27(9):1575–85. doi:10.1021/tx500211x.
  • Kim J-K, Jang H-D. 6-shogaol attenuates H2O2-induced oxidative stress via upregulation of Nrf2-mediated γ-glutamylcysteine synthetase and heme oxygenase expression in HepG2 cells. Food Sci Biotechnol. 2016;25(1):319–27. doi:10.1007/s10068-016-0045-3.
  • Hur J, Lee Y, Lee CJ, Park H-Y, Choi SY. 6-shogaol suppresses oxidative damage in L6 muscle cells. Appl Biol Chem. 2020;63(1):1–6. doi:10.1186/s13765-020-00544-8.
  • Peng S, Yao J, Liu Y, Duan D, Zhang X, Fang J. Activation of Nrf2 target enzymes conferring protection against oxidative stress in PC12 cells by ginger principal constituent 6-shogaol. Food Funct. 2015;6(8):2813–23. doi:10.1039/c5fo00214a.
  • Gan F-F, Ling H, Ang X, Reddy SA, Lee SS-H, Yang H, Tan S-H, Hayes JD, Chui W-K, Chew E-H, et al. A novel shogaol analog suppresses cancer cell invasion and inflammation, and displays cytoprotective effects through modulation of NF-κB and Nrf2-Keap1 signaling pathways. Toxicol Appl Pharmacol. 2013;272(3):852–62. doi:10.1016/j.taap.2013.07.011.
  • Kim MO, Lee M-H, Oi N, Kim S-H, Bae KB, Huang Z, Kim DJ, Reddy K, Lee S-Y, Park SJ, et al. [6]-Shogaol inhibits growth and induces apoptosis of non-small cell lung cancer cells by directly regulating Akt1/2. Carcinogenesis. 2014;35(3):683–91. doi:10.1093/carcin/bgt365.
  • Chen F, Tang Y, Sun Y, Veeraraghavan VP, Mohan SK, Cui C. 6-shogaol, a active constiuents of ginger prevents UVB radiation mediated inflammation and oxidative stress through modulating NrF2 signaling in human epidermal keratinocytes (HaCaT cells). J Photochem Photobiol B. 2019;197:111518. doi:10.1016/j.jphotobiol.2019.111518.
  • Bischoff-Kont I, Fürst R. Benefits of ginger and its constituent 6-shogaol in inhibiting inflammatory processes. Pharmaceuticals (Basel). 2021;14(6):571. doi:10.3390/ph14060571.
  • Du Y-T, Zheng Y-L, Ji Y, Dai F, Hu Y-J, Zhou B. Applying an electrophilicity-based strategy to develop a novel nrf2 activator inspired from dietary [6]-shogaol. J Agric Food Chem. 2018;66(30):7983–94. doi:10.1021/acs.jafc.8b02442.
  • Yun N, Kang JW, Lee SM. Protective effects of chlorogenic acid against ischemia/reperfusion injury in rat liver: molecular evidence of its antioxidant and anti-inflammatory properties. J Nutr Biochem. 2012;23(10):1249–55. doi:10.1016/j.jnutbio.2011.06.018.
  • Liang N, Dupuis JH, Yada RY, Kitts DD. Chlorogenic acid isomers directly interact with Keap 1-Nrf2 signaling in Caco-2 cells. Mol Cell Biochem. 2019;457(1–2):105–18. doi:10.1007/s11010-019-03516-9.
  • Shi A, Shi H, Wang Y, Liu X, Cheng Y, Li H, Zhao H, Wang S, Dong L. Activation of Nrf2 pathway and inhibition of NLRP3 inflammasome activation contribute to the protective effect of chlorogenic acid on acute liver injury. Int Immunopharmacol. 2018;54:125–30. doi:10.1016/j.intimp.2017.11.007.
  • Bao L, Li J, Zha D, Zhang L, Gao P, Yao T, Wu X. Chlorogenic acid prevents diabetic nephropathy by inhibiting oxidative stress and inflammation through modulation of the Nrf2/HO-1 and NF-ĸB pathways. Int Immunopharmacol. 2018;54:245–53. doi:10.1016/j.intimp.2017.11.021.
  • Zhao X-L, Yu L, Zhang S-D, Ping K, Ni H-Y, Qin X-Y, Zhao C-J, Wang W, Efferth T, Fu Y-J, et al. Cryptochlorogenic acid attenuates LPS-induced inflammatory response and oxidative stress via upregulation of the Nrf2/HO-1 signaling pathway in RAW 264.7 macrophages. Int Immunopharmacol. 2020;83:106436. doi:10.1016/j.intimp.2020.106436.
  • Bender O, Atalay A. Polyphenol chlorogenic acid, antioxidant profile, and breast cancer. In: Preedy VR, Patel VB, editors. Cancer. Academic Press; 2021. p. 311–21. doi: 10.1016/B978-0-12-819547-5.00028-6.
  • Reczek CR, Chandel NS. The two faces of reactive oxygen species in cancer. Annu Rev Cancer Biol. 2017;1:79–98. doi:10.1146/annurev-cancerbio-041916-065808.
  • Hayakawa S, Ohishi T, Miyoshi N, Oishi Y, Nakamura Y, Isemura M. Anti-cancer effects of green tea epigallocatchin-3-gallate and coffee chlorogenic acid. Molecules. 2020;25(19):4553. doi:10.3390/molecules25194553.
  • Park JJ, Hwang SJ, Park J-H, Lee H-J. Chlorogenic acid inhibits hypoxia-induced angiogenesis via down-regulation of the HIF-1α/AKT pathway. Cell Oncol. 2015;38(2):111–8.
  • Lampiasi N, Montana G. An in vitro inflammation model to study the Nrf2 and NF-κB crosstalk in presence of ferulic acid as modulator. Immunobiology. 2018;223(4–5):349–55. doi:10.1016/j.imbio.2017.10.046.
  • Chaudhary A, Jaswal VS, Choudhary S, Sonika, Sharma A, Beniwal V, Tuli HS, Sharma S. Ferulic acid: a promising therapeutic phytochemical and recent patents advances. Recent Pat Inflamm Allergy Drug Discov. 2019;13(2):115–23. doi:10.2174/1872213X13666190621125048.
  • Mahmoud AM, Hussein OE, Hozayen WG, Bin-Jumah M, Abd El-Twab SM. Ferulic acid prevents oxidative stress, inflammation, and liver injury via upregulation of Nrf2/HO-1 signaling in methotrexate-induced rats. Environ Sci Pollut Res Int. 2020;27(8):7910–21. doi:10.1007/s11356-019-07532-6.
  • Yu CL, Zhao XM, Niu YC. Ferulic acid protects against lead acetate-induced inhibition of neurite outgrowth by upregulating HO-1 in PC12 cells: involvement of ERK1/2-Nrf2 pathway. Mol Neurobiol. 2016;53(9):6489–500. doi:10.1007/s12035-015-9555-x.
  • Ma ZC, Hong Q, Wang YG, Liang QD, Tan HL, Xiao CR, Tang XL, Shao S, Zhou SS, Gao Y, et al. Ferulic acid induces heme oxygenase-1 via activation of ERK and Nrf2. Drug Discov Ther. 2011;5(6):299–305. doi:10.5582/ddt.2011.v5.6.299.
  • Das U, Manna K, Khan A, Sinha M, Biswas S, Sengupta A, Chakraborty A, Dey S. Ferulic acid (FA) abrogates γ-radiation induced oxidative stress and DNA damage by up-regulating nuclear translocation of Nrf2 and activation of NHEJ pathway. Free Radic Res. 2017;51(1):47–63. doi:10.1080/10715762.2016.1267345.
  • Olivier S, Robe P, Bours V. Can NF-kappaB be a target for novel and efficient anti-cancer agents? Biochem Pharmacol. 2006;72(9):1054–68. doi:10.1016/j.bcp.2006.07.023.
  • Lampiasi N, Montana G. The molecular events behind ferulic acid mediated modulation of IL-6 expression in LPS-activated Raw 264.7 cells. Immunobiology. 2016;221(3):486–93. doi:10.1016/j.imbio.2015.11.001.
  • Choi YE, Park E. Ferulic acid in combination with PARP inhibitor sensitizes breast cancer cells as chemotherapeutic strategy. Biochem Biophys Res Commun. 2015;458(3):520–4. doi:10.1016/j.bbrc.2015.01.147.
  • Kovacs K, Vaczy A, Fekete K, Kovari P, Atlasz T, Reglodi D, Gabriel R, Gallyas F, Sumegi B. PARP inhibitor protects against chronic hypoxia/reoxygenation-induced retinal injury by regulation of MAPKs, HIF1α, Nrf2, and NFκB. Invest Ophthalmol Vis Sci. 2019;60(5):1478–90. doi:10.1167/iovs.18-25936.
  • Serreli G, Naitza MR, Zodio S, Leoni VP, Spada M, Melis MP, et al. Ferulic acid metabolites attenuate LPS-induced inflammatory response in enterocyte-like cells. Nutrients. 2021;13(9):3152. doi:10.3390/nu13093152.
  • Satoh T, Kosaka K, Itoh K, Kobayashi A, Yamamoto M, Shimojo Y, Kitajima C, Cui J, Kamins J, Okamoto S-i, et al. Carnosic acid, a catechol-type electrophilic compound, protects neurons both in vitro and in vivo through activation of the Keap1/Nrf2 pathway via S-alkylation of targeted cysteines on Keap1. J Neurochem. 2008;104(4):1116–31. doi:10.1111/j.1471-4159.2007.05039.x.
  • de Oliveira MR, Ferreira GC, Schuck PF, Dal Bosco SM. Role for the PI3K/Akt/Nrf2 signaling pathway in the protective effects of carnosic acid against methylglyoxal-induced neurotoxicity in SH-SY5Y neuroblastoma cells. Chem Biol Interact. 2015;242:396–406. doi:10.1016/j.cbi.2015.11.003.
  • Xie Z, Zhong L, Wu Y, Wan X, Yang H, Xu X, Li P. Carnosic acid improves diabetic nephropathy by activating Nrf2/ARE and inhibition of NF-κB pathway. Phytomedicine. 2018;47:161–73. doi:10.1016/j.phymed.2018.04.031.
  • Lin Y, Bai L, Chen W, Xu S. The NF-kappaB activation pathways, emerging molecular targets for cancer prevention and therapy. Expert Opin Ther Targets. 2010;14(1):45–55. doi:10.1517/14728220903431069.
  • Khan MA, Jain VK, Rizwanullah M, Ahmad J, Jain K. PI3K/AKT/mTOR pathway inhibitors in triple-negative breast cancer: a review on drug discovery and future challenges. Drug Discov Today 2019;24(11):2181–91. doi:10.1016/j.drudis.2019.09.001
  • Ghoneum A, Said N. PI3K-AKT-mTOR and NFκB pathways in ovarian cancer: implications for targeted therapeutics. Cancers (Basel). 2019;11(7):949. doi:10.3390/cancers11070949.
  • Yang N, Xia Z, Shao N, Li B, Xue L, Peng Y, Zhi F, Yang Y. Carnosic acid prevents dextran sulfate sodium-induced acute colitis associated with the regulation of the Keap1/Nrf2 pathway. Sci Rep. 2017;7(1):11036. doi:10.1038/s41598-017-11408-5.
  • Gong J, Xie J, Bedolla R, Rivas P, Chakravarthy D, Freeman JW, Reddick R, Kopetz S, Peterson A, Wang H, et al. Combined targeting of STAT3/NF-κB/COX-2/EP4 for effective management of pancreatic cancer. Clin Cancer Res. 2014;20(5):1259–73. doi:10.1158/1078-0432.Ccr-13-1664.
  • Gao Q, Liu H, Yao Y, Geng L, Zhang X, Jiang L, Shi B, Yang F. Carnosic acid induces autophagic cell death through inhibition of the Akt/mTOR pathway in human hepatoma cells. J Appl Toxicol. 2015;35(5):485–92. doi:10.1002/jat.3049.
  • Einbond LS, Wu H-A, Kashiwazaki R, He K, Roller M, Su T, Wang X, Goldsberry S. Carnosic acid inhibits the growth of ER-negative human breast cancer cells and synergizes with curcumin. Fitoterapia. 2012;83(7):1160–8. doi:10.1016/j.fitote.2012.07.006.
  • Shi B, Wang LF, Meng WS, Chen L, Meng ZL. Carnosic acid and fisetin combination therapy enhances inhibition of lung cancer through apoptosis induction. Int J Oncol. 2017;50(6):2123–35. doi:10.3892/ijo.2017.3970.
  • Chen CC, Chen HL, Hsieh CW, Yang YL, Wung BS. Upregulation of NF-E2-related factor-2-dependent glutathione by carnosol provokes a cytoprotective response and enhances cell survival. Acta Pharmacol Sin. 2011;32(1):62–9. doi:10.1038/aps.2010.181.
  • Foresti R, Bucolo C, Platania CM, Drago F, Dubois-Randé JL, Motterlini R. Nrf2 activators modulate oxidative stress responses and bioenergetic profiles of human retinal epithelial cells cultured in normal or high glucose conditions. Pharmacol Res. 2015;99:296–307. doi:10.1016/j.phrs.2015.07.006.
  • Wu KC, McDonald PR, Liu J, Klaassen CD. Screening of natural compounds as activators of the keap1-nrf2 pathway. Planta Med. 2014;80(1):97–104. doi:10.1055/s-0033-1351097.
  • Martin D, Rojo AI, Salinas M, Diaz R, Gallardo G, Alam J, De Galarreta CMR, Cuadrado A. Regulation of heme oxygenase-1 expression through the phosphatidylinositol 3-kinase/Akt pathway and the Nrf2 transcription factor in response to the antioxidant phytochemical carnosol. J Biol Chem. 2004;279(10):8919–29. doi:10.1074/jbc.M309660200.
  • Vergara D, Simeone P, Bettini S, Tinelli A, Valli L, Storelli C, Leo S, Santino A, Maffia M. Antitumor activity of the dietary diterpene carnosol against a panel of human cancer cell lines. Food Funct. 2014;5(6):1261–9. doi:10.1039/c4fo00023d.
  • Park K-W, Kundu J, Chae I-G, Kim D-H, Yu M-H, Kundu JK, Chun K-S. Carnosol induces apoptosis through generation of ROS and inactivation of STAT3 signaling in human colon cancer HCT116 cells. Int J Oncol. 2014;44(4):1309–15. doi:10.3892/ijo.2014.2281.
  • Sanli T, Linher-Melville K, Tsakiridis T, Singh G. Sestrin2 modulates AMPK subunit expression and its response to ionizing radiation in breast cancer cells. PLoS One. 2012;7(2):e32035. doi:10.1371/journal.pone.0032035.
  • O’Neill EJ, Hartogh D, Azizi DJ, Tsiani K. E. Anticancer properties of carnosol: a summary of in vitro and in vivo evidence. Antioxidants. 2020;9(10):961.
  • Ding Y, Zhang B, Zhou K, Chen M, Wang M, Jia Y, Song Y, Li Y, Wen A. Dietary ellagic acid improves oxidant-induced endothelial dysfunction and atherosclerosis: role of Nrf2 activation. Int J Cardiol. 2014;175(3):508–14. doi:10.1016/j.ijcard.2014.06.045.
  • Ding X, Jian T, Wu Y, Zuo Y, Li J, Lv H, Ma L, Ren B, Zhao L, Li W, et al. Ellagic acid ameliorates oxidative stress and insulin resistance in high glucose-treated HepG2 cells via miR-223/keap1-Nrf2 pathway. Biomed Pharmacother. 2019;110:85–94. doi:10.1016/j.biopha.2018.11.018.
  • ALTamimi JZ, AlFaris NA, Aljabryn DH, Alagal RI, Alshammari GM, Aldera H, et al. Ellagic acid improved diabetes mellitus-induced testicular damage and sperm abnormalities by activation of Nrf2. Saudi J Biol Sci. 2021;28(8):4300–4310. doi:10.1016/j.sjbs.2021.04.005.
  • Yang H-L, Lin C-P, Vudhya Gowrisankar Y, Huang P-J, Chang W-L, Shrestha S, Hseu Y-C. The anti-melanogenic effects of ellagic acid through induction of autophagy in melanocytes and suppression of UVA-activated α-MSH pathways via Nrf2 activation in keratinocytes. Biochem Pharmacol. 2021;185:114454. doi:10.1016/j.bcp.2021.114454.
  • Edderkaoui M, Odinokova I, Ohno I, Gukovsky I, Go VLW, Pandol SJ, Gukovskaya AS. Ellagic acid induces apoptosis through inhibition of nuclear factor kappa B in pancreatic cancer cells. World J Gastroenterol. 2008;14(23):3672–80. doi:10.3748/wjg.14.3672.
  • Ebrahimi R, Sepand MR, Seyednejad SA, Omidi A, Akbariani M, Gholami M, Sabzevari O. Ellagic acid reduces methotrexate-induced apoptosis and mitochondrial dysfunction via up-regulating Nrf2 expression and inhibiting the IĸBα/NFĸB in rats. Daru. 2019;27(2):721–33. doi:10.1007/s40199-019-00309-9.
  • Ceci C, Tentori L, Atzori MG, Lacal PM, Bonanno E, Scimeca M, et al. Ellagic acid inhibits bladder cancer invasiveness and in vivo tumor growth. Nutrients. 2016;8(11):744.
  • Zhu H, Jin H, Pi J, Bai H, Yang F, Wu C, et al. Apigenin induced apoptosis in esophageal carcinoma cells by destruction membrane structures. Scanning. 2016;38(4):322–8.
  • Huang C-S, Lii C-K, Lin A-H, Yeh Y-W, Yao H-T, Li C-C, Wang T-S, Chen H-W. Protection by chrysin, apigenin, and luteolin against oxidative stress is mediated by the Nrf2-dependent up-regulation of heme oxygenase 1 and glutamate cysteine ligase in rat primary hepatocytes. Arch Toxicol. 2013;87(1):167–78. doi:10.1007/s00204-012-0913-4.
  • Paredes-Gonzalez X, Fuentes F, Jeffery S, Saw CL-L, Shu L, Su Z-Y, Kong A-NT. Induction of NRF2‐mediated gene expression by dietary phytochemical flavones apigenin and luteolin. Biopharm Drug Dispos. 2015;36(7):440–51. doi:10.1002/bdd.1956.
  • Zhang B, Wang J, Zhao G, Lin M, Lang Y, Zhang D, Feng D, Tu C. Apigenin protects human melanocytes against oxidative damage by activation of the Nrf2 pathway. Cell Stress Chaperones. 2020;25(2):277–85. doi:10.1007/s12192-020-01071-7.
  • Ozbey U, Attar R, Romero MA, Alhewairini SS, Afshar B, Sabitaliyevich UY, Hanna-Wakim L, Ozcelik B, Farooqi AA. Apigenin as an effective anticancer natural product: spotlight on TRAIL, WNT/β‐catenin, JAK‐STAT pathways, and microRNAs. J Cell Biochem. 2019;120(2):1060–7. doi:10.1002/jcb.27575.
  • Sharma A, Ghani A, Sak K, Tuli HS, Sharma AK, Setzer WN, Sharma S, Das AK. Probing into therapeutic anti-cancer potential of apigenin: recent trends and future directions. Recent Pat Inflamm Allergy Drug Discov. 2019;13(2):124–33. doi:10.2174/1872213X13666190816160240.
  • Javed Z, Sadia H, Iqbal MJ, Shamas S, Malik K, Ahmed R, et al. Apigenin role as cell-signaling pathways modulator: implications in cancer prevention and treatment. Cancer Cell Int. 2021;21(1):1–11. doi:10.1186/s12935-021-01888-x.
  • Soni RP, Katoch M, Kumar A, Ladohiya R, Verma P. Tea: production, composition, consumption and its potential as an antioxidant and antimicrobial agent. Int J Food Ferment Technol. 2015;5(2):95–106. doi:10.5958/2277-9396.2016.00002.7.
  • Jing X, Zhang J, Huang Z, Sheng Y, Ji L. The involvement of Nrf2 antioxidant signalling pathway in the protection of monocrotaline-induced hepatic sinusoidal obstruction syndrome in rats by (+)-catechin hydrate. Free Radic Res. 2018;52(4):402–14. doi:10.1080/10715762.2018.1437914.
  • Fan F-Y, Sang L-X, Jiang M. Catechins and their therapeutic benefits to inflammatory bowel disease. Molecules. 2017;22(3):484.
  • Suganuma M, Saha A, Fujiki H. New cancer treatment strategy using combination of green tea catechins and anticancer drugs. Cancer Sci. 2011;102(2):317–23. doi:10.1111/j.1349-7006.2010.01805.x.
  • Cheng Y-T, Wu C-H, Ho C-Y, Yen G-C. Catechin protects against ketoprofen-induced oxidative damage of the gastric mucosa by up-regulating Nrf2 in vitro and in vivo. J Nutr Biochem. 2013;24(2):475–83. doi:10.1016/j.jnutbio.2012.01.010.
  • Shay J, Elbaz HA, Lee I, Zielske SP, Malek MH, Hüttemann M. Molecular mechanisms and therapeutic effects of (−)-epicatechin and other polyphenols in cancer, inflammation, diabetes, and neurodegeneration. Oxid Med Cell Longev. 2015;2015:181260. doi:10.1155/2015/181260.
  • Chiou Y-S, Huang Q, Ho C-T, Wang Y-J, Pan M-H. Directly interact with Keap1 and LPS is involved in the anti-inflammatory mechanisms of (-)-epicatechin-3-gallate in LPS-induced macrophages and endotoxemia. Free Radic Biol Med. 2016;94:1–16. doi:10.1016/j.freeradbiomed.2016.02.010.
  • Bahia PK, Rattray M, Williams RJ. Dietary flavonoid (−) epicatechin stimulates phosphatidylinositol 3‐kinase‐dependent anti‐oxidant response element activity and up‐regulates glutathione in cortical astrocytes. J Neurochem. 2008;106(5):2194–204.doi:10.1111/j.1471-4159.2008.05542.x.
  • Granado-Serrano AB, Martín MA, Haegeman G, Goya L, Bravo L, Ramos S. Epicatechin induces NF-κB, activator protein-1 (AP-1) and nuclear transcription factor erythroid 2p45-related factor-2 (Nrf2) via phosphatidylinositol-3-kinase/protein kinase B (PI3K/AKT) and extracellular regulated kinase (ERK) signalling in HepG2 cells. Br J Nutr. 2010;103(2):168–79. doi:10.1017/S0007114509991747.
  • Mayer IA, Arteaga CL. The PI3K/AKT pathway as a target for cancer treatment. Annu Rev Med. 2016;67:11–28. doi:10.1146/annurev-med-062913-051343.
  • Wu C, Hsu M, Hsieh C, Lin J, Lai P, Wung B. Upregulation of heme oxygenase-1 by Epigallocatechin-3-gallate via the phosphatidylinositol 3-kinase/Akt and ERK pathways. Life Sci. 2006;78(25):2889–97. doi:10.1016/j.lfs.2005.11.013.
  • Pearson G, Robinson F, Beers Gibson T, Xu BE, Karandikar M, Berman K, Cobb MH. Mitogen-activated protein (MAP) kinase pathways: regulation and physiological functions. Endocr Rev. 2001;22(2):153–83. doi:10.1210/edrv.22.2.0428.
  • Andreadi CK, Howells LM, Atherfold PA, Manson MM. Involvement of Nrf2, p38, B-Raf, and nuclear factor-κB, but not phosphatidylinositol 3-kinase, in induction of hemeoxygenase-1 by dietary polyphenols. Mol Pharmacol. 2006;69(3):1033–40. doi:10.1124/mol.105.018374.
  • Shimizu M, Shirakami Y, Moriwaki H. Targeting receptor tyrosine kinases for chemoprevention by green tea catechin, EGCG. Int J Mol Sci. 2008;9(6):1034–49. doi:10.3390/ijms9061034.
  • Han SG, Han S-S, Toborek M, Hennig B. EGCG protects endothelial cells against PCB 126-induced inflammation through inhibition of AhR and induction of Nrf2-regulated genes. Toxicol Appl Pharmacol. 2012;261(2):181–8. doi:10.1016/j.taap.2012.03.024.
  • Na H-K, Surh Y-J. Modulation of Nrf2-mediated antioxidant and detoxifying enzyme induction by the green tea polyphenol EGCG. Food Chem Toxicol. 2008;46(4):1271–8. doi:10.1016/j.fct.2007.10.006.
  • Chen C, Yu R, Owuor ED, Kong A-NT. Activation of antioxidant-response element (ARE), mitogen-activated protein kinases (MAPKs) and caspases by major green tea polyphenol components during cell survival and death. Arch Pharm Res. 2000;23(6):605–12. doi:10.1007/BF02975249.
  • Hwang J-T, Ha J, Park I-J, Lee S-K, Baik HW, Kim YM, Park OJ. Apoptotic effect of EGCG in HT-29 colon cancer cells via AMPK signal pathway. Cancer Lett. 2007;247(1):115–21. doi:10.1016/j.canlet.2006.03.030.
  • Wada Y, Takata A, Ikemoto T, Morine Y, Imura S, Iwahashi S, Saito Y, Shimada M. The protective effect of epigallocatechin 3-gallate on mouse pancreatic islets via the Nrf2 pathway. Surg Today. 2019;49(6):536–45. doi:10.1007/s00595-019-1761-0.
  • Shankar S, Ganapathy S, Hingorani SR, Srivastava RK. EGCG inhibits growth, invasion, angiogenesis and metastasis of pancreatic cancer. Front Biosci. 2008;13(1):440–52. doi:10.2741/2691.
  • Zhang W, Zhang W, Sun L, Xiang L, Lai X, Li Q, et al. The effects and mechanisms of epigallocatechin-3-gallate on reversing multidrug resistance in cancer. Trends Food Sci Technol. 2019;93:221–33. doi:10.1016/j.tifs.2019.09.017.
  • Li M-J, Yin Y-C, Wang J, Jiang Y-F. Green tea compounds in breast cancer prevention and treatment. World J Clin Oncol. 2014;5(3):520–8. doi:10.5306/wjco.v5.i3.520.
  • Huang Y-J, Wang K-L, Chen H-Y, Chiang Y-F, Hsia S-M. Protective effects of epigallocatechin gallate (EGCG) on endometrial, breast, and ovarian cancers. Biomolecules. 2020;10(11):1481. doi:10.3390/biom10111481.
  • Higa S, Hirano T, Kotani M, Matsumoto M, Fujita A, Suemura M, Kawase I, Tanaka T. Fisetin, a flavonol, inhibits TH2-type cytokine production by activated human basophils. J Allergy Clin Immunol. 2003;111(6):1299–306. doi:10.1067/mai.2003.1456.
  • Zhang H, Zheng W, Feng X, Yang F, Qin H, Wu S, et al. Nrf2–ARE signaling acts as master pathway for the cellular antioxidant activity of fisetin. Molecules. 2019;24(4):708.
  • Li Y, Liu Y, Chen J, Hu J. Protective effect of Fisetin on the lipopolysaccharide-induced preeclampsia-like rats. Hyperten Pregnancy. 2022;2021:1–8. doi:10.1080/10641955.2021.2013874.
  • Ahmad S, Khan A, Ali W, Jo MH, Park J, Ikram M, et al. Fisetin rescues the mice brains against D-galactose-induced oxidative stress, neuroinflammation and memory impairment. Front Pharmacol. 2021;12:612078. doi:10.3389/fphar.2021.612078.
  • Lee SE, Jeong SI, Yang H, Park CS, Jin YH, Park YS. Fisetin induces Nrf2‐mediated HO‐1 expression through PKC‐δ and p38 in human umbilical vein endothelial cells. J Cell Biochem. 2011;112(9):2352–60. doi:10.1002/jcb.23158.
  • Yen J-H, Wu P-S, Chen S-F, Wu M-J. Fisetin protects PC12 cells from tunicamycin-mediated cell death via reactive oxygen species scavenging and modulation of Nrf2-driven gene expression, SIRT1 and MAPK signaling in PC12 cells. Int J Mol Sci. 2017;18(4):852. doi:10.3390/ijms18040852.
  • Sun Y, Liu W-Z, Liu T, Feng X, Yang N, Zhou H-F. Signaling pathway of MAPK/ERK in cell proliferation, differentiation, migration, senescence and apoptosis. J Recept Signal Transduct Res. 2015;35(6):600–4. doi:10.3109/10799893.2015.1030412.
  • Yang P-M, Tseng H-H, Peng C-W, Chen W-S, Chiu S-J. Dietary flavonoid fisetin targets caspase-3-deficient human breast cancer MCF-7 cells by induction of caspase-7-associated apoptosis and inhibition of autophagy. Int J Oncol. 2012;40(2):469–78. doi:10.3892/ijo.2011.1203.
  • Sundarraj K, Raghunath A, Panneerselvam L, Perumal E. Fisetin inhibits autophagy in HepG2 cells via PI3K/Akt/mTOR and AMPK pathway. Nutr Cancer. 2021;73(11–12):2502–14. doi:10.1080/01635581.2020.1836241.
  • Hada Y, Uchida HA, Wada J. Fisetin attenuates lipopolysaccharide-induced inflammatory responses in macrophage. Biomed Res Int. 2021;2021:5570885. doi:10.1155/2021/5570885.
  • Molagoda IMN, Athapaththu A, Choi YH, Park C, Jin CY, Kang CH, et al. Fisetin inhibits NLRP3 inflammasome by suppressing TLR4/MD2-mediated mitochondrial ROS production. Antioxidants (Basel, Switzerland). 2021;10(8):1215. doi:10.3390/antiox10081215.
  • Chenxu G, Xianling D, Qin K, Linfeng H, Yan S, Mingxin X, Jun T, Minxuan X. Fisetin protects against high fat diet-induced nephropathy by inhibiting inflammation and oxidative stress via the blockage of iRhom2/NF-κB signaling. Int Immunopharmacol. 2021;92:107353. doi:10.1016/j.intimp.2020.107353.
  • Pu J-L, Huang Z-T, Luo Y-H, Mou T, Li T-T, Li Z-T, Wei X-F, Wu Z-J. Fisetin mitigates hepatic ischemia-reperfusion injury by regulating GSK3β/AMPK/NLRP3 inflammasome pathway. Hepatobiliary Pancreat Dis Int. 2021;20(4):352–60. doi:10.1016/j.hbpd.2021.04.013.
  • Sim H, Choo S, Kim J, Baek M-C, Bae J-S. Fisetin suppresses pulmonary inflammatory responses through heme oxygenase-1 mediated downregulation of inducible nitric oxide synthase. J Med Food. 2020;23(11):1163–8. doi:10.1089/jmf.2020.4755.
  • Peng H-L, Huang W-C, Cheng S-C, Liou C-J. Fisetin inhibits the generation of inflammatory mediators in interleukin-1β–induced human lung epithelial cells by suppressing the NF-κB and ERK1/2 pathways. Int Immunopharmacol. 2018;60:202–10. doi:10.1016/j.intimp.2018.05.004.
  • Bhat TA, Nambiar D, Pal A, Agarwal R, Singh RP. Fisetin inhibits various attributes of angiogenesis in vitro and in vivo—implications for angioprevention. Carcinogenesis. 2012;33(2):385–93.
  • Li L, Wang M, Yang H, Li Y, Huang X, Guo J, et al. Fisetin inhibits trypsin activity and suppresses the growth of colorectal cancer in vitro and in vivo. Nat Prod Commun. 2022;17(8). doi:10.1177/1934578X221115511.
  • Sarkar FH, Li Y. Mechanisms of cancer chemoprevention by soy isoflavone genistein. Cancer Metastasis Rev. 2002;21(3–4):265–80. doi:10.1023/a:1021210910821.
  • Zhang T, Wang F, Xu H-X, Yi L, Qin Y, Chang H, et al. Activation of nuclear factor erythroid 2-related factor 2 and PPARγ plays a role in the genistein-mediated attenuation of oxidative stress-induced endothelial cell injury. Br J Nutr. 2013;109(2):223–35.
  • Zhai X, Lin M, Zhang F, Hu Y, Xu X, Li Y, Liu K, Ma X, Tian X, Yao J, et al. Dietary flavonoid genistein induces Nrf2 and phase II detoxification gene expression via ERKs and PKC pathways and protects against oxidative stress in Caco‐2 cells. Mol Nutr Food Res. 2013;57(2):249–59. doi:10.1002/mnfr.201200536.
  • Xi Y-D, Yu H-L, Ding J, Ma W-W, Yuan L-H, Feng J-F, Xiao Y-X, Xiao R. Flavonoids protect cerebrovascular endothelial cells through Nrf2 and PI3K from β-amyloid peptide-induced oxidative damage. Curr Neurovasc Res. 2012;9(1):32–41. doi:10.2174/156720212799297092.
  • Noorolyai S, Shajari N, Baghbani E, Sadreddini S, Baradaran B. The relation between PI3K/AKT signalling pathway and cancer. Gene. 2019;698:120–8. 10.1016/j.gene.2019.02.076
  • Liu F-C, Wang C-C, Lu J-W, Lee C-H, Chen S-C, Ho Y-J, et al. Chondroprotective effects of genistein against osteoarthritis induced joint inflammation. Nutrients. 2019;11(5):1180.
  • Wang L, Li A, Liu Y, Zhan S, Zhong L, Du Y, Xu D, Wang W, Huang W. Genistein protects against acetaminophen-induced liver toxicity through augmentation of SIRT1 with induction of Nrf2 signalling. Biochem Biophys Res Commun. 2020;527(1):90–7. doi:10.1016/j.bbrc.2020.04.100.
  • Kim E-J, Um S-J. SIRT1: roles in aging and cancer. BMB Rep. 2008;41(11):751–6.
  • Zhang L, Li H, Gao M, Zhang T, Wu Z, Wang Z, Chong T. Genistein attenuates di‑(2‑ethylhexyl) phthalate-induced testicular injuries via activation of Nrf2/HO‑1 following prepubertal exposure. Int J Mol Med. 2018;41(3):1437–46. doi:10.3892/ijmm.2018.3371.
  • Gundogdu G, Dodurga Y, Elmas L, Tasci SY, Karaoglan ES. Investigation of the anticancer mechanism of isoorientin isolated from Eremurus spectabilis leaves via cell cycle pathways in HT-29 human colorectal adenocarcinoma cells. Eurasian J Med. 2018;50(3):168. doi:10.5152/eurasianjmed.2018.17403.
  • Ma L, Zhang B, Liu J, Qiao C, Liu Y, Li S, Lv H. Isoorientin exerts a protective effect against 6-OHDA-induced neurotoxicity by activating the AMPK/AKT/Nrf2 signalling pathway. Food Funct. 2020;11(12):10774–85. doi:10.1039/d0fo02165b.
  • Yuan L, Wang J, Wu W, Liu Q, Liu X. Effect of isoorientin on intracellular antioxidant defence mechanisms in hepatoma and liver cell lines. Biomed Pharmacother. 2016;81:356–62. doi:10.1016/j.biopha.2016.04.025.
  • Yuan L, Wang J, Xiao H, Wu W, Wang Y, Liu X. MAPK signaling pathways regulate mitochondrial-mediated apoptosis induced by isoorientin in human hepatoblastoma cancer cells. Food Chem Toxicol. 2013;53:62–8. doi:10.1016/j.fct.2012.11.048.
  • Anilkumar K, Reddy GV, Azad R, Yarla NS, Dharmapuri G, Srivastava A, et al. Evaluation of anti-inflammatory properties of isoorientin isolated from tubers of Pueraria tuberosa. Oxid Med Cell Longev. 2017;2017:5498054. doi:10.1155/2017/5498054.
  • Li Y, Zhao Y, Tan X, Liu J, Zhi Y, Yi L, Bai S, Du Q, Li QX, Dong Y, et al. Isoorientin inhibits inflammation in macrophages and endotoxemia mice by regulating glycogen synthase kinase 3β. Mediators Inflamm. 2020;2020:8704146–11. doi:10.1155/2020/8704146.
  • Boots AW, Haenen GR, Bast A. Health effects of quercetin: from antioxidant to nutraceutical. Eur J Pharmacol. 2008;585(2–3):325–37. doi:10.1016/j.ejphar.2008.03.008.
  • Tanigawa S, Fujii M, Hou D-X. Action of Nrf2 and Keap1 in ARE-mediated NQO1 expression by quercetin. Free Radic Biol Med. 2007;42(11):1690–703. doi:10.1016/j.freeradbiomed.2007.02.017.
  • Granado-Serrano AB, Martín MA, Bravo L, Goya L, Ramos S. Quercetin modulates Nrf2 and glutathione-related defenses in HepG2 cells: involvement of p38. Chem Biol Interact. 2012;195(2):154–64. doi:10.1016/j.cbi.2011.12.005.
  • Liao W, Chen L, Ma X, Jiao R, Li X, Wang Y. Protective effects of kaempferol against reactive oxygen species-induced hemolysis and its antiproliferative activity on human cancer cells. Eur J Med Chem. 2016;114:24–32. doi:10.1016/j.ejmech.2016.02.045.
  • Weng C-J, Chen M-J, Yeh C-T, Yen G-C. Hepatoprotection of quercetin against oxidative stress by induction of metallothionein expression through activating MAPK and PI3K pathways and enhancing Nrf2 DNA-binding activity. N Biotechnol. 2011;28(6):767–77. doi:10.1016/j.nbt.2011.05.003.
  • Sun L, Xu G, Dong Y, Li M, Yang L, Lu W. Quercetin protects against lipopolysaccharide-induced intestinal oxidative stress in broiler chickens through activation of Nrf2 pathway. Molecules. 2020;25(5):1053. doi:10.3390/molecules25051053.
  • Ramyaa P, Krishnaswamy R, Padma VV. Quercetin modulates OTA-induced oxidative stress and redox signalling in HepG2 cells—up regulation of Nrf2 expression and down regulation of NF-κB and COX-2. Biochim Biophys Acta. 2014;1840(1):681–92. doi:10.1016/j.bbagen.2013.10.024.
  • Zhao X, Wang J, Deng Y, Liao L, Zhou M, Peng C, Li Y. Quercetin as a protective agent for liver diseases: a comprehensive descriptive review of the molecular mechanism. Phytother Res. 2021;35(9):4727–47. doi:10.1002/ptr.7104.
  • Liu C-M, Ma J-Q, Xie W-R, Liu S-S, Feng Z-J, Zheng G-H, Wang A-M. Quercetin protects mouse liver against nickel-induced DNA methylation and inflammation associated with the Nrf2/HO-1 and p38/STAT1/NF-κB pathway. Food Chem Toxicol. 2015;82:19–26. doi:10.1016/j.fct.2015.05.001.
  • Lau A, Villeneuve NF, Sun Z, Wong PK, Zhang DD. Dual roles of Nrf2 in cancer. Pharmacol Res. 2008;58(5–6):262–70. doi:10.1016/j.phrs.2008.09.003.
  • Saw CLL, Guo Y, Yang AY, Paredes-Gonzalez X, Ramirez C, Pung D, Kong A-NT. The berry constituents quercetin, kaempferol, and pterostilbene synergistically attenuate reactive oxygen species: involvement of the Nrf2-ARE signaling pathway. Food Chem Toxicol. 2014;72:303–11. doi:10.1016/j.fct.2014.07.038.
  • Tang S-M, Deng X-T, Zhou J, Li Q-P, Ge X-X, Miao L. Pharmacological basis and new insights of quercetin action in respect to its anti-cancer effects. Biomed Pharmacother. 2020;121:109604. 10.1016/j.biopha.2019.109604
  • Lin Y, Shi R, Wang X, Shen HM. Luteolin, a flavonoid with potential for cancer prevention and therapy. Curr Cancer Drug Targets. 2008;8(7):634–46. doi:10.2174/156800908786241050.
  • Lin C-W, Wu M-J, Liu IY-C, Su J-D, Yen J-H. Neurotrophic and cytoprotective action of luteolin in PC12 cells through ERK-dependent induction of Nrf2-driven HO-1 expression. J Agric Food Chem. 2010;58(7):4477–86. doi:10.1021/jf904061x.
  • Yan Y, Jun C, Lu Y, Jiangmei S. Combination of metformin and luteolin synergistically protects carbon tetrachloride‐induced hepatotoxicity: mechanism involves antioxidant, anti‐inflammatory, antiapoptotic, and Nrf2/HO‐1 signaling pathway. Biofactors. 2019;45(4):598–606. 10.1002/biof.1521
  • Li L, Luo W, Qian Y, Zhu W, Qian J, Li J. Luteolin protects against diabetic cardiomyopathy by inhibiting NF-κB-mediated inflammation and activating the Nrf2-mediated antioxidant responses. Phytomedicine. 2019;59:152774.
  • Xiao C, Xia M-L, Wang J, Zhou X-R, Lou Y-Y, Tang L-H, et al. Luteolin attenuates cardiac ischemia/reperfusion injury in diabetic rats by modulating Nrf2 antioxidative function. Oxid Med Cell Longev. 2019;2019:2719252. doi:10.1155/2019/2719252.
  • Rajput SA, Shaukat A, Wu K, Rajput I, Dost R, Baloch M, et al. Luteolin alleviates aflatoxinB1-induced apoptosis and oxidative stress in the liver of mice through activation of Nrf2 signaling pathway. Antioxidants (Basel). 2021;10(8):1268. doi:10.3390/antiox10081268.
  • Yang H, Liu BF, Xie FJ, Yang WL, Cao N. Luteolin induces mitochondrial apoptosis in HT29 cells by inhibiting the Nrf2/ARE signaling pathway. Exp Ther Med. 2020;19(3):2179–2187. doi:10.3892/etm.2020.8464.
  • Wang X-J, Sun Z, Villeneuve NF, Zhang S, Zhao F, Li Y, Chen W, Yi X, Zheng W, Wondrak GT, et al. Nrf2 enhances resistance of cancer cells to chemotherapeutic drugs, the dark side of Nrf2. Carcinogenesis. 2008;29(6):1235–43. doi:10.1093/carcin/bgn095.
  • Chian S, Thapa R, Chi Z, Wang XJ, Tang X. Luteolin inhibits the Nrf2 signaling pathway and tumor growth in vivo. Biochem Biophys Res Commun. 2014;447(4):602–8. doi:10.1016/j.bbrc.2014.04.039.
  • Tsai KJ, Tsai HY, Tsai CC, Chen TY, Hsieh TH, Chen CL, et al. Luteolin inhibits breast cancer stemness and enhances chemosensitivity through the Nrf2-mediated pathway. Molecules. 2021;26(21):6452. doi:10.3390/molecules26216452.
  • Chian S, Li Y-Y, Wang X-J, Tang X-W. Luteolin sensitizes two oxaliplatin-resistant colorectal cancer cell lines to chemotherapeutic drugs via inhibition of the Nrf2 pathway. Asian Pac J Cancer Prev. 2014;15(6):2911–6. doi:10.7314/apjcp.2014.15.6.2911.
  • Nouri Z, Fakhri S, Nouri K, Wallace CE, Farzaei MH, Bishayee A. Targeting multiple signaling pathways in cancer: the rutin therapeutic approach. Cancers. 2020;12(8):2276. 10.3390/cancers12082276
  • Manna K, Khan A, Biswas S, Das U, Sengupta A, Dey S, et al. editors. Rutin reverses radiation-induced oxidative DNA damage and inflammation through the modulation of p38/Nf-Kb and Keap1/Nrf2 pathway. Proceedings of the fourteenth annual meeting of the Society for Free Radical Research-India and international conference on translational research in ionizing radiation, free radicals, antioxidants and functional food; 2016.
  • Kongara S, Karantza V. The interplay between autophagy and ROS in tumorigenesis. Front Oncol. 2012;2:171. doi:10.3389/fonc.2012.00171.
  • Domitrović R, Jakovac H, Vasiljev Marchesi V, Vladimir-Knežević S, Cvijanović O, Tadić Z, Romić Z, Rahelić D. Differential hepatoprotective mechanisms of rutin and quercetin in CCl 4-intoxicated BALB/cN mice. Acta Pharmacol Sin. 2012;33(10):1260–70. 10.1038/aps.2012.62
  • Fulda S. Autophagy in cancer therapy. Front Oncol. 2017;7:128. doi:10.3389/fonc.2017.00128.
  • Oluranti OI, Alabi BA, Michael OS, Ojo AO, Fatokun BP. Rutin prevents cardiac oxidative stress and inflammation induced by bisphenol A and dibutyl phthalate exposure via NRF-2/NF-κB pathway. Life Sci. 2021;284:119878. doi:10.1016/j.lfs.2021.119878.
  • Gęgotek A, Ambrożewicz E, Jastrząb A, Jarocka-Karpowicz I, Skrzydlewska E. Rutin and ascorbic acid cooperation in antioxidant and antiapoptotic effect on human skin keratinocytes and fibroblasts exposed to UVA and UVB radiation. Arch Dermatol Res. 2019;311(3):203–19. doi:10.1007/s00403-019-01898-w.
  • Hussein RM, Mohamed WR, Omar HA. A neuroprotective role of kaempferol against chlorpyrifos-induced oxidative stress and memory deficits in rats via GSK3β-Nrf2 signaling pathway. Pestic Biochem Physiol. 2018;152:29–37. doi:10.1016/j.pestbp.2018.08.008.
  • Alshehri AS. Kaempferol attenuates diabetic nephropathy in streptozotocin-induced diabetic rats by a hypoglycaemic effect and concomitant activation of the Nrf-2/Ho-1/antioxidants axis. Arch Physiol Biochem. 2021;1–14. doi:10.1080/13813455.2021.1890129.
  • Wang Z, Sun W, Sun X, Wang Y, Zhou M. Kaempferol ameliorates Cisplatin induced nephrotoxicity by modulating oxidative stress, inflammation and apoptosis via ERK and NF-κB pathways. AMB Express. 2020;10(1):1–11.
  • Kumar AN, Bevara GB, Kaja LK, Badana AK, Malla RR. Protective effect of 3-O-methyl quercetin and kaempferol from Semecarpus anacardium against H2O2 induced cytotoxicity in lung and liver cells. BMC Complement Altern Med. 2016;16(1):1–13. doi:10.1186/s12906-016-1354-z.
  • Kitakaze T, Makiyama A, Nakai R, Kimura Y, Ashida H. Kaempferol modulates TCDD-and t-BHQ-induced drug-metabolizing enzymes and luteolin enhances this effect. Food Funct. 2020;11(4):3668–80. doi:10.1039/c9fo02951f.
  • Yao H, Sun J, Wei J, Zhang X, Chen B, Lin Y. Kaempferol protects blood vessels from damage induced by oxidative stress and inflammation in association with the Nrf2/HO-1 signaling pathway. Front Pharmacol. 2020;11:1118. doi:10.3389/fphar.2020.01118.
  • Yang Y, Wang Y, Wang T, Jiang X. O26 Screening active components of Modified Xiaoyao Powder for chemoprevention in breast cancer cells: involvement of the NRF2/NQO1 signalling pathway. Biochem Pharmacol. 2017;139:117–8. doi:10.1016/j.bcp.2017.06.091.
  • El-Kott AF, Bin-Meferij MM, Eleawa SM, Alshehri MM. Kaempferol protects against cadmium chloride-induced memory loss and hippocampal apoptosis by increased intracellular glutathione stores and activation of PTEN/AMPK induced inhibition of Akt/mTOR signaling. Neurochem Res. 2020;45(2):295–309. doi:10.1007/s11064-019-02911-4.
  • Polivka J, Jr., Janku F. Molecular targets for cancer therapy in the PI3K/AKT/mTOR pathway. Pharmacol Ther. 2014;142(2):164–75. doi:10.1016/j.pharmthera.2013.12.004.
  • Wang Y, Chen B, Wang Z, Zhang W, Hao K, Chen Y, et al. Marsdenia tenacissimae extraction (MTE) inhibits the proliferation and induces the apoptosis of human acute T cell leukemia cells through inactivating PI3K/AKT/mTOR signaling pathway via PTEN enhancement. Oncotarget. 2016;7(50):82851.
  • Cho HJ, Park JHY. Kaempferol induces cell cycle arrest in HT-29 human colon cancer cells. J Cancer Prev. 2013;18(3):257–63. doi:10.15430/jcp.2013.18.3.257.
  • Al Sabaani N. Kaempferol protects against hydrogen peroxide-induced retinal pigment epithelium cell inflammation and apoptosis by activation of SIRT1 and inhibition of PARP1. J Ocul Pharmacol Ther. 2020;36(7):563–77. doi:10.1089/jop.2019.0151.
  • Gao Y, Yin J, Rankin GO, Chen YC. Kaempferol induces G2/M cell cycle arrest via checkpoint kinase 2 and promotes apoptosis via death receptors in human ovarian carcinoma A2780/CP70 cells. Molecules. 2018;23(5):1095.
  • Kang G-l, Jing Z-x Kaempferol alleviates ox-LDL-mediated endothelial cell injury via regulating AMPK/Nrf2/HO-1 signaling pathway. Chin J Microbiol Immunol. 2018;34(4):525–30.
  • Du Y, Han J, Zhang H, Xu J, Jiang L, Ge W. Kaempferol prevents against Ang II-induced cardiac remodeling through attenuating Ang II-induced inflammation and oxidative stress. J Cardiovasc Pharmacol. 2019;74(4):326–35. doi:10.1097/fjc.0000000000000713.