Molecular Mechanism for Quercetin-Induced Cancer and SCD: Comparison
Please note this is a comparison between Version 2 by Jessie Wu and Version 4 by Jessie Wu.

The term “ROS (Reactive Oxygen Species)” refers to radicals and ions that contain an unpaired numbered electron in its outmost electron field which are highly reactive metabolic byproducts that can have both harmful and useful effects within the cell. Quercetin (QC), a plant-derived bioflavonoid, is known for its ROS scavenging properties and was recently discovered to have various antitumor properties in a variety of solid tumors. Adaptive stress responses may be induced by persistent ROS stress, allowing cancer cells to survive with high levels of ROS while maintaining cellular viability. Large amounts of ROS make cancer cells extremely susceptible to quercetin, one of the most available dietary flavonoids. Because of the molecular and metabolic distinctions between malignant and normal cells, targeting ROS metabolism might help overcome medication resistance and achieve therapeutic selectivity while having little or no effect on normal cells. The powerful bioactivity and modulatory role of quercetin has prompted extensive research into the chemical, which has identified a number of pathways that potentially work together to prevent cancer, alongside, QC has a great number of evidences to use as a therapeutic agent in cancer stem cells.

  • ROS
  • quercetin
  • cancer stem cells

1. Introduction

Due to its antioxidant properties, quercetin is used therapeutically for a variety of disorders. Mainly, the scavenging activity of reactive oxygen species (ROS) is converted to hydroxyl ions (OH) and holds the electron exchange capability [1]. Quercetin (QC) is lipophilic in nature and can easily pass through the cell membranes and latterly activate numerous intracellular signaling pathways, which are health effective [2]. It has been demonstrated that it mediates not only intrinsic (mainly mitochondrial) but also extrinsic (Fas/FasL) apoptotic cancer cell death (Figure 1) [3][4]. In addition to apoptosis, previous research has suggested that QC plays an important role in arresting the cell cycle and controlling the expression of CDK enzyme (cyclin-dependent kinases) [5]. Recent reports mention that it hinders the activity of cytochrome P450 (CYP) enzymes in hepatocytes and that most of them have crucial roles in the metabolism of drugs [6]. QC has also been shown to suppress metastatic protein expression such as MMPs (matrix metalloproteases) [7]. Actually, the metastasis process is also supported via neo-angiogenesis, and interestingly, QC is also noted to hinder neovascularization within the microenvironment of tumors [8]. Another antitumor property of QC is its ability to inhibit inflammatory mediators including IFN-γ, IL-6, COX-2, IL-8, iNOS, TNF-α, and many other cancer inflammatory mechanisms [9]. Investigating the mechanistic basis for these bioactive metabolites’ actions will aid in our understanding of cancer biology (Table 1). Besides, here is also included an extensive number of scientific outcomes, where QC is a vital therapeutic agent in cancer stem cells and mitigates the devastating role of cancers, and the results are represented in Table 2.
Figure 1. The anticancer pathways and mechanisms of action induced by quercetin.

2. Reactive Oxygen Species-Mediated Regulation of the MAPK/ERK1/2 Pathways

Quercetin plays a crucial role in several signal transduction pathways. Herein, it directly activates the MAPK/ERK-mediated pathways, leading to the apoptosis process and the mechanism discovered within the A549 cell lines of lung carcinoma [10]. Besides, it induces the production of inflammatory and/or proinflammatory cytokines, and inhibits the formation of cyto-kinase LPS, which results in the suppression of iNOS by following the ERK, MAPK, as well as p38 mechanism [11]. Basically, ERK1 or 2 is the major section of the MAPK cascade mechanism, these build up kinase family proteins namely MEK, Raf, and ERK1/2 that work successively [12]. The active ERK 1 or 2 causes are reprogramming events linked to the expression of genes through the active phosphorylation process of various intracellular molecular target proteins and other transcription factors [13].
Table 2. Tabular representation of quercetin-mediated cancer stem cell killing with their mechanistic illustration.

3. Reactive Oxygen Species-Mediated Regulation of the p53 Pathway

The p53 pathway has been found to modulate cellular stress responses and to be vital in the direction of apoptosis in cancer. It is directly associated with MDM2 because of the complexities of p53, and is regulated at low levels by the regular proteasomal breakdown in unstressed cells [46]. Similar to other cell signaling molecules, reactive oxygen species (ROS) function as an influencer in the activation of p53 through upstream signal transduction and also act as a downstream component to induce cell death (Figure 2) [47]. Quercetin, like other omnipresent bioactive plant flavonoids, has been demonstrated to reduce overall tumor growth and metastasis [48]. Quercetin inhibited tumor cell proliferation by concurrently stimulating two divergent regulatory networks, p53 and NF-κB [49]. An investigation reported that in A549 human lung cancer cells expressing wild-type p53, cell growth arrest and apoptosis were accelerated by quercetin considerably [50].
Figure 2. Mechanistic illustration of the function of quercetin in cancers cells by regulating ROS and p53-related pathways.
Another study found that while quercetin increased p53 phosphorylation, it did not increase p53 mRNA transcription. Quercetin increased p21 expression while suppressing cyclin D1 expression, favoring cell cycle arrest. In favor of apoptosis, the enhancement of the Bax/Bcl-2 ratio with this sort of therapy is also mediated by quercetin. Noticeably, the prevention of p53 mRNA degradation at the post-transcriptional stage occurred through quercetin-mediated signaling [51].
Lee, Yoon-Jin et al. showed that experiment on MM MSTO-211H cells, quercetin induced apoptotic cell death at low concentrations. According to the study, the up-regulation of p53 induced by quercetin can be visualized at both mRNA and protein levels without altering its ubiquitination as well as enhancing caspase-3/7 activity and procaspase-3 and PARP cleavages. The subset of cells in the G2/M phase that has been obstructed by the pan-caspase inhibitor escalated later in a quercetin concentration-dependent approach [52]. Another study found that quercetin interacts with the human glioblastoma A172 as well as LBC3 cell lines, confirming increased ROS production, unregulated SOD1 and SOD2 expression, ATP depletion, and CHOP8 overexpression, and as a result, apoptosis of these brain cancer cell lines. They also reported enhanced caspase-9 expression and activity, supporting a mitochondrial apoptosis pathway [53].

4. Reactive Oxygen Species-Mediated Regulation of the JAK/STAT and TRAIL Pathways

The JAK-STAT pathway is a well-characterized signaling pathway involved in the production of a variety of inflammatory and/or proinflammatory cytokines along with growth-controlling factors. The connection of ligand molecules to specific receptors leads to JAK mechanism activation that also improves the auto-phosphorylation process and activates the STAT signaling pathways [54]. Several research studies revealed that the active form of mast cells directly increases the formation of Th2 cells such as cytokines, and at the same time reduces the secretion of cytokines mainly Th1. Similarly, mast cells regulate the expression of the JAK/STAT gene, which stimulates IL-13 development in the Th2 cell line [55].
Quercetin can effectively inhibit the signaling pathways of JAK-STAT in various inflammatory conditions. Additionally, quercetin therapy of activated T-cells suppressed the phosphorylation of the TiK-2 (tyrosine kinase-2), JAK2, STAT-3,4 enzymes stimulated by interleukin-12, resulting in diminished T-cell proliferation, and Th1 variation [56]. As a result, quercetin’s anti-inflammatory and anti-apoptotic properties play an important role in cancer prevention by modulating the TLR-2 (toll-like receptor-2) and JAK-2/STAT-3 pathways and significantly inhibiting STAT-3 tyrosine phosphorylation within the inflammatory cells [57]. QC-mediated pretreatment of cholangiocarcinoma cells inhibited the JAK/STAT cascade pathway-mediated activation of iNOS and activation of the ICAM-1 (intercellular adhesion molecule-1). Additionally, QC inhibited the inflammatory cytokine IL-6 and interferon-alpha activation [58].

5. Reactive Oxygen Species-Mediated Regulation of the AMPKα1/ASK1/p38 Pathways

Understanding the mechanisms of apoptotic pathways can lead to the development of effective cancer preventive and treatment options. Particularly, outside stimuli and mitogen-activated protein (MAP) kinases are two well-known intracellular signaling pathways that possess effective signals of the apoptosis transducing channel named the kinase cascade. Apoptosis signal-regulating kinase (ASK-1) is a redox sensor that belongs to the ROS-sensitive MAP kinase kinases family [59]. The chemopreventive drug, quercetin, reported potent apoptosis induction in cancer cells mediated by ASK-1. Quercetin is responsible for free radical-induced cell death by the production of sufficient reactive oxygen species (ROS) though it has conventional anti-oxidation activity in the body. Through the apoptosis signal-regulating kinase (ASK)-1 and mitogen-activated protein kinase pathways, reswearchers studied the regulation mechanism of quercetin-induced apoptosis [60]. A therapeutic technique for cancer treatment is to induce apoptosis in cancer cells [61]. Quercetin (3,30,40,5,7-pentahydroxyflavone) has a broad-spectrum of biological effects, along-with cancer prevention and tumor growth suppression [62]. After scavenging the peroxyl radicals present, quercetin is transformed to quercetin-O-, a radical form that emerges to be one of the multiple pathways participating in the formation of ROS by quercetin [60]. Due to the generation of quercetin radicals (quercetin-O•) throughout its interaction with peroxyl radicals, quercetin also enhances intracellular reactive oxygen species (ROS), inducing free radical-induced apoptosis via the ROS/AMPK1/ASK1/p38 and AMPK1/COX2 signaling pathways [63]. AMPK1 and ASK1 are activated when the quantity of ROS produced increases in which AMPK1 and ASK1 activate p38 and boost the activity of several caspases [62]. In t he current study, it was reported that the quercetin-mediated apoptosis process is regulated by the activation of AMP-activated protein kinase (AMPK)/p38 MAPK signaling pathways and enhanced expression of sestrin 2. According to the findings, quercetin triggered apoptosis via raising the expression of sestrin 2 and creating intracellular reactive oxygen species (ROS). The activation of the AMPK/p38 signaling pathway by quercetin reported the induction of apoptosis, which was dependent on sestrin 2. Nevertheless, silencing sestrin 2 with short-interfering RNA (siRNA) demonstrated that quercetin had no effect on AMPK or p38 phosphorylation inside the cells where sestrin 2 was silenced [61]. Another AMPK1 downstream target involved in Quercetin-induced apoptosis is COX-2 [64]. Quercetin radicals have also been reported to reduce the intracellular glutathione (GSH) pool in a concentration-mediated manner, as well as initiate apoptosis via mitochondrial depolarization. A research study demonstrated that quercetin exhibited interaction with the estrogen binding site type II [63]. Apoptosis signal-regulating kinase 1 (ASK1) is a mitogen-activated protein (MAP) kinase that stimulates both the MKK3/MKK6–p38 MAP and MKK4/MKK7–JNK kinase pathways and is a key signaling route in many forms of stress-induced apoptosis. ASK1 has been discovered to have an important role in apoptosis triggered by stress, particularly oxidative stress, and endoplasmic reticulum (ER) stress [65]. AMP-activated protein kinase (AMPK) appeared to be an important regulator of quercetin-mediated ASK1/p38 activation. When the activity of AMPKa1 was inhibited with a synthetic inhibitor or siRNA, quercetin-activated ASK1 was unable to promote p38 activity. As a result, it is thought that quercetin’s apoptotic effects are mediated by the ROS/AMPKa1/ASK1/p38 signaling pathway, and that AMPKa1 is required for ASK1-induced apoptosis. The alanine substitution of NR4A2’s oxidative stress-mediated phosphorylation sites reduces NR4A2’s cytoplasmic translocation and pro-necrosis activity, implying that p38-dependent phosphorylation of NR4A2 is a critical regulation to bestow the necrosis-inducing activity to NR4A2 [66]. These research outcomes imply that the ROS/AMPKa1/ASK1/p38 signaling pathway is involved in quercetin-induced apoptosis, and that AMPKa1 is a key regulator of ASK1 [60].

6. Reactive Oxygen Species-Mediated Regulation of the RAGE/PI3K/AKT/mTOR Axis

AKT signaling is required for the maintenance of normal physiological conditions. AKT signaling is frequently activated in cancer, which keeps the cellular microenvironment of tumors in a highly oxidative state, which is necessary for tumor development. As a result, antioxidants are hypothesized to possess anticancer effects through their ability to disrupt the tumor microenvironment. According to the findings of the study, quercetin (QC) prevented AKT signaling, which led to less cell survival, inflammation, and blood vessel growth in mice with lymphoma [67]. In human pancreatic adenocarcinoma cells, QC enhanced cell death and chemosensitivity via the regulation of RAGE/PI3K/AKT/mTOR axis [43].
Angiogenesis is a critical phase in cancer growth and spread because it allows the growing tumor to get oxygen and nutrients. In vitro, in vivo, and ex vivo antiangiogenic potential of quercetin reported that the progression of human prostate cancer was suppressed through the regulation of VEGFR-2-mediated Akt/mTOR/P70S6K signaling pathways [68]. Quercetin treatment dramatically reduced the overexpression of pro-fibrotic factors (IL-6, IL-8, COL-1, COL-3, and LC3) as well as the increase in pro-fibrotic signaling mediators (mTOR and AKT) generated by LPS in WI-38 cells. Thus, nasogastric injection of QC significantly reversed an elevation in profibrotic markers (VEGF, IL-6, TGF, COL-1, and COL-3) and fibrotic signaling mediators (mTOR and Akt) in the rabbit tracheal stenosis research model, while also inactivating ATG5 [69]. QC blocked the PI3K/AKT/mTOR and STAT3 signaling pathways in primary effusion lymphoma cells, which induced cell death and autophagy [70]. In addition, quercetin showed its anti-breast cancer effect via inhibiting the Akt/AMPK/mTOR signaling cascade [71]. Because it inhibited the PI3K/Akt/mTOR signaling pathway and has been proven to reduce breast cancer stem cells (CD44+/CD24), it was identified as a candidate for the treatment of breast cancer [72].
Q-6-C-b-D-glucopyranoside, a naturally occurring derivative of quercetin, demonstrated anti-prostate cancer action via blocking the Akt-mTOR pathway through the aryl hydrocarbon receptor [73]. The generation of human breast cancer stem cells was inhibited due to the inhibition of the Notch1 and PI3K/Akt signaling pathways by quercetin-3-methyl ether [74]. In addition, quercetin-3-methyl ether stopped the growth of cancer in the esophagus by blocking the Akt/mTOR/P70S6k and MAPK pathways, which are important for the growth of cancer [75]. P53, Akt/mTOR pathway, and cancer cell growth were all downregulated in breast cancer cells when the vanadium quercetin complex is used, which is also associated with many apoptosis events [76]. On the contrary, Burkitt’s lymphoma cells are significantly more susceptible to death because of quercetin’s induction of a drop in c-Myc expression in addition to PI3K/AKT/mTOR signaling [77]. Quercetin inhibited glycolysis via activating the Akt-mTOR pathway, resulting in the activation of autophagy, hence inhibiting the migration of breast cancer cells [78]. In glycolysis-mediated HCC cells, it suppressed the hexokinase 2 and the Akt-mTOR pathway, which are involved in glycolysis-dependent cell proliferation [79]. Quercetin has been shown to suppress the Akt-mTOR pathway and hypoxia-induced factor 1 signaling pathway in gastric cancer cells, resulting in preventative autophagy [80].
Furthermore, quercetin possessed strong synergistic activity in combination with anti-sense oligo gene therapy (suppressing snail gene expression), it significantly inhibited the progression of renal cell carcinoma Caki-2 by the regulation of Akt/mTOR/ERK1/2 signaling pathways [81]. Quercetin nanoparticles inhibited the AKT/ERK/caspase-3 signaling pathway to induce autophagy and apoptosis in human neuroglioma cell lines in vitro and in vivo research models [82]. In breast cancer cell lines (MCF-7 and MDA-MB-231), quercetin coupled with gold nanoparticles promoted apoptosis by inhibiting the EGFR/P13K/Akt-mediated pathway [42].

7. Reactive Oxygen Species-Mediated Regulation of the HMGB1 and NF-κB Pathways

The HMGB1 (high mobility group box 1) protein that is exposed to the outside environment has the potential to trigger the production of TNF (tumor necrosis factor), IL-1, and other inflammatory cytokines from monocytes [83]. QC enhances the suppression of HMGB1-induced TNF and IL-1 mRNA production, indicating that it transmits signals inside cells that modulate the action of proinflammatory cytokines [84]. In HMGB1-induced gene expression, the activation of MAPK is important for the release of inflammatory cytokines such as IL-1 and TNF from neutrophils, macrophages, and endothelial cells. The release of cytokines induced by HMGB1 partially interferes with MAPK pathways in which the p38 phosphorylation or C-Junction of N-terminal kinases with extracellular signals regulate kinases in macrophages in a time-dependent manner by HMGB1 or LPS. Here, such a bioactive compound significantly inhibits the activity of each kinase enzyme by maintaining the phosphorylation process [85]. Additionally, a recent study reported that QC can easily modulate the signaling proteins (i.e., Cox-2, NF-κB) that may lead to the apoptosis process, and furthermore, suppress anti-apoptotic proteins, e.g., Bcl-2 as well as Bcl-xL, up-regulating Bax with several pro-apoptotic proteins [86][87]. In addition to activating MAPK, the nuclear factor-κB (NF-κB) pathway also includes HMGB1-induced cellular expression, and NF-κB-dependent gene expression is of significant importance when exposing cytokine [88][89]. Through their interaction with the IκBα (members of the IκB family), subunits of NF-κB (p50 as well as p65) take place in cells as inactive trimers within the cell cytosol [90]. Interestingly, QC suppresses the degradation of IκBα factor and NF-κB p65 in a significant manner. As a result, following stimulation with the following p65, HMGB1 or LPS, the critical mediator of NF-κB factor will become available to be regulated nuclear NF-κB genes and the compound QC most effectively inhibits the nuclear localization process [13].

8. Reactive Oxygen Species -Mediated Regulation of the Nrf2-Induced Phase II Enzyme and Signaling Pathways

Carcinogenesis can be suppressed through inductions from Phase II enzymes, including the GST (glutathione S transferase), NAD(P)H/NO (quinone oxidoreductase), UDP-GLT (UDP-glucuronosyl transferases), and HO-1 (heme oxygenase-1) [91][92]. The genes of such enzymes bearing the ARE (antioxidant replication elements), which are strictly regulated by Nrf2 nuclear erythroid, and which, in turn, are associated by Keap-1 (Kelch-like CEH-associated protein-1), therefore, an Nrf2 represent, also encourage their degradation through the path of the ubicin-dependent proteasome [93][94][95]. When cells are treated with ARE-mediated stimulants such as QC, the complex formation of “Nrf2-Keap1” dissociates, allowing the Nrf2 factor to be translocated from the cytosol to the nuclear part of the cell. By binding to ARE and subsequently forms a heterodimer structure with other transcription factors (TFs), which simultaneously promotes the transcription of phase II enzyme genes. [96][97]. QC has also been demonstrated to stimulate ARE binding activity and regulate the mRNA expression level of NQO1, the total mechanism that occurred in the HepG2 cell line in a dose-dependent manner [97]. In addition, QC regulated the prohibition of Nrf-2 protein degradation and maintained the reduced amount of posttranslational products of Keap-1 proteins without any impairment of the complex structure dissociation like- “Keap-1-Nrf2” [97]. A cross-mediated increase in enzymes of phase II Caco-2 adenocarcinoma with duodenum adenocarcinoma of HuTu 80 cells in human CRAC (colorectal adenocarcinoma) was shown [98]. The effect of QC on nuclear translocation of Nrf-2 in a time-dependent manner, and increased expression level in HepG2, MgM (malignant mesothelioma) MSTO-211H, and H2452 cells at mRNA and protein quantity has been reported recently [99]. Moreover, the therapeutic efficacy of QC has also been defined in rat models through the activation of Nrf-2/HO-1 against high glucose-induced damage [100]. Until now, little is known about the precise biochemical pathways by which QC triggers Nof2-based gene expression. Nevertheless, evidence suggests that the improvement of diverse signal transduction cascades, along with the MAPK, modulates cross-induced interpretation of transcriptional genes (MAPK). The p38 and ERK-generated translocation of Nrf2 to nuclei between MAPK signal pathways were shown to be responsible for the successive initiation and activity of HO-1 expression [101][102][103]. Indeed, QC (50 µM) protected RAW264.7 macrophages from the induction of apoptosis by up-regulating phase II enzymes, which include HO-1, via ERK pathway-dependent mechanisms [102]. Similarly, cross-treated human hepatocytes were mentioned in MAPK HO-1-dependent upregulation and successive preservation against ethanol-induced oxidative damage [103].

References

  1. Boots, A.W.; Haenen, G.R.; Hartog, G.J.D.; Bast, A. Oxidative damage shifts from lipid peroxidation to thiol arylation by catechol-containing antioxidants. Biochim. Biophys. Acta (BBA) Mol. Cell Biol. Lipids 2002, 1583, 279–284.
  2. Chen, X. Protective effects of quercetin on liver injury induced by ethanol. Pharmacogn. Mag. 2010, 6, 135.
  3. Chien, S.Y.; Wu, Y.C.; Chung, J.G.; Yang, J.S.; Lu, H.F.; Tsou, M.F.; Wood, W.G.; Kuo, S.J.; Chen, D.R. Quercetin-induced apoptosis acts through mitochondrial- and caspase-3-dependent pathways in human breast cancer MDA-MB-231 cells. Hum Exp. Toxicol. 2009, 28, 493–503.
  4. Kashyap, D.; Mittal, S.; Sak, K.; Singhal, P.; Tuli, H.S. Molecular mechanisms of action of quercetin in cancer: Recent advances. Tumor Biol. 2016, 37, 12927–12939.
  5. Kumar, S.; Pandey, A.K.J. Chemistry and biological activities of flavonoids: An overview. Sci. World J. 2013, 2013, 162750.
  6. Lautraite, S.; Musonda, A.; Doehmer, J.; Edwards, G.; Chipman, J.J.M. Flavonoids inhibit genetic toxicity produced by carcinogens in cells expressing CYP1A2 and CYP1A1. Mutagenesis 2002, 17, 45–53.
  7. Ong, C.S.; Tran, E.; Nguyen, T.T.; Ong, C.K.; Lee, S.K.; Lee, J.J.; Ng, C.P.; Leong, C.; Huynh, H.J. Quercetin-induced growth inhibition and cell death in nasopharyngeal carcinoma cells are associated with increase in Bad and hypophosphorylated retinoblastoma expressions. Oncol. Rep. 2004, 11, 727–733.
  8. Tan, W.-f.; Lin, L.-p.; Li, M.-h.; Zhang, Y.-X.; Tong, Y.-g.; Xiao, D.; Ding, J. Quercetin, a dietary-derived flavonoid, possesses antiangiogenic potential. Eur. J. Pharmacol. 2003, 459, 255–262.
  9. Vijayababu, M.; Arunkumar, A.; Kanagaraj, P.; Venkataraman, P.; Krishnamoorthy, G.; Arunakaran, J. Quercetin downregulates matrix metalloproteinases 2 and 9 proteins expression in prostate cancer cells (PC-3). Mol. Cell. Biochem. 2006, 287, 109–116.
  10. Nguyen, T.; Tran, E.; Nguyen, T.; Do, P.; Huynh, T.; Huynh, H.J.C. The role of activated MEK-ERK pathway in quercetin-induced growth inhibition and apoptosis in A549 lung cancer cells. Carcinogenesis 2004, 25, 647–659.
  11. Cho, S.-Y.; Park, S.-J.; Kwon, M.-J.; Jeong, T.-S.; Bok, S.-H.; Choi, W.-Y.; Jeong, W.-I.; Ryu, S.-Y.; Do, S.-H.; Lee, C.-S.J.M.; et al. Quercetin suppresses proinflammatory cytokines production through MAP kinases and NF-κB pathway in lipopolysaccharide-stimulated macrophage. Mol. Cell. Biochem. 2003, 243, 153–160.
  12. Raman, M.; Chen, W.; Cobb, M.J.O. Differential regulation and properties of MAPKs. Oncogene 2007, 26, 3100–3112.
  13. Khan, F.; Niaz, K.; Maqbool, F.; Ismail Hassan, F.; Abdollahi, M.; Nagulapalli Venkata, K.C.; Nabavi, S.M.; Bishayee, A. Molecular Targets Underlying the Anticancer Effects of Quercetin: An Update. Nutrients 2016, 8, 529.
  14. Lee, T.-J.; Kim, O.H.; Kim, Y.H.; Lim, J.H.; Kim, S.; Park, J.-W.; Kwon, T.K. Quercetin arrests G2/M phase and induces caspase-dependent cell death in U937 cells. Cancer Lett. 2006, 240, 234–242.
  15. Lee, W.J.; Hsiao, M.; Chang, J.L.; Yang, S.F.; Tseng, T.H.; Cheng, C.W.; Chow, J.M.; Lin, K.H.; Lin, Y.W.; Liu, C.C.; et al. Quercetin induces mitochondrial-derived apoptosis via reactive oxygen species-mediated ERK activation in HL-60 leukemia cells and xenograft. Arch. Toxicol. 2015, 89, 1103–1117.
  16. Niu, G.; Yin, S.; Xie, S.; Li, Y.; Nie, D.; Ma, L.; Wang, X.; Wu, Y. Quercetin induces apoptosis by activating caspase-3 and regulating Bcl-2 and cyclooxygenase-2 pathways in human HL-60 cells. Acta Biochim. Biophys. Sin. 2011, 43, 30–37.
  17. Mandal, A.K.; Das, S.; Mitra, M.; Chakrabarti, R.N.; Chatterjee, M.; Das, N. Vesicular flavonoid in combating diethylnitrosamine induced hepatocarcinoma in rat model. J. Exp. Ther. Oncol. 2008, 7, 123–133.
  18. Zhou, J.; Fang, L.; Liao, J.; Li, L.; Yao, W.; Xiong, Z.; Zhou, X. Investigation of the anti-cancer effect of quercetin on HepG2 cells in vivo. PLoS ONE 2017, 12, e0172838.
  19. Howells, L.M.; Britton, R.G.; Mazzoletti, M.; Greaves, P.; Broggini, M.; Brown, K.; Steward, W.P.; Gescher, A.J.; Sale, S. Preclinical colorectal cancer chemopreventive efficacy and p53-modulating activity of 3′,4′,5′-trimethoxyflavonol, a quercetin analogue. Cancer Prev. Res. 2010, 3, 929–939.
  20. Del Follo-Martinez, A.; Banerjee, N.; Li, X.; Safe, S.; Mertens-Talcott, S. Resveratrol and quercetin in combination have anticancer activity in colon cancer cells and repress oncogenic microRNA-27a. Nutr. Cancer 2013, 65, 494–504.
  21. Shan, B.E.; Wang, M.X.; Li, R.Q. Quercetin inhibit human SW480 colon cancer growth in association with inhibition of cyclin D1 and survivin expression through Wnt/beta-catenin signaling pathway. Cancer Investig. 2009, 27, 604–612.
  22. Dihal, A.A.; de Boer, V.C.; van der Woude, H.; Tilburgs, C.; Bruijntjes, J.P.; Alink, G.M.; Rietjens, I.M.; Woutersen, R.A.; Stierum, R.H. Quercetin, but not its glycosidated conjugate rutin, inhibits azoxymethane-induced colorectal carcinogenesis in F344 rats. J. Nutr. 2006, 136, 2862–2867.
  23. Chen, W.; Wang, X.; Zhuang, J.; Zhang, L.; Lin, Y. Induction of death receptor 5 and suppression of survivin contribute to sensitization of TRAIL-induced cytotoxicity by quercetin in non-small cell lung cancer cells. Carcinogenesis 2007, 28, 2114–2121.
  24. Youn, H.; Jeong, J.C.; Jeong, Y.S.; Kim, E.J.; Um, S.J. Quercetin potentiates apoptosis by inhibiting nuclear factor-kappaB signaling in H460 lung cancer cells. Biol. Pharm. Bull. 2013, 36, 944–951.
  25. Yeh, S.L.; Yeh, C.L.; Chan, S.T.; Chuang, C.H. Plasma rich in quercetin metabolites induces G2/M arrest by upregulating PPAR-γ expression in human A549 lung cancer cells. Planta Med. 2011, 77, 992–998.
  26. Gao, X.; Wang, B.; Wei, X.; Men, K.; Zheng, F.; Zhou, Y.; Zheng, Y.; Gou, M.; Huang, M.; Guo, G.; et al. Anticancer effect and mechanism of polymer micelle-encapsulated quercetin on ovarian cancer. Nanoscale 2012, 4, 7021–7030.
  27. Catanzaro, D.; Ragazzi, E.; Vianello, C.; Caparrotta, L.; Montopoli, M. Effect of Quercetin on Cell Cycle and Cyclin Expression in Ovarian Carcinoma and Osteosarcoma Cell Lines. Nat. Prod. Commun. 2015, 10, 1365–1368.
  28. Duo, J.; Ying, G.G.; Wang, G.W.; Zhang, L. Quercetin inhibits human breast cancer cell proliferation and induces apoptosis via Bcl-2 and Bax regulation. Mol. Med. Rep. 2012, 5, 1453–1456.
  29. Srinivasan, A.; Thangavel, C.; Liu, Y.; Shoyele, S.; Den, R.B.; Selvakumar, P.; Lakshmikuttyamma, A. Quercetin regulates β-catenin signaling and reduces the migration of triple negative breast cancer. Mol. Carcinog. 2016, 55, 743–756.
  30. Cao, L.; Yang, Y.; Ye, Z.; Lin, B.; Zeng, J.; Li, C.; Liang, T.; Zhou, K.; Li, J. Quercetin-3-methyl ether suppresses human breast cancer stem cell formation by inhibiting the Notch1 and PI3K/Akt signaling pathways. Int. J. Mol. Med. 2018, 42, 1625–1636.
  31. Iriti, M.; Kubina, R.; Cochis, A.; Sorrentino, R.; Varoni, E.M.; Kabała-Dzik, A.; Azzimonti, B.; Dziedzic, A.; Rimondini, L.; Wojtyczka, R.D. Rutin, a Quercetin Glycoside, Restores Chemosensitivity in Human Breast Cancer Cells. Phytother. Res. 2017, 31, 1529–1538.
  32. Shen, X.; Si, Y.; Wang, Z.; Wang, J.; Guo, Y.; Zhang, X. Quercetin inhibits the growth of human gastric cancer stem cells by inducing mitochondrial-dependent apoptosis through the inhibition of PI3K/Akt signaling. Int. J. Mol. Med. 2016, 38, 619–626.
  33. Yu, D.; Ye, T.; Xiang, Y.; Shi, Z.; Zhang, J.; Lou, B.; Zhang, F.; Chen, B.; Zhou, M. Quercetin inhibits epithelial-mesenchymal transition, decreases invasiveness and metastasis, and reverses IL-6 induced epithelial-mesenchymal transition, expression of MMP by inhibiting STAT3 signaling in pancreatic cancer cells. OncoTargets Ther. 2017, 10, 4719–4729.
  34. Yang, F.; Jiang, X.; Song, L.; Wang, H.; Mei, Z.; Xu, Z.; Xing, N. Quercetin inhibits angiogenesis through thrombospondin-1 upregulation to antagonize human prostate cancer PC-3 cell growth in vitro and in vivo. Oncol. Rep. 2016, 35, 1602–1610.
  35. Atashpour, S.; Fouladdel, S.; Movahhed, T.K.; Barzegar, E.; Ghahremani, M.H.; Ostad, S.N.; Azizi, E. Quercetin induces cell cycle arrest and apoptosis in CD133(+) cancer stem cells of human colorectal HT29 cancer cell line and enhances anticancer effects of doxorubicin. Iran. J. Basic Med. Sci. 2015, 18, 635–643.
  36. Wang, R.; Yang, L.; Li, S.; Ye, D.; Yang, L.; Liu, Q.; Zhao, Z.; Cai, Q.; Tan, J.; Li, X. Quercetin Inhibits Breast Cancer Stem Cells via Downregulation of Aldehyde Dehydrogenase 1A1 (ALDH1A1), Chemokine Receptor Type 4 (CXCR4), Mucin 1 (MUC1), and Epithelial Cell Adhesion Molecule (EpCAM). Med. Sci. Monit. 2018, 24, 412–420.
  37. Tang, S.N.; Singh, C.; Nall, D.; Meeker, D.; Shankar, S.; Srivastava, R.K. The dietary bioflavonoid quercetin synergizes with epigallocathechin gallate (EGCG) to inhibit prostate cancer stem cell characteristics, invasion, migration and epithelial-mesenchymal transition. J. Mol. Signal. 2010, 5, 14.
  38. Crans, D.C.; Yang, L.; Haase, A.; Yang, X. Health benefits of vanadium and its potential as an anticancer agent. Met. Ions Life Sci. 2018, 18, 251–279.
  39. Erdogan, S.; Turkekul, K.; Dibirdik, I.; Doganlar, O.; Doganlar, Z.B.; Bilir, A.; Oktem, G. Midkine downregulation increases the efficacy of quercetin on prostate cancer stem cell survival and migration through PI3K/AKT and MAPK/ERK pathway. Biomed. Pharmacother. 2018, 107, 793–805.
  40. Tang, S.N.; Fu, J.; Nall, D.; Rodova, M.; Shankar, S.; Srivastava, R.K. Inhibition of sonic hedgehog pathway and pluripotency maintaining factors regulate human pancreatic cancer stem cell characteristics. Int. J. Cancer 2012, 131, 30–40.
  41. Hoca, M.; Becer, E.; Kabadayı, H.; Yücecan, S.; Vatansever, H.S. The Effect of Resveratrol and Quercetin on Epithelial-Mesenchymal Transition in Pancreatic Cancer Stem Cell. Nutr. Cancer 2020, 72, 1231–1242.
  42. Balakrishnan, S.; Mukherjee, S.; Das, S.; Bhat, F.A.; Raja Singh, P.; Patra, C.R.; Arunakaran, J. Gold nanoparticles-conjugated quercetin induces apoptosis via inhibition of EGFR/PI3K/Akt-mediated pathway in breast cancer cell lines (MCF-7 and MDA-MB-231). Cell Biochem. Funct. 2017, 35, 217–231.
  43. Lan, C.Y.; Chen, S.Y.; Kuo, C.W.; Lu, C.C.; Yen, G.C. Quercetin facilitates cell death and chemosensitivity through RAGE/PI3K/AKT/mTOR axis in human pancreatic cancer cells. J. Food Drug Anal. 2019, 27, 887–896.
  44. Psahoulia, F.H.; Drosopoulos, K.G.; Doubravska, L.; Andera, L.; Pintzas, A. Quercetin enhances TRAIL-mediated apoptosis in colon cancer cells by inducing the accumulation of death receptors in lipid rafts. Mol. Cancer Ther. 2007, 6, 2591–2599.
  45. Wei, L.; Liu, T.T.; Wang, H.H.; Hong, H.M.; Yu, A.L.; Feng, H.P.; Chang, W.W. Hsp27 participates in the maintenance of breast cancer stem cells through regulation of epithelial-mesenchymal transition and nuclear factor-κB. Breast Cancer Res. BCR 2011, 13, R101.
  46. Qin, J.J.; Li, X.; Hunt, C.; Wang, W.; Wang, H.; Zhang, R. Natural products targeting the p53-MDM2 pathway and mutant p53: Recent advances and implications in cancer medicine. Genes Dis 2018, 5, 204–219.
  47. Liu, B.; Chen, Y.; Clair, D.K.S. ROS and p53: A versatile partnership. Free Radic. Biol. Med. 2008, 44, 1529–1535.
  48. Kuo, P.C.; Liu, H.F.; Chao, J.I. Survivin and p53 modulate quercetin-induced cell growth inhibition and apoptosis in human lung carcinoma cells. J. Biol. Chem. 2004, 279, 55875–55885.
  49. Vidya Priyadarsini, R.; Senthil Murugan, R.; Maitreyi, S.; Ramalingam, K.; Karunagaran, D.; Nagini, S. The flavonoid quercetin induces cell cycle arrest and mitochondria-mediated apoptosis in human cervical cancer (HeLa) cells through p53 induction and NF-κB inhibition. Eur. J. Pharmacol. 2010, 649, 84–91.
  50. Chan, S.T.; Yang, N.C.; Huang, C.S.; Liao, J.W.; Yeh, S.L. Quercetin enhances the antitumor activity of trichostatin A through upregulation of p53 protein expression in vitro and in vivo. PLoS ONE 2013, 8, e54255.
  51. Tanigawa, S.; Fujii, M.; Hou, D.X. Stabilization of p53 is involved in quercetin-induced cell cycle arrest and apoptosis in HepG2 cells. Biosci. Biotechnol. Biochem. 2008, 72, 797–804.
  52. Lee, Y.-J.; Lee, Y.-J.; Park, I.-S.; Song, J.-H.; Oh, M.-H.; Nam, H.-S.; Cho, M.-K.; Woo, K.-M.; Lee, S.-H. Quercetin exerts preferential cytotoxic effects on malignant mesothelioma cells by inducing p53 expression, caspase-3 activation, and apoptosis. Mol. Cell. Toxicol. 2015, 11, 295–305.
  53. Kusaczuk, M.; Krętowski, R.; Naumowicz, M.; Stypułkowska, A.; Cechowska-Pasko, M. A Preliminary Study of the Effect of Quercetin on Cytotoxicity, Apoptosis, and Stress Responses in Glioblastoma Cell Lines. Int. J. Mol. Sci. 2022, 23, 1345.
  54. Bao, L.; Zhang, H.; Chan, L.S. The involvement of the JAK-STAT signaling pathway in chronic inflammatory skin disease atopic dermatitis. JAK-STAT 2013, 2, e24137.
  55. Horr, B.; Borck, H.; Thurmond, R.; Grösch, S.; Diel, F. STAT1 phosphorylation and cleavage is regulated by the histamine (H4) receptor in human atopic and non-atopic lymphocytes. Int. Immunopharmacol. 2006, 6, 1577–1585.
  56. Muthian, G.; Bright, J.J. Quercetin, a flavonoid phytoestrogen, ameliorates experimental allergic encephalomyelitis by blocking IL-12 signaling through JAK-STAT pathway in T lymphocyte. J. Clin. Immunol. 2004, 24, 542–552.
  57. Liao, Y.-R.; Lin, J.-Y.J. Quercetin, but not its metabolite quercetin-3-glucuronide, exerts prophylactic immunostimulatory activity and therapeutic antiinflammatory effects on lipopolysaccharide-treated mouse peritoneal macrophages ex vivo. J. Agric. Food Chem. 2014, 62, 2872–2880.
  58. Senggunprai, L.; Kukongviriyapan, V.; Prawan, A.; Kukongviriyapan, U. Quercetin and EGCG exhibit chemopreventive effects in cholangiocarcinoma cells via suppression of JAK/STAT signaling pathway. Phyther. Res. 2014, 28, 841–848.
  59. Matsuzawa, A.; Ichijo, H.J.B. Redox control of cell fate by MAP kinase: Physiological roles of ASK1-MAP kinase pathway in stress signaling. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2008, 1780, 1325–1336.
  60. Lee, Y.-K.; Hwang, J.-T.; Kwon, D.Y.; Surh, Y.-J.; Park, O.J. Induction of apoptosis by quercetin is mediated through AMPKα1/ASK1/p38 pathway. Cancer Lett. 2010, 292, 228–236.
  61. Kim, G.T.; Lee, S.H.; Kim, J.I.; Kim, Y.M. Quercetin regulates the sestrin 2-AMPK-p38 MAPK signaling pathway and induces apoptosis by increasing the generation of intracellular ROS in a p53-independent manner. Int. J. Mol. Med. 2014, 33, 863–869.
  62. Dutta, S.; Sadhukhan, P.; Saha, S.; Sil, P.C. Regulation of oxidative stress by different naturally occurring polyphenolic compounds: An emerging anticancer therapeutic approach. React. Oxyg. Species 2017, 3, 81–95.
  63. Arzuman, L.; Beale, P.; Chan, C.; Yu, J.Q.; Huq, F. Synergism from combinations of tris(benzimidazole) monochloroplatinum(II) chloride with capsaicin, quercetin, curcumin and cisplatin in human ovarian cancer cell lines. Anticancer. Res. 2014, 34, 5453–5464.
  64. Gibellini, L.; Pinti, M.; Nasi, M.; De Biasi, S.; Roat, E.; Bertoncelli, L.; Cossarizza, A. Interfering with ROS Metabolism in Cancer Cells: The Potential Role of Quercetin. Cancers 2010, 2, 1288–1311.
  65. Matsuzawa, A.; Nishitoh, H.; Tobiume, K.; Takeda, K.; Ichijo, H. Physiological roles of ASK1-mediated signal transduction in oxidative stress- and endoplasmic reticulum stress-induced apoptosis: Advanced findings from ASK1 knockout mice. Antioxid. Redox Signal. 2002, 4, 415–425.
  66. Watanabe, T.; Sekine, S.; Naguro, I.; Sekine, Y.; Ichijo, H. Apoptosis Signal-regulating Kinase 1 (ASK1)-p38 Pathway-dependent Cytoplasmic Translocation of the Orphan Nuclear Receptor NR4A2 Is Required for Oxidative Stress-induced Necrosis. J. Biol. Chem. 2015, 290, 10791–10803.
  67. Maurya, A.K.; Vinayak, M. Quercetin Attenuates Cell Survival, Inflammation, and Angiogenesis via Modulation of AKT Signaling in Murine T-Cell Lymphoma. Nutr. Cancer 2017, 69, 470–480.
  68. Pratheeshkumar, P.; Budhraja, A.; Son, Y.O.; Wang, X.; Zhang, Z.; Ding, S.; Wang, L.; Hitron, A.; Lee, J.C.; Xu, M.; et al. Quercetin inhibits angiogenesis mediated human prostate tumor growth by targeting VEGFR- 2 regulated AKT/mTOR/P70S6K signaling pathways. PLoS ONE 2012, 7, e47516.
  69. Xiao, Y.; Zhou, L.; Zhang, T.; Qin, C.; Wei, P.; Luo, L.; Luo, L.; Huang, G.; Chen, A.; Liu, G. Anti-fibrosis activity of quercetin attenuates rabbit tracheal stenosis via the TGF-β/AKT/mTOR signaling pathway. Life Sci. 2020, 250, 117552.
  70. Granato, M.; Rizzello, C.; Gilardini Montani, M.S.; Cuomo, L.; Vitillo, M.; Santarelli, R.; Gonnella, R.; D’Orazi, G.; Faggioni, A.; Cirone, M. Quercetin induces apoptosis and autophagy in primary effusion lymphoma cells by inhibiting PI3K/AKT/mTOR and STAT3 signaling pathways. J. Nutr. Biochem. 2017, 41, 124–136.
  71. Rivera Rivera, A.; Castillo-Pichardo, L.; Gerena, Y.; Dharmawardhane, S. Anti-Breast Cancer Potential of Quercetin via the Akt/AMPK/Mammalian Target of Rapamycin (mTOR) Signaling Cascade. PLoS ONE 2016, 11, e0157251.
  72. Li, X.; Zhou, N.; Wang, J.; Liu, Z.; Wang, X.; Zhang, Q.; Liu, Q.; Gao, L.; Wang, R. Quercetin suppresses breast cancer stem cells (CD44(+)/CD24(-)) by inhibiting the PI3K/Akt/mTOR-signaling pathway. Life Sci. 2018, 196, 56–62.
  73. Hamidullah; Kumar, R.; Saini, K.S.; Kumar, A.; Kumar, S.; Ramakrishna, E.; Maurya, R.; Konwar, R.; Chattopadhyay, N. Quercetin-6-C-β-D-glucopyranoside, natural analog of quercetin exhibits anti-prostate cancer activity by inhibiting Akt-mTOR pathway via aryl hydrocarbon receptor. Biochimie 2015, 119, 68–79.
  74. He, K.; Yu, X.; Wang, X.; Tang, L.; Cao, Y.; Xia, J.; Cheng, J. Baicalein and Ly294002 induces liver cancer cells apoptosis via regulating phosphatidyl inositol 3-kinase/Akt signaling pathway. J. Cancer Res. Ther. 2018, 14, 519.
  75. Zhao, S.; Jiang, Y.; Zhao, J.; Li, H.; Yin, X.; Wang, Y.; Xie, Y.; Chen, X.; Lu, J.; Dong, Z.; et al. Quercetin-3-methyl ether inhibits esophageal carcinogenesis by targeting the AKT/mTOR/p70S6K and MAPK pathways. Mol. Carcinog. 2018, 57, 1540–1552.
  76. Roy, S.; Banerjee, S.; Chakraborty, T. Vanadium quercetin complex attenuates mammary cancer by regulating the P53, Akt/mTOR pathway and downregulates cellular proliferation correlated with increased apoptotic events. Biometals 2018, 31, 647–671.
  77. Granato, M.; Rizzello, C.; Romeo, M.A.; Yadav, S.; Santarelli, R.; D’Orazi, G.; Faggioni, A.; Cirone, M. Concomitant reduction of c-Myc expression and PI3K/AKT/mTOR signaling by quercetin induces a strong cytotoxic effect against Burkitt’s lymphoma. Int. J. Biochem. Cell Biol. 2016, 79, 393–400.
  78. Jia, L.; Huang, S.; Yin, X.; Zan, Y.; Guo, Y.; Han, L. Quercetin suppresses the mobility of breast cancer by suppressing glycolysis through Akt-mTOR pathway mediated autophagy induction. Life Sci. 2018, 208, 123–130.
  79. Wu, H.; Pan, L.; Gao, C.; Xu, H.; Li, Y.; Zhang, L.; Ma, L.; Meng, L.; Sun, X.; Qin, H. Quercetin Inhibits the Proliferation of Glycolysis-Addicted HCC Cells by Reducing Hexokinase 2 and Akt-mTOR Pathway. Molecules 2019, 24, 1993.
  80. Wang, K.; Liu, R.; Li, J.; Mao, J.; Lei, Y.; Wu, J.; Zeng, J.; Zhang, T.; Wu, H.; Chen, L.; et al. Quercetin induces protective autophagy in gastric cancer cells: Involvement of Akt-mTOR- and hypoxia-induced factor 1α-mediated signaling. Autophagy 2011, 7, 966–978.
  81. Meng, F.D.; Li, Y.; Tian, X.; Ma, P.; Sui, C.G.; Fu, L.Y.; Jiang, Y.H. Synergistic effects of snail and quercetin on renal cell carcinoma Caki-2 by altering AKT/mTOR/ERK1/2 signaling pathways. Int. J. Clin. Exp. Pathol. 2015, 8, 6157–6168.
  82. Lou, M.; Zhang, L.N.; Ji, P.G.; Feng, F.Q.; Liu, J.H.; Yang, C.; Li, B.F.; Wang, L. Quercetin nanoparticles induced autophagy and apoptosis through AKT/ERK/Caspase-3 signaling pathway in human neuroglioma cells: In vitro and in vivo. Biomed. Pharm. 2016, 84, 1–9.
  83. Rendon-Mitchell, B.; Ochani, M.; Li, J.; Han, J.; Wang, H.; Yang, H.; Susarla, S.; Czura, C.; Mitchell, R.A.; Chen, G. IFN-γ induces high mobility group box 1 protein release partly through a TNF-dependent mechanism. J. Immunol. 2003, 170, 3890–3897.
  84. Andersson, U.; Wang, H.; Palmblad, K.; Aveberger, A.-C.; Bloom, O.; Erlandsson-Harris, H.; Janson, A.; Kokkola, R.; Zhang, M.; Yang, H.; et al. High mobility group 1 protein (HMG-1) stimulates proinflammatory cytokine synthesis in human monocytes. J. Exp. Med. 2000, 192, 565–570.
  85. Degryse, B.; Bonaldi, T.; Scaffidi, P.; Müller, S.; Resnati, M.; Sanvito, F.; Arrigoni, G.; Bianchi, M.E. The high mobility group (HMG) boxes of the nuclear protein HMG1 induce chemotaxis and cytoskeleton reorganization in rat smooth muscle cells. J. Cell Biol. 2001, 152, 1197–1206.
  86. Kim, H.; Moon, J.Y.; Ahn, K.S.; Cho, S.K. Quercetin induces mitochondrial mediated apoptosis and protective autophagy in human glioblastoma U373MG cells. Oxidative Med. Cell. Longev. 2013, 2013, 596496.
  87. Han, Y.; Yu, H.; Wang, J.; Ren, Y.; Su, X.; Shi, Y.J.H. Quercetin alleviates myocyte toxic and sensitizes anti-leukemic effect of adriamycin. Hematology 2015, 20, 276–283.
  88. Park, J.S.; Svetkauskaite, D.; He, Q.; Kim, J.-Y.; Strassheim, D.; Ishizaka, A.; Abraham, E. Involvement of toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J. Biol. Chem. 2004, 279, 7370–7377.
  89. Kokkola, R.; Andersson, Å.; Mullins, G.; Östberg, T.; Treutiger, C.J.; Arnold, B.; Nawroth, P.; Andersson, U.; Harris, R.; Harris, H. RAGE is the major receptor for the proinflammatory activity of HMGB1 in rodent macrophages. Scand. J. Immunol. 2005, 61, 1–9.
  90. Luo, J.-L.; Kamata, H.; Karin, M. IKK/NF-κB signaling: Balancing life and death–a new approach to cancer therapy. J. Clin. Investig. 2005, 115, 2625–2632.
  91. Kwak, M.-K.; Itoh, K.; Yamamoto, M.; Sutter, T.R.; Kensler, T.W. Role of transcription factor Nrf2 in the induction of hepatic phase 2 and antioxidative enzymes in vivo by the cancer chemoprotective agent, 3 H-1, 2-dithiole-3-thione. Mol. Med. 2001, 7, 135–145.
  92. Tan, X.-L.; Spivack, S.D. Dietary chemoprevention strategies for induction of phase II xenobiotic-metabolizing enzymes in lung carcinogenesis: A review. Lung Cancer 2009, 65, 129–137.
  93. Kansanen, E.; Jyrkkänen, H.-K.; Volger, O.L.; Leinonen, H.; Kivelä, A.M.; Häkkinen, S.-K.; Woodcock, S.R.; Schopfer, F.J.; Horrevoets, A.J.; Ylä-Herttuala, S.; et al. Nrf2-dependent and-independent responses to nitro-fatty acids in human endothelial cells. J. Biol. Chem. 2009, 284, 33233–33241.
  94. Taguchi, K.; Motohashi, H.; Yamamoto, M. Molecular mechanisms of the Keap1–Nrf2 pathway in stress response and cancer evolution. Genes Cells 2011, 16, 123–140.
  95. Kansanen, E.; Jyrkkänen, H.-K.; Levonen, A.-L. Activation of stress signaling pathways by electrophilic oxidized and nitrated lipids. Free Radic. Biol. Med. 2012, 52, 973–982.
  96. Dinkova-Kostova, A.T.; Holtzclaw, W.D.; Cole, R.N.; Itoh, K.; Wakabayashi, N.; Katoh, Y.; Yamamoto, M.; Talalay, P. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proc. Natl. Acad. Sci. USA 2002, 99, 11908–11913.
  97. 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, 1690–1703.
  98. Odenthal, J.; van Heumen, B.W.; Roelofs, H.M.; te Morsche, R.H.; Marian, B.; Nagengast, F.M.; Peters, W.H. The influence of curcumin, quercetin, and eicosapentaenoic acid on the expression of phase II detoxification enzymes in the intestinal cell lines HT-29, Caco-2, HuTu 80, and LT97. Nutr. Cancer 2012, 64, 856–863.
  99. Ramyaa, P.; Padma, V.V. 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 (BBA)-Gen. Subj. 2014, 1840, 681–692.
  100. Shi, Y.; Liang, X.-c.; Zhang, H.; Wu, Q.-l.; Qu, L.; Sun, Q. Quercetin protects rat dorsal root ganglion neurons against high glucose-induced injury in vitro through Nrf-2/HO-1 activation and NF-κB inhibition. Acta Pharmacol. Sin. 2013, 34, 1140–1148.
  101. Lee, Y.-J.; Song, J.-H.; Oh, M.-H.; Lee, Y.-J.; Kim, Y.-B.; Im, J.-H.; Lee, S.-H. ERK1/2 activation in quercetin-treated BEAS-2B cell plays a role in Nrf2-driven HO-1 expression. Mol. Cell. Toxicol. 2011, 7, 347–355.
  102. Chow, J.-M.; Shen, S.-C.; Huan, S.K.; Lin, H.-Y.; Chen, Y.-C. Quercetin, but not rutin and quercitrin, prevention of H2O2-induced apoptosis via anti-oxidant activity and heme oxygenase 1 gene expression in macrophages. Biochem. Pharmacol. 2005, 69, 1839–1851.
  103. Yao, P.; Nussler, A.; Liu, L.; Hao, L.; Song, F.; Schirmeier, A.; Nussler, N. Quercetin protects human hepatocytes from ethanol-derived oxidative stress by inducing heme oxygenase-1 via the MAPK/Nrf2 pathways. J. Hepatol. 2007, 47, 253–261.
More
Video Production Service