Next Article in Journal
Laherradurin Inhibits Colorectal Cancer Cell Growth by Induction of Mitochondrial Dysfunction and Autophagy Induction
Previous Article in Journal
Effects of Tryptophan and Physical Exercise on the Modulation of Mechanical Hypersensitivity in a Fibromyalgia-like Model in Female Rats
Previous Article in Special Issue
Identification of Poliovirus Receptor-like 3 Protein as a Prognostic Factor in Triple-Negative Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vulnerability of Antioxidant Drug Therapies on Targeting the Nrf2-Trp53-Jdp2 Axis in Controlling Tumorigenesis

1
School of Dentistry, College of Dental Medicine, Kaohsiung Medical University, Kaohsiung 807, Taiwan
2
Graduate Institute of Medicine, Kaohsiung Medical University, Kaohsiung 807, Taiwan
3
Regenerative Medicine and Cell Therapy Research Center, Kaohsiung Medical University, Kaohsiung 807, Taiwan
4
Cell Therapy and Research Center, Kaohsiung Medical University Hospital, Kaohsiung 807, Taiwan
5
Division of Gastroenterology, Department of Internal Medicine, Kaohsiung Medical University Hospital, Kaohsiung 807, Taiwan
*
Authors to whom correspondence should be addressed.
Cells 2024, 13(19), 1648; https://doi.org/10.3390/cells13191648
Submission received: 15 August 2024 / Revised: 23 September 2024 / Accepted: 26 September 2024 / Published: 3 October 2024
(This article belongs to the Special Issue Molecular Mechanisms of Tumor Pathogenesis)

Abstract

:
Control of oxidation/antioxidation homeostasis is important for cellular protective functions, and disruption of the antioxidation balance by exogenous and endogenous ligands can lead to profound pathological consequences of cancerous commitment within cells. Although cancers are sensitive to antioxidation drugs, these drugs are sometimes associated with problems including tumor resistance or dose-limiting toxicity in host animals and patients. These problems are often caused by the imbalance between the levels of oxidative stress-induced reactive oxygen species (ROS) and the redox efficacy of antioxidants. Increased ROS levels, because of abnormal function, including metabolic abnormality and signaling aberrations, can promote tumorigenesis and the progression of malignancy, which are generated by genome mutations and activation of proto-oncogene signaling. This hypothesis is supported by various experiments showing that the balance of oxidative stress and redox control is important for cancer therapy. Although many antioxidant drugs exhibit therapeutic potential, there is a heterogeneity of antioxidation functions, including cell growth, cell survival, invasion abilities, and tumor formation, as well as the expression of marker genes including tumor suppressor proteins, cell cycle regulators, nuclear factor erythroid 2-related factor 2, and Jun dimerization protein 2; their effectiveness in cancer remains unproven. Here, we summarize the rationale for the use of antioxidative drugs in preclinical and clinical antioxidant therapy of cancer, and recent advances in this area using cancer cells and their organoids, including the targeting of ROS homeostasis.

1. Introduction

It is known that the control of balance between reactive oxygen species (ROS) and redox status is an important way to address the occurrence and development of tumor cells without compromising normal cells [1]. This idea depends upon the theory that cancers have a homeostatic balance of oxidation, which is produced by phase I enzyme ligands, and antioxidation, which is mediated by phase II enzyme ligands and is arranged to favor the hallmark actions of normal cells or cancer cells, such as proliferation, survival, migration, and metastasis (Figure 1) [2]. Thus, the ROS balance is critical for regulating oxidative stress and antioxidation and mediated by transcription factors such as the aryl hydrocarbon receptor (AHR) and nuclear factor erythroid 2-related factor 2 (NRF2), respectively. The extensive redox conditions damage biocomponents, and cancers are known to compensate for this damage by enhancing the expression of antioxidants. Thus, the development of cancers depends on these antioxidant proteins to regulate the oxidation and antioxidation status of macromolecules within cells to maintain the ROS balance. A switch in cancer redox activity produced by antioxidation or increased ROS production causes the threshold of the ROS balance of cancer cells to be altered, which causes cell cycle arrest or cell death [3]. Redox-modulating drugs that cause apoptosis are important in defining ROS-induced cell death, which is important for anticancer therapy [3]. AHR and NRF2 are critical factors for protection against environmental stresses. The phase I enzyme system facilitates the oxidative reduction of xenobiotics, which converts intermediate metabolites, including electrophiles. The phase II enzyme system involves conjugating a hydrophilic moiety (i.e., glucuronate, sulfate, glutathione, or glycine) to phase I metabolites. This reaction involves enzymes, such as transferases, and the phase I metabolites are subsequently transformed into water-soluble molecules. Thus, these metabolites or AHR-induced ROS can trigger the phase II enzyme system to reduce the endogenous ROS using the NRF2-KEAP1 (Kelch-like ECH associated protein 1) system to maintain the endogenous ROS balance within the cells. AHR controls phase I enzymes, including the CYP450 family (i.e., CYP1A1, CYP1A2, and CYP1B1), nitrogen oxides (NOXs), cyclooxygenase (COX), and aldo-keto reductase (AKR), while NRF2 controls phase II enzymes, such as glutathione S-transferase α1 (GSTA1), glutathione S-transferase π1 (GSTP1), and the uridine diphosphate glucuronosyltransferase family (UGTs). The interaction between the AHR and NRF2 pathways is essential for cellular protection against the toxic byproducts of AHR. The oxidative reactive species (ROS) are toxic byproducts of aerobic metabolism that are harmful to cells and cause oxidative damage to DNA, proteins, and lipids, which might lead to cell death. ROS production is compensated by reversible reactions, such as antioxidation, to maintain a balance of ROS levels. An elevated level of ROS is detrimental to the control of oxygen toxicity, whereas physiologically low levels of ROS induced by antioxidative mechanisms serve as intracellular signaling mechanisms. Thus, the endogenous control of ROS homeostasis is a cellular defense mechanism against toxicants and oxidative stress (Figure 1).
Although drugs that control ROS are expected to have therapeutic applications in cancer, drugs targeting ROS have shown only limited success in preclinical trials compared with other anticancer drugs [4]. These results raise the possibility that tumor cells may not be more sensitive to ROS than normal cells. Thus, these trials produced less success than anticipated. The lack of biomarkers to measure endogenous redox levels might also prevent success [5]. Furthermore, the complexity and redundancy of drugs targeting ROS/redox are not well characterized. Recent reports have shown the heterogeneity of phase II enzyme reagents against cancer organoids [6]. Thus, the use of antioxidants for cancer treatment should be approached with caution. In this review, the heterogeneity of phase II enzyme agents is discussed in the context of cancer treatment and how to avoid these difficulties in preclinical trials of phase II drugs for cancer treatment.

2. ROS in Cancer Cells

Assuming a crucial role for ROS in cancer cells, controlling endogenous ROS by altering the function of mitochondrial constituents is a possible anticancer strategy. These strategies may inhibit ROS-mediated cancer occurrence and progression by causing oxidative damage, such as ROS-mediated apoptosis [7,8]. Thus, preclinical research on antioxidant production and weak pro-oxidants was conducted to understand their merits. Compared with normal cells, cancer cells generate excessive ROS, increasing their sensitivity to further increases in ROS-related cell damage and committing them to apoptosis. Pro-oxidants may thus have antitumor functions. Therapeutic trials of antioxidants targeting ROS control have included nonenzymatic drugs, such as NRF2 agonists [9] and vitamins [10,11], or targeting ROS via the enzymatic production of antioxidants, such as nitrogen oxide inhibitors [12,13], superoxide dismutase mimetics [14], N-acetylcysteine, and glutathione (GSH) esters [15,16]. Many such pre and clinical trials have been conducted (Table 1), and many such agents against ROS production have been developed, but their actions involve ROS homeostasis and antioxidation, which are not specific to cancer.
For example, the repression of glutathione peroxidase 2 (GPx2) in breast cancer may enhance cancer progression due to hypoxic signals, aberrant vascularization, and a metabolic switch to the aerobic glycolysis/oxidative phosphorylation (OXPHOS) axis. The complex expression of GPx2 depends on metabolic-specific drives and microenvironments, such as hyperproliferation-derived hypoxia.
Another consideration is that ROS have a dual function in cancer occurrence, which might depend on endogenous ROS (Figure 1) [17].
Table 1. The variation of redox levels in different cancer cells. Various effects of antioxidant functions are also shown.
Table 1. The variation of redox levels in different cancer cells. Various effects of antioxidant functions are also shown.
Tumor TypeRedox LevelEffectsReferences
Urinary tract & bladder
  • GPx2 ↓
  • Poor prognosis patient samples
[18]
Esophageal carcinoma
  • GPx2 ↓
  • Poor prognosis patient samples
[19]
Early-stage of lung adeno carcinoma (LuAD)
  • GPx2, ALDH3A, GLRX, PPK ↓
  • Increase ROS dependent caspase activity
  • Aggressive phenotype in early stage of hypoxic tumor
[20]
T1 Bladder carcinoma
  • pT1 stage
  • GPx2 ↓
  • Early-phase invasion
[21]
Breast cancer
  • GPx2 ↓
  • Tumor heterogeneity
  • Metabolic plasticity
  • Malignant progression
  • ROS/HIF1/VEGFA pathways activation
[22]
Pancreas cancer
  • Proliferation ↓
  • Invasion ↓
  • Metastasis ↓
[23]
Hepatocellular carcinoma
  • Poor prognosis
[24]
Gastric carcinoma
  • Markers of primary tumor and metastatic form
[25]
Lung carcinoma
  • Tumor initiation
  • Cisplatin resistance
[26]
Lung adenocarcinoma
  • Tumor formation ↑
  • Independent prognostic factor of recurrence free survival (RFS)
[27]
Non-small cell lung adenocarcinoma
  • Gpx2 ↑: as an oncogene by inhibit of apoptosis, promoting MET, glc uptakes
  • BMP ↓: as an Tumor suppressor via inducing apoptosis cell cycle arrest
[28]
Prostate cancer
  • PC2, RFS prognostic model ↑
  • Wnt/β-catenin/EMT pathway ↑
[29]
kRas mutated lung cancer cells
  • Malignant progression
  • Cisplatin resistance
[30]
Colon cancer
  • Microsatellite instability ↑
[31]
Gastric cancer
  • Patient survival, Tumor progression in ROS-KYNU-KYN-AhR signal in GPx2 knock down cells
[32]
NSCLC
(Non-small cell lung cancer)
  • EMT metastasis of NSCLC by PI3K/Akt/mTOR/snail axis to remove ROS
  • Prognosis is OK
[33]
NSCLC
  • Breast cancer metastasis suppressor 1 like (BRMS1L) ↓: is associate with cancer growth
  • Gpx2 mediated oxidative stress repair ↓
  • BRMS1L knockdown characterized NSCLC cells to ROS induce
[34]
Excessive ROS production leads to a defective antioxidant defense mechanism and thereby impairs the balance between antioxidants and pro-oxidants. Recent reports revealed the dichotomous nature of ROS in tumor cells, depending on the stage of cancer progression (early stage or late stage) or differences in endogenous ROS levels. Elevated ROS production in cancer cells initiates adaptation, which maintains the ROS balance by antioxidation reactions. In precancerous stages or early stages of cancer progression, moderate ROS levels induced tumorigenesis, tumor promulgation, metastasis, and survival [35]. With tumor progression, ROS levels elevated beyond a toxic threshold lead to cell death, apoptosis, and senescence [36,37].
This dichotomy further complicates the antioxidation response to the decomposition of H2O2 through a change in the glutathione-SH (GSH)–glutathione sulfide (GSSG) axis. Hypoxia-inducible factor-1β, also known as AHR nuclear translocator (ARNT), is a component of the complex of phase I enzymes with AHR. Thus, crosstalk between AHR signals and hypoxia signals might occur, leading to overlap. In addition, the development of cancers also involves contrasting levels of GPx2 expression, which are correlated with the progression of cancer and poor prognosis. Decreased GPx2 expression was detected in colon cancer, pancreatic cancer, cancers of the bladder and urinary tract, and esophageal colon carcinomas, whereas increased GPx2 expression was detected in several carcinomas [38]. This apparent contradiction of GPx2 levels might be due to differences in the levels of the intact endogenous ROS production machinery [39].

3. NRF2 Has Dual Roles in Tumorigenesis

NRF2 has dual and contradictory roles in cancer [39]. Abnormal increased expression of NRF2 can be related to a poor prognosis. The constitutive level of NRF2 in many cancer cells might induce the expression of pro-survival genes and promote the proliferation of tumor cells via metabolic alterations, inhibit cancer specific-apoptosis, and increase the self-renewal function of cancer stem cells. Moreover, NRF2 contributes to chemoresistance, radio-resistance, and inflammation, inducing carcinogenesis. Many NRF2 inhibitors have been reported to treat cancers [40]. These trials targeting NRF2 might provide a new tool in cancer treatment.

3.1. Tumor Suppressive Actions of NRF2

The NRF2 protein is activated by tumor suppressor gene products such as breast cancer susceptibility gene I and p21Cip1 by inhibiting KEAP1/NRF2 complex formation in microenvironmental cells in tumors [41,42]. The degradation of the oncogene tyrosine kinase Fyn blocks its activation [43]. Nrf2-deficient mice reduced the production of antioxidants and repressed the expression of markers specific for phase II enzyme-encoded genes. Quinone induced severe oxidative stress rendered Nrf2-deficient mice more prone to skin cancer, while the Nrf2-mediated expressions of NAD(P)H quinone dehydrogenase 1 (Nqo1) and Gst were inhibited [44]. Similarly, Nrf2-deficient mice exposed to carcinogens developed more cancers, including stomach [45], liver [46], and urinary bladder [47], as compared with wild type mice, suggesting that Nrf2 is the critical factor controlling inflammation.
Many compounds isolated from plants, including carnosol, curcumin, resveratrol, and sulforaphane, and synthetic chemicals like oltipraz and oleanane triterpenoids, exhibit chemo-preventive functions by activating the Nrf2-antioxidant response element (ARE)-regulated genes. These Nrf2 activators mainly control the antioxidation reaction by changing the intermolecular disulfide (S–S) bonds between two cystine residues of Keap1 proteins at Cys273 and Cys288 to commit the accumulation of Nrf2 [48]. Sulforaphane (SFN) is committed to inducing the transcription of phase II enzyme-encoding genes while inhibiting phase I enzyme-encoding genes and facilitating the cell death of cancer cells through the TP53 mechanism [49]. In addition, SFN targeted nuclear factor kappa B (NF-κB) [50] and activation protein-1 (AP-1) complexes with an anti-inflammatory effect [51]. Other synthetic oleanane triterpenoids were found to inhibit tumorigenesis by inhibiting the transcription of oncogenes, including k-RAS, TP53, BRCA1, and Erb-B2 receptor tyrosine kinase 2, in some cancers [52,53].
A lot of therapeutics that control NRF2 activity have been currently applied in clinical trials; these drugs play a significant role in repressing cancers and could be augmented further by chemo-preventive compounds. Paradoxically, NRF2-deficient cancer cells were found to be more prone to the death of cancerous cells by oxidative stress but more resistant to chemo-preventive compounds. Thus, effective drugs targeting NRF2 pathways represent a crucial strategy to develop useful chemo-preventive medications. However, we need to accumulate further pre-clinical data to prove the efficiency of these drugs against cancer in human patients.
In Nrf2-knockout mice, the level of nitric oxide synthase activated by cyclooxygenase 2 and the level of tumor necrosis factor (Tnf) are greater than those in normal mice, indicating that Nrf2 is able to suppress the proinflammatory molecules [54]. In addition, Nrf2-dependent activation of Nqo1 decreased the levels of Tnf and interleukin 1 after incubation with lipopolysaccharide. Although ROS are the molecular targets of NRF2-mediated anti-inflammatory effects, NRF2 might function as an anti-inflammation mediator without inducing ROS, which was reported by regulating the expressions of the axis encoding the macrophage receptor with collagenous structure and CD36, which are not concerned with the oxidative response [55]. Furthermore, NRF2 protected against cell damage induced by H2O2 exposure via the p38/MAPK pathway [56,57]. Similarly, NRF2 inhibited the NF-κB pathway by stabilizing the inhibitor of nuclear factor κB kinase subunit-α (IKK-α) to inhibit the degradation of IKK-β [58]. By contrast, the NF-κB p65 subunit was reported to compete with NRF2 through the CH-KIK domain of the cyclic AMP response element binding protein (CREB) binding protein CBP coactivator [59]. Thus, these different mechanistic inputs of NRF2 should be clarified to determine how it suppresses oncogenicity by controlling the level of ROS within cells.

3.2. Oncogenic Functions of NRF2

NRF2 is usually constitutively activated in tumorigenic cells. For example, somatic mutations in KEAP1 and/or NRF2 genes, exon skipping in the NRF2 genome, methylation/demethylation of the KEAP1 promoter, accumulation of p62/sequestosome-1, and mutation of fumarate hydrolase have been reported [60,61] Constitutively activated NRF2 promoted cell proliferation through metabolic changes; inhibiting apoptosis; promoting angiogenesis, and invasion/metastasis; and promoting drug resistance in various cancers has also been reported [62,63]. Forced expression of NRF2 promoted the transcriptions of the oncogenes MYC, KRAS, and BRAF [64]. Moreover, NRF2 activated oncogenes by suppressing the activity of tumor suppressor gene products like phosphatase tensin homolog/glycogen synthetase kinase 3/β-transduction repeat-containing E3 ubiquitin-protein ligase [63]. NRF2 increased the proliferation of tumor cells by increasing the expression of glycolytic enzymes including glucose-6-phosphate dehydrogenase, phosphoglucomutase dehydrogenase, transketolase and trans-aldolase [64], which control genes involved in fatty acid-lipid metabolism [65], growth-associated genes [66], and cell cycle regulators [67]. The activation of NRF2 via the glucose-regulated protein 78/phosphorylated protein kinase RNA-like ER kinase/NRF2 signaling pathway enhanced the transcriptions of glycolytic genes and simultaneously induced the inhibition of the TCA cycle related genes, which enhanced the Warburg effect [68].
Increased levels of NRF2 suppressed the endogenous ROS level in cancer stem cells (CSCs) compared with noncancer stem cells, suggesting the enrichment of stemness phenotypes [69,70,71,72]. For example, CSCs with reduced mitochondrial-derived ROS exhibited cancer stemness-associated properties, such as epithelial-to-mesenchymal transition (EMT) [68,73]. Similarly, persistent activation of NRF2 improved the self-renewal ability of CSCs by maintaining cell quiescence and reducing the intracellular levels of ROS [74,75]. Mesenchymal stem cells (MSCs) in the microenvironment cells promoted the spread of cancer cells by enhancing their motility and invasion ability [76,77]. Thus, NRF2 was required to maintain the stemness of MSCs and causing their cell deaths under an oxidative stress condition [78]. Furthermore, because the continuous expression of NRF2 significantly benefited the cancer cells, these cells frequently developed NRF2 dependency [79,80].
In addition, NRF2 -derived oncogenic activity also stimulated angiogenesis capacity, mainly by activating heme oxygenase 1 (HO-1) [81], which in turn stimulated the vascular endothelial growth factor to enhance angiogenesis [82]. siRNA against NRF2 downregulated HO-1 and sensitized acute myeloid leukemia cells to TNF-induced cell death. This effect indicates that NRF2 inhibits the apoptosis of cancer cells by regulating HO-1 [83]. In addition, NRF2 upregulated the antiapoptotic protein B-cell lymphoma 2 (Bcl2), while it downregulated the activity of BAX and caspase 3 to protect against etoposide/radiation-mediated cell death following drug resistance [84]. Furthermore, NRF2 inhibited the activation of proapoptotic JNKs [85] and induced selective autophagy reactions via Keap1 degradation [86,87]. However, the forced expression of NRF2 allowed autophagy-dependent cancer cells to overcome the loss of autophagy function [88].
Increased accumulation of NRF2 in the nucleus was associated with proliferation capacity. For instance, phosphoinositide 3-kinase (PI3K)-Akt activation coupled with depletion of Keap1 resulted in Nrf2-dependent proliferation of hepatocytes and cholangiocytes [89,90]. However, because the ability of stable Nrf2 was not sufficient to alter it from a cancer defender to a cancer driver, the additional mutations of oncogenes and tumor suppressor genes were essential [91,92,93]. Keap1 mutations, such as Kras/Hras mutations and Trp53 loss of function, are required to produce NRF2-dependent cancer models [94]. Furthermore, NRF2-dependent malignancies with somatic mutations of KEAP1 and NRF2 differ depending on the specific organs. Thus, tissue-specific variation is another issue for the study of NRF2-dependent cancer, which should be discussed further.
In resistance to medical therapy, NRF2-controlled drug efflux transporters are predictors of resistance for drug inoculations. For example, multidrug resistance protein 1 (MDR1), multidrug resistance-associated proteins 1-5 (MRP1-5), and breast cancer-resistant protein (BCRP) are critical for controlling issues to regulate the therapy because they induced the abnormal NRF2 activation to induce the widespread chemoresistance [95,96,97,98]. These issues might comprise complementary trials for drug delivery to cancer cells.

4. NRF2-Targetted Drugs in Cancer Prevention

In the initiation stage of cancer, NRF2 activation can suppress carcinogenesis. Under normal conditions, NRF2 maintains antioxidation homeostasis and exerts anti-inflammatory effects and antitumorigenic effects, thus supporting cell survival (Figure 2). NRF2 can control the transcription of phase II enzyme genes and activate a series of defective mechanisms through the Keap1/Nrf2/ARE axis, maintaining cellular redox homeostasis. NRF2′s anti-inflammatory and anticancer activities avoid cell damage, thus benefiting the survival of normal cells. In this respect, NRF2 activation is critical for cancer prevention. By contrast, the overactivation of NRF2 also protects cancer cells and promotes their growth. Excessive NRF2 signaling plays a tumor-promoting role, mainly by maintaining the proliferation signals, infinite replication, continuous angiogenesis, resistance to apoptosis and ferroptosis escaped from immune destruction, and promotion of invasion and migration [43,48].

Heterogeneity of Antioxidation Drugs against Cancer

Wu et al. (2023) recently reported the heterogeneity of redox ligands controlling the growth and progression of human gastric cancer organoids [6]. They performed experiments to compare the efficiency of three drugs, perillaldehyde (PEA) and cinnamaldehyde (CE), and the regular antioxidative drug sulforaphane (1-isothiocyanato-4-(methane sulfinyl) butane; SFN), and measured cancer occurrence by assessing xenotransplantation in SCID mice. PEA extracted from Perilla frutescens [99] showed antifungal, and antioxidant activities in addition to other biological functions [100,101,102]. PEA activates NRF2 and represses oxidative stress-induced innate immunity in human keratinocytes [103]. CIN is a β-unsaturated aldehyde from cinnamon [104]. In addition, the antioxidant activity of CIN was demonstrated in mice [105]. Exposure to CIN-induced autophagy-dependent cell death through epigenetic alteration by the histone methylating enzyme G9a in the endoplasmic reticulum [106]. CIN inhibited the AHR axis and induced an NRF2-mediated antioxidation response [107]. Thus, PEA and CIN might function in NRF2-dependent antioxidation to inhibit ROS generation without affecting AHR-mediated oxidative stress pathway. SFN, as a dietary isothiocyanate synthesized from a precursor in Brassica, was also examined. It is one of the most typical ligands of phase II enzymes pathway, including NQO1, GST α-1, and HO-1, which are required to eliminate chemically damaged DNA. SFN induced cell cycle arrest at S-phase and apoptosis in a TP53-dependent manner in gastric cancer (GC) cells [108]. In addition, SFN inhibited the metastasis of breast cancer [109].
Another issue is the differential outputs on function of AHR- and NRF2-mediated phase I and II pathways. Both PEA and CIN induced an NRF2-mediated phase II enzyme response, which was mediated by ARE-containing genes [109] but was not involved in the phase I pathway [107]. By contrast, the effects of SFN were reported to be involved in both enzymatic pathways; SFN produced NRF2-dependent phase II promoter activation and AHR-dependent activation of phase II promoters such as NQO1 [110,111]. This activation provides unambiguous evidence of the different actions of PEA/CIN and SFN in preventing cancer-inducing activity. It is surprising to observe the contrasting actions of PEA/CIN and SFN in inhibiting gastric cancer development, ROS generation, and cellular apoptosis. TP53 mutation might explain these differential effects because 3-D organoids exhibited TP53 mutations, which might alter the phase I and phase II enzyme reactions (Figure 3).

5. Phase I Drugs in Clinical Trials

Phase I and II enzyme pathways are known to overlap and are regulated by the “ARF–NRF2” gene battery [112]. In addition, the AP1/ATF transcription factor and histone chaperone Jun dimerization protein 2 (Jdp2) were found to play a role in regulating AHR promoter activity via the NRF2 complex [113]. This mechanism was named the AHR–NRF2–JDP2 gene battery.
In the case of AHR therapeutics, the present application of AHR antagonists in cancer disorders is a reasonable strategy for preclinical trials. However, although many trials of AHR-targeted drugs and NRF2-target genes for cancer or disease have been conducted, most of them did not result in completed FDA approvals. The current trials are listed (Table 2) [114,115]. AHR antagonists inhibit immune suppression, and AHR agonists overcome chronic inflammation and autoimmune disorders in cancer patients. The AHR antagonist IK-175 is well known to be effective in an anticancer therapeutic trial when combined with anti-PD1 antibodies. A phase Ia/b open-label study of OK175, alone or in combination with nivolumab, which suppressed locally advanced solid tumors and urothelial carcinoma, has been reported (NCT04200963) [34]. The microbial metabolite indole-3-aldehyde (3-IAld) exhibited strong anti-inflammatory activity via AHR [116,117]. Compelling results with 3-IAld were obtained in a dextran sodium sulfate (DSS)-mediated inflammatory bowel disease of mouse model. 3-IAld repaired colon damage and improved epithelial barrier integrity through AHR, and IL22 cured immune checkpoint inhibitor (ICI)-induced colitis but did not block the antitumor effect of ICI [118]. These data indicate that the AHR ligand of 3-IAld could be developed as an agent for treating human disease.
BAY 2416964 is a novel oral AhR inhibitor that prevented AHR ligand-induced immunosuppressive effects and enhanced the proinflammatory activity of antigen-in-human phase I clinical trials [119]. Studies in vitro showed that BAY 2416964 restored immunological activity in human and mouse cells, stimulated antigen-specific cytotoxic T-cell responses, and killed tumors. Studies in vivo showed that its oral application was well tolerated and demonstrated antitumor activity in a syngeneic mouse cancer model.

6. Phase II Drugs in Clinical Trials

Clinical current trials using phase II enzyme drugs have been reported (Table 2) [124,148,149,150,151].
For example, SFN has achieved only limited success in prostate cancer patients [152]. However, SFN is not effective for treating breast cancer patients [153]. SFN inhibited the progression of GC [108,154,155,156,157]. Thus, further studies are required to determine the usefulness of SFN.
In addition, the mutation status of oncogenes and antioncogenes is also crucial for the efficacy of treatment with phase II enzyme drugs. The expressions of TP53, NRF2, and JDP2 were tremendously repressed in the growth of organoids exposed to PEA and CIN as compared with those in control normal organoids. However, a 1.5- to 1.75-fold increase in NRF2 and TP53 expression was found in SFN-treated organoids compared with control organoids. Exposure to PEA or CIN repressed the expression of proteins of the TP53–NRF2 axis and a third factor histone chaperone JDP2, but SFN increased the expression of TP53 and NRF2 proteins. In the case of a tumor, the p53 mutation itself did not induce the tumor formation, but tumors developing from areas with p53 mutation and loss of heterozygosity were larger and demonstrated extensive chromosomal instability compared with lesions arising in normal epithelium [158]. Most TP53 mutations in cancers are found as missense mutations rather than truncations or deletions. Both dominant-negative effects and gain-of-function activities were observed in the case of mutant p53 [159]. Mutant TP53 interacts with NRF2, but both positive and negative effects of NRF2 have been reported [160,161,162]. Non-small cell lung cancers carrying mutant TP53 but not mutant NRF2 or KEAPl displayed higher levels of NRF2 mRNA than wild-type TP53 tumors [163]. Similarly, the oncogenes KRAS, BRAF, and MYC promoted the increased transcription of NRF2 gene and its target genes, which might induce a greater decrease in cellular molecules [62]. Mutant p53 prolonged TNFα-induced NF-κB signaling [164], and mutation of p53 can upregulate Nrf2 via the NF-κB axis. However, in 3-D gastric cancer organoids, PEA and CIN treatment reduced the protein expression of TP53 and Nrf2 and reduced ROS and caspase 3 activity [6].
By contrast, SFN exposure induced NRF2 and other redox functions to be dominant. Some reports have indicated that the levels of phase II enzymes, such as NQO1, are increased in cancer tissues compared with healthy tissues and that NQO1 stabilized the wild-type TP53, especially under oxidative stress [165]. Compared with those of wild-type TP53, the TP53 mutants exhibited increased binding to NQO1 [166]. By contrast, the function of NRF2 as an antioxidation response factor was blocked in the case of R273H in p53-expressing cancer cells upon oxidative stress, and the NRF2 antioxidant response was impaired because of the decreased expression of NQO1 and HO-1 [167].
Mutant p53 enhanced the Warburg effect by activating the glucose transporters GLUT5/6 and GLUT3 via the NF-κB axis and GLUT1 by stimulating its transfer to the plasma membrane. Various glycolytic enzymes are also stimulated by mutant p53, which caused increased glucose uptake and glycolysis [168]. The effect of the microenvironment on p53 mutations can also affect cancer generation.
One TP53 mutant, APR-246, has been tested in clinical applications [169]. The APR-246 mutant inhibited thioredoxin reductase 1 (TrxR1) [170], thioredoxin (Trx), glutaredoxin, and ribonucleotide reductase [171] and depleted cellular GSH [172,173]. Recent studies revealed that APR-246 was not specific to cancer cells, but its effects might be cell-type specific [174].
In general, the expression of p53 counteracted the expression of ARE-containing antioxidant genes such as cystine/glutamate transporter (SLC7a11; x-CT), NQO1, and GST-α1, which are Nrf2 targets [175]. p53 deletion increased ROS, DNA oxidation, and mutations. The introduction of dietary supplementation with the antioxidant N-acetylcysteine subsequently improved karyotype stability and prevented the early onset of mutations for tumorigenesis. Three p53 mutations, K117R, K161R, and K162R, were generated and resulted in impaired p53-mediated cell cycle arrest, senescence, and apoptosis. Unlike p53-knockout mice, these p53 mutant mice did not develop early-stage lymphoma [176,177]. KRAS depletion and the depletion of RalB and IκB-related TANK-binding kinase 2 (TBK1) induced activation of p53 in a ROS- and NRF2-dependent manner. Similarly, the IκB kinase inhibitor BAY 11-7085 and dominant-negative mutant IκB-αM inhibited NF-κB activity and increased the levels of phosphorylated p53, p53, and p21Cip1 in a ROS-dependent manner [178]. The p53–Mdm2 protooncogene (MDM2)–ARF network can lead to unconventional and unique innovative approaches for developing a new generation of genetically informed and clinically effective cancer therapies [179].
Wild-type TP53 repressed Jdp2 promoter activity, but TP53-knockout mutant cells did not, which was demonstrated using HCT116 p53−/− cells [180]. JDP2 enhances the NRF2-dependent antioxidant response [181]. JDP2 functions not only as a transcription factor but also as a histone chaperone for targeting specific acetylated histones. In addition, it has inhibitory effects on p300-mediated histone acetylation [182]. Therefore, the expression of TP53 and JDP2 was reported to be coregulated during transcription and chromatin regulation. These alterations might influence the function of NRF2 in TP53-mutant cancer cells or organoids. Thus, the molecular interactions of mutated TP53, JDP2, and NRF2 should be clarified.
Therefore, the contradictory results with PEA or CIN and SFN should be studied at the molecular level to identify the distinct pathways involved in human GC.

7. Discussion

This review highlights the role of the transcriptional balance of key factors between cellular antioxidation and oxidation in cancer development. The dedicated interplay to control NRF2 and AHR master regulators of ROS homeostasis should be defined as the context-dependent nature of NRF2/AHR actions in each cancer case. Ingenuity pathway analysis (IPA) of the signal networks between AHR and NRF2 revealed that they connected to shared target genes, such as genes related to the cell cycle, and chromatin regulators, including cyclin-dependent kinase inhibitor 1 (CDKN1A; p21CIP1) and EP300 (histone acetylase p300) (Figure 4). JDP2, such as CDKN1A, has been reported to play a role in cell cycle regulation by controlling cell growth via cyclin A2 [183] and p300/CBP acetyltransferase to function as a HAT on histones [182].
Given the complexity of antioxidation mechanisms in different cancer patients, the cancer types, and stages of cancer progression, manipulating redox status may be inappropriate. Thus, the precise regulation of redox balance is important and has become a key target in searching for the next generation of redox-regulating agents for cancer therapy. Repeating the examples to perform experimental validation would provide a further understanding of redox mechanisms in cancer initiation and progression.
A variety of responses to NRF2-mediated cancer development, which are obtained by 2-D cell culture, should be reexamined using 3-D organoids. These responses might be due to the intratumor heterogeneity of CSCs and to committed subpopulations with distinct microenvironments. The following questions should be addressed: (i) Do antioxidant exposures, as cancer treatments, demonstrate the positive and negative dual effects in cancer types, cancer stages or an endogenous ROS-dependent manner? (ii) Are there specific treatments with different doses in different tumor tissues that need treatment with antioxidant therapies for the required clinical outcomes? (iii) Is JDP2–NRF2-induced metabolic reprogramming in response to phase II enzyme ligands required for CSCs or niches that are dependent upon the status of TP53 mutation? Such questions are currently ongoing. The techniques used to measure endogenous ROS levels in vivo and in 3-D organoid tips might be the next targets for applying redox drug therapy for cancer treatment, in addition to further identifying markers for CSCs and microenvironments. These technologies might validate antioxidant therapy for tumors. Thus, we need more experiments involving mechanistic analysis of phase II drugs against cancer development.
Many preclinical trials of phase I and II agonists and antagonists have been conducted to date; however, only limited clinical trials have been conducted to evaluate the efficacy of antioxidative stress agents and antioxidants. The clinical and preclinical trials of phase I and II enzyme agents were performed to examine their efficacies for disease and cancer. However, none of them produced satisfactory outcomes in cancer therapy.
We have listed current applications [124,148,149,150,151]. N-acetylcysteine has been applied in various diseases, including cancers, and pulmonary, renal, liver (non-alcoholic fatty liver) diseases, and infections, including with human immunodeficiency virus, obesity, Parkinson’s disease, and depressive disorders. Melatonin has also been used in many diseases, including necrotizing enterocolitis, multiple sclerosis, and septic shock. Curcumin has been tested for a wide variety of conditions, including renal transplantation disorder, coronary artery disease, metabolic syndrome, and kidney disease (Table 2). These data were collected from the website http://clinicaltrials.gov (accessed on 28 August 2024) using the keywords antioxidative stress, antioxidation, diseases, drug treatments, and nutrients, and included ongoing and completed clinical trials.

8. Conclusions

Antioxidant drugs targeting phase II enzyme pathways have been extensively developed to focus on controlling ROS balance in patients with disease or cancer. The phase II enzyme regulating transcription factor NRF2 plays a critical role in modulating cancer development against oxidative stresses and the somatic mutations of genes and epigenetics. More than 600 genes, about 200 of which encode cytoprotective proteins, are involved in disease and cancer. In conclusion, modulating NRF2 activity is a promising approach to achieving homeostasis during every stage of tumorigenesis, where oxidative stress is an essential player. Nevertheless, a crucial strategy for cancer therapy is a better understanding of how activation or inhibition of NRF2-AHR pathways and ROS control functions. The cell specificity, the stage of cancer development, and the expression of genes encoding TP53, p21Cip1, NRF2, and JDP2 should be clarified. In addition, whether antioxidant drugs could affect NRF2 expression only or the AHRNRF2JDP2 battery is also a critical issue. Thus, a deeper understanding of the role of antioxidants in cancer is required for developing new drugs [17,151,184]. The targeted therapeutics should be developed carefully to focus on ROS balance in each tumor type, cancer stem cells, and their stem cell niches.

Author Contributions

Y.-C.L. wrote, reviewed the original manuscript, and generated the figures. C.-C.K. conceived and designed the work and wrote the original draft, and K.W. reviewed the references in Table 1 and drew the figures as well as the revising the draft. D.-C.W. performed the investigation and reviewed the manuscript. K.K.Y. wrote the original and revised manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the Ministry of Science and Technology (MOST111-2314-B-037-009), the National Health Research Institutes (NHRI-EX109-10720SI), Kaohsiung Medical University Hospital (KMUH111-1R77; KMUH-DK (A)112002), the National Sun-Yat-sen University-Kaohsiung Medical University Joint Research Project (NYCUKMU-113-I006), and Kaohsiung Medical University (KMU-TC113A02).

Acknowledgments

We thank Chang-Shen Lin, Ming-Ho Tsai, Richard Eckner, and Ami Aronheim for revising the manuscript.

Conflicts of Interest

The author declare no conflict of interest.

Abbreviations

AhRaryl hydrocarbon receptor
AP-1activation protein-1
AREantioxidant response element
BAXBcl-2-associated X protein
BCL2B-cell lymphoma 2
BCRPbreast cancer-resistant protein
BRCA1breast cancer susceptibility gene 1
CBPCREB binding protein
CDKN1Acyclin-dependent kinase inhibitor 1
CEcinnamaldehyde
COX2cyclooxygenase 2
CRCcolorectal cancer
CREBcyclic AMP response element binding protein
CSCscancer stem cells
DPx2glutathione peroxidase 2
DSSdextran sodium sulfate
EMTepithelial-to-mesenchymal transition
ERBB2Erb-B2 receptor tyrosine kinase 2
GCgastric cancer
GLUTsglucose transporters
G6PDglucose-6-phosphate dehydrogenase
GPR78glucose-regulated protein 78
GSHglutathione-SH
GSK3glycogen synthetase kinase 3
GSSGglutathione sulfide
H2O2hydrogen peroxide
HIF-1βhypoxia inducible factor beta
HNSCChead and neck squamous cell carcinoma
HO-1heme oxygenase 1
3-IAldindole-3-aldehyde
ICIimmune checkpoint inhibitor
IKK-α:-βinhibitor of nuclear factor κB kinase subunit-α,-β
IPAingenuity pathway analysis
iNOSnitric oxide synthase
JDP2Jun dimerization protein 2
JNKsc-Jun N-terminal kinases
Keap1Kelch-like ECH associated protein 1
LPSlipopolysaccharide
MARCOmacrophage receptor with collagenous structure
MDR1multidrug resistance protein 1
MRP1-5multidrug resistance-associated protein 1-5
MSCsmesenchymal stem cells
NACN-acetylcysteine
NFkBnuclear factor kappa B
NOnitrogen oxide
NQO1NAD(P)H: quinone dehydrogenase 1
NRF2nuclear factor erythroid 2-related factor 2
NSCLCnon-small cell lung cancer
OXPHOSoxidative phosphorylation
PEAperillaldehyde
PGDphosphoglucomutase dehydrogenase
PI3Kphosphoinositide 3-kinase
p-PERKphosphorylated protein kinase RNA-like ER kinase
PSMA2proteasome 20S alpha 2
PSMC426S proteasome regulatory subunit 4
ROSreactive oxygen species
SFNsulforaphane
SODsuperoxide dismutase
SQSRM1sequestosome-1
TALDO1trans-aldolase
TBK1TANK-binding kinase 2
TKTtransketolase
TNFtumor necrosis factor
β-TrCPβ-transduction repeat-containing E3 ubiquitin-protein ligase
Trxthioredoxin
TrxR1thioredoxin reductase 1
VEGFvascular endothelial growth factor
x-CTSLC7a11 cystine/glutamate transporter

References

  1. Gorrini, C.; Harris, I.S.; Mak, T.W. Modulation of oxidative stress as an anticancer strategy. Nat. Rev. Drug Discov. 2013, 12, 931–947. [Google Scholar] [CrossRef] [PubMed]
  2. Xing, F.; Hu, Q.; Qin, Y.; Xu, J.; Zhang, B.; Yu, X.; Wang, W. The Relationship of Redox with Hallmarks of Cancer: The Importance of Homeostasis and Context. Front. Oncol. 2022, 12, 862743. [Google Scholar] [CrossRef] [PubMed]
  3. Perillo, B.; Di Donato, M.; Pezone, A.; Di Zazzo, E.; Giovannelli, P.; Galasso, G.; Castoria, G.; Migliaccio, A. ROS in cancer therapy: The bright side of the moon. Exp. Mol. Med. 2020, 52, 192–203. [Google Scholar] [CrossRef] [PubMed]
  4. Kirkpatrick, D.L.; Powis, G. Clinically Evaluated Cancer Drugs Inhibiting Redox Signaling. Antioxid. Redox Signal. 2017, 26, 262–273. [Google Scholar] [CrossRef]
  5. Wu, X.; Zhou, Z.; Li, K.; Liu, S. Nanomaterials-Induced Redox Imbalance: Challenged and Opportunities for Nanomaterials in Cancer Therapy. Adv. Sci. 2024, 11, e2308632. [Google Scholar] [CrossRef]
  6. Wu, D.C.; Ku, C.C.; Pan, J.B.; Wuputra, K.; Yang, Y.H.; Liu, C.J.; Liu, Y.C.; Kato, K.; Saito, S.; Lin, Y.C.; et al. Heterogeneity of Phase II Enzyme Ligands on Controlling the Progression of Human Gastric Cancer Organoids as Stem Cell Therapy Model. Int. J. Mol. Sci. 2023, 24, 15911. [Google Scholar] [CrossRef] [PubMed]
  7. Poprac, P.; Jomova, K.; Simunkova, M.; Kollar, V.; Rhodes, C.J.; Valko, M. Targeting Free Radicals in Oxidative Stress-Related Human Diseases. Trends Pharmacol. Sci. 2017, 38, 592–607. [Google Scholar] [CrossRef]
  8. Liu, S.; Yue, M.; Lu, Y.; Wang, Y.; Luo, S.; Liu, X.; Jiang, J. Advancing the frontiers of colorectal cancer treatment: Harnessing ferroptosis regulation. Apoptosis 2024, 29, 86–102. [Google Scholar] [CrossRef]
  9. Schmidlin, C.J.; Shakya, A.; Dodson, M.; Chapman, E.; Zhang, D.D. The intricacies of NRF2 regulation in cancer. Semin. Cancer Biol. 2021, 76, 110–119. [Google Scholar] [CrossRef]
  10. Young, M.R.I.; Xiong, Y. Influence of vitamin D on cancer risk and treatment: Why the variability? Trends Cancer Res. 2018, 13, 43–53. [Google Scholar]
  11. Bakalova, R.; Zhelev, Z.; Miller, T.; Aoki, I.; Higashi, T. New potential biomarker for stratification of patients for pharmacological vitamin C in adjuvant settings of cancer therapy. Redox Biol. 2020, 28, 101357. [Google Scholar] [CrossRef] [PubMed]
  12. Augsburger, F.; Filippova, A.; Rasti, D.; Seredenina, T.; Lam, M.; Maghzal, G.; Mahiout, Z.; Jansen-Durr, P.; Knaus, U.G.; Doroshow, J.; et al. Pharmacological characterization of the seven human NOX isoforms and their inhibitors. Redox Biol. 2019, 26, 101272. [Google Scholar] [CrossRef] [PubMed]
  13. Liang, S.; Ma, H.Y.; Zhong, Z.; Dhar, D.; Liu, X.; Xu, J.; Koyama, Y.; Nishio, T.; Karin, D.; Karin, G.; et al. NADPH Oxidase 1 in Liver Macrophages Promotes Inflammation and Tumor Development in Mice. Gastroenterology 2019, 156, 1156–1172.e6. [Google Scholar] [CrossRef] [PubMed]
  14. Batinic-Haberle, I.; Tovmasyan, A.; Spasojevic, I. Mn Porphyrin-Based Redox-Active Drugs: Differential Effects as Cancer Therapeutics and Protectors of Normal Tissue Against Oxidative Injury. Antioxid. Redox Signal. 2018, 29, 1691–1724. [Google Scholar] [CrossRef]
  15. Wu, Z.Y.; Kim, H.J.; Lee, J.W.; Chung, I.Y.; Kim, J.S.; Lee, S.B.; Son, B.H.; Eom, J.S.; Kim, S.B.; Gong, G.Y.; et al. Breast Cancer Recurrence in the Nipple-Areola Complex After Nipple-Sparing Mastectomy with Immediate Breast Reconstruction for Invasive Breast Cancer. JAMA Surg. 2019, 154, 1030–1037. [Google Scholar] [CrossRef]
  16. Wu, J.H.; Batist, G. Glutathione and glutathione analogues; therapeutic potentials. Biochim. Biophys. Acta 2013, 1830, 3350–3353. [Google Scholar] [CrossRef] [PubMed]
  17. Iqbal, M.J.; Kabeer, A.; Abbas, Z.; Siddiqui, H.A.; Calina, D.; Sharifi-Rad, J.; Cho, W.C. Interplay of oxidative stress, cellular communication and signaling pathways in cancer. Cell Commun Signal. 2024, 22, 7. [Google Scholar] [CrossRef]
  18. Chang, I.W.; Lin, V.C.; Hung, C.H.; Wang, H.P.; Lin, Y.Y.; Wu, W.J.; Huang, C.N.; Li, C.C.; Li, W.M.; Wu, J.Y.; et al. GPX2 underexpression indicates poor prognosis in patients with urothelial carcinomas of the upper urinary tract and urinary bladder. World J Urol 2015, 33, 1777–1789. [Google Scholar] [CrossRef]
  19. Lei, Z.; Tian, D.; Zhang, C.; Zhao, S.; Su, M. Clinicopathological and prognostic significance of GPX2 protein expression in esophageal squamous cell carcinoma. BMC Cancer 2016, 16, 410. [Google Scholar] [CrossRef]
  20. Moreno Leon, L.; Gautier, M.; Allan, R.; Ilie, M.; Nottet, N.; Pons, N.; Paquet, A.; Lebrigand, K.; Truchi, M.; Fassy, J.; et al. The nuclear hypoxia-regulated NLUCAT1 long non-coding RNA contributes to an aggressive phenotype in lung adenocarcinoma through regulation of oxidative stress. Oncogene 2019, 38, 7146–7165. [Google Scholar] [CrossRef]
  21. Minato, A.; Noguchi, H.; Ohnishi, R.; Tomisaki, I.; Nakayama, T.; Fujimoto, N. Reduced Expression Level of GPX2 in T1 Bladder Cancer and its Role in Early-phase Invasion of Bladder Cancer. In Vivo 2021, 35, 753–759. [Google Scholar] [CrossRef] [PubMed]
  22. Ren, Z.; Liang, H.; Galbo, P.M., Jr.; Dharmaratne, M.; Kulkarni, A.S.; Fard, A.T.; Aoun, M.L.; Martinez-Lopez, N.; Suyama, K.; Benard, O.; et al. Redox signaling by glutathione peroxidase 2 links vascular modulation to metabolic plasticity of breast cancer. Proc. Natl. Acad. Sci. USA 2022, 119, e2107266119. [Google Scholar] [CrossRef] [PubMed]
  23. Li, F.; Dai, L.; Niu, J. GPX2 silencing relieves epithelial-mesenchymal transition, invasion, and metastasis in pancreatic cancer by downregulating Wnt pathway. J. Cell Physiol. 2020, 235, 7780–7790. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, T.; Kan, X.F.; Ma, C.; Chen, L.L.; Cheng, T.T.; Zou, Z.W.; Li, Y.; Cao, F.J.; Zhang, W.J.; Yao, J.; et al. GPX2 overexpression indicates poor prognosis in patients with hepatocellular carcinoma. Tumour Biol. 2017, 39, 1–10. [Google Scholar] [CrossRef] [PubMed]
  25. Saha, A.; Harowicz, M.R.; Cain, E.H.; Hall, A.H.; Hwang, E.S.; Marks, J.R.; Marcom, P.K.; Mazurowski, M.A. Intra-tumor molecular heterogeneity in breast cancer: Definitions of measures and association with distant recurrence-free survival. Breast Cancer Res. Treat. 2018, 172, 123–132. [Google Scholar] [CrossRef]
  26. Du, H.; Chen, B.; Jiao, N.L.; Liu, Y.H.; Sun, S.Y.; Zhang, Y.W. Elevated Glutathione Peroxidase 2 Expression Promotes Cisplatin Resistance in Lung Adenocarcinoma. Oxid. Med. Cell Longev. 2020, 2020, 7370157. [Google Scholar] [CrossRef]
  27. Hashinokuchi, A.; Matsubara, T.; Ono, Y.; Shunichi, S.; Matsudo, K.; Nagano, T.; Kinoshita, F.; Akamine, T.; Kohno, M.; Takenaka, T.; et al. Clinical and Prognostic Significance of Glutathione Peroxidase 2 in Lung Adenocarcinoma. Ann. Surg. Oncol. 2024, 31, 4822–4829. [Google Scholar] [CrossRef]
  28. Derakhshan Nazari, M.H.; Askari Dastjerdi, R.; Ghaedi Talkhouncheh, P.; Bereimipour, A.; Mollasalehi, H.; Mahshad, A.A.; Razi, A.; Nazari, M.H.; Ebrahimi Sadrabadi, A.; Taleahmad, S. GPX2 and BMP4 as Significant Molecular Alterations in The Lung Adenocarcinoma Progression: Integrated Bioinformatics Analysis. Cell J. 2022, 24, 302–308. [Google Scholar]
  29. Yang, M.; Zhu, X.; Shen, Y.; He, Q.; Qin, Y.; Shao, Y.; Yuan, L.; Ye, H. GPX2 predicts recurrence-free survival and triggers the Wnt/beta-catenin/EMT pathway in prostate cancer. PeerJ. 2022, 10, e14263. [Google Scholar] [CrossRef]
  30. Xu, H.; Hu, C.; Wang, Y.; Shi, Y.; Yuan, L.; Xu, J.; Zhang, Y.; Chen, J.; Wei, Q.; Qin, J.; et al. Glutathione peroxidase 2 knockdown suppresses gastric cancer progression and metastasis via regulation of kynurenine metabolism. Oncogene 2023, 42, 1994–2006. [Google Scholar] [CrossRef]
  31. Cui, Z.; Xu, L.; Wu, H.; Wang, M.; Lu, L.; Wu, S. Glutathione peroxidase 2: A key factor in the development of microsatellite instability in colon cancer. Pathol. Res. Pract. 2023, 243, 154372. [Google Scholar] [CrossRef] [PubMed]
  32. Peng, F.; Xu, Q.; Jing, X.; Chi, X.; Zhang, Z.; Meng, X.; Liu, X.; Yan, J.; Liu, X.; Shao, S. GPX2 promotes EMT and metastasis in non-small cell lung cancer by activating PI3K/AKT/mTOR/Snail signaling axis. FASEB Bioadv. 2023, 5, 233–250. [Google Scholar] [CrossRef] [PubMed]
  33. Cao, P.; Gu, J.; Liu, M.; Wang, Y.; Chen, M.; Jiang, Y.; Wang, X.; Zhu, S.; Gao, X.; Li, S. BRMS1L confers anticancer activity in non-small cell lung cancer by transcriptionally inducing a redox imbalance in the GPX2-ROS pathway. Transl. Oncol. 2024, 41, 101870. [Google Scholar] [CrossRef] [PubMed]
  34. McGovern, K.; Castro, A.C.; Cavanaugh, J.; Coma, S.; Walsh, M.; Tchaicha, J.; Syed, S.; Natarajan, P.; Manfredi, M.; Zhang, X.M.; et al. Discovery and Characterization of a Novel Aryl Hydrocarbon Receptor Inhibitor, IK-175, and Its Inhibitory Activity on Tumor Immune Suppression. Mol. Cancer Ther. 2022, 21, 1261–1272. [Google Scholar] [CrossRef] [PubMed]
  35. Kumari, S.; Badana, A.K.; Murali Moran, G.; Shailender, G.; Malla, R. Reactive oxygen species: A key constituent in cancer survival. BioMark Insights 2018, 13, 1–9. [Google Scholar] [CrossRef]
  36. Raza, M.H.; Siraj, S.; Arshad, A.; Waheed, U.; Aldakheel, F.; Alduraywish, S.; Arshad, M.J. ROS-modulated therapeutic approaches in cancer treatment. Cancer Res. Clin. Oncol. 2017, 143, 1789–1809. [Google Scholar] [CrossRef]
  37. Assi, M. The differential role of reactive oxygen species in early and late stages of cancer. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2017, 313, R646–R653. [Google Scholar] [CrossRef]
  38. Esworthy, R.S.; Doroshow, J.H.; Chu, F.F. The beginning of GPX2 and 30 years later. Free Radic. Biol. Med. 2022, 188, 419–433. [Google Scholar] [CrossRef]
  39. Selistre-de-Araujo, H.S.; Pachane, B.C.; Altei, W.F. Tumor heterogeneity and the dilemma of antioxidant therapies in cancer. Ann. Transl. Med. 2022, 10, 1074. [Google Scholar] [CrossRef]
  40. Jiang, H.; Zuo, J.; Li, B.; Chen, R.; Luo, K.; Xiang, X.; Lu, S.; Huang, C.; Liu, L.; Tang, J.; et al. Drug-induced oxidative stress in cancer treatments: Angel or devil? Redox. Biol. 2023, 63, 102754. [Google Scholar] [CrossRef]
  41. Gorrini, C.; Baniasadi, P.S.; Harris, I.S.; Silvester, J.; Inoue, S.; Snow, B.; Joshi, P.A.; Wakeham, A.; Molyneux, S.D.; Martin, B.; et al. BRCA1 interacts with Nrf2 to regulate antioxidant signaling and cell survival. J. Exp. Med. 2013, 210, 1529–1544. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, W.; Sun, Z.; Wang, X.J.; Jiang, T.; Huang, Z.; Fang, D.; Zhang, D.D. Direct interaction between Nrf2 and p21 (Cip1/WAF1) upregulates the Nrf2-mediated antioxidant response. Mol. Cell 2009, 34, 663–673. [Google Scholar] [CrossRef] [PubMed]
  43. Niture, S.K.; Khatri, R.; Jaiswal, A.K. Regulation of Nrf2-an update. Free Radic. Biol. Med. 2014, 66, 36–44. [Google Scholar] [CrossRef]
  44. Long, D.J., 2nd; Waikel, R.L.; Wang, X.J.; Perlaky, L.; Roop, D.R.; Jaiswal, A.K. NAD(P)H:quinone oxidoreductase 1 deficiency increases susceptibility to benzo(a)pyrene-induced mouse skin carcinogenesis. Cancer Res. 2000, 60, 5913–5915. [Google Scholar]
  45. Ramos-Gomez, M.; Kwak, M.K.; Dolan, P.M.; Itoh, K.; Yamamoto, M.; Talalay, P.; Kensler, T.W. Sensitivity to carcinogenesis is increased and chemoprotective efficacy of enzyme inducers is lost in nrf2 transcription factor-deficient mice. Proc. Natl. Acad. Sci. USA 2001, 98, 3410–3415. [Google Scholar] [CrossRef]
  46. Kitamura, Y.; Umemura, T.; Kanki, K.; Kodama, Y.; Kitamoto, S.; Saito, K.; Itoh, K.; Yamamoto, M.; Masegi, T.; Nishikawa, A.; et al. Increased susceptibility to hepatocarcinogenicity of Nrf2-deficient mice exposed to 2-amino-3-methylimidazo[4,5-f]quinoline. Cancer Sci. 2007, 98, 19–24. [Google Scholar] [CrossRef] [PubMed]
  47. Iida, K.; Itoh, K.; Maher, J.M.; Kumagai, Y.; Oyasu, R.; Mori, Y.; Shimazui, T.; Akaza, H.; Yamamoto, M. Nrf2 and p53 cooperatively protect against BBN-induced urinary bladder carcinogenesis. Carcinogenesis 2007, 28, 2398–2403. [Google Scholar] [CrossRef]
  48. Wakabayashi, N.; Dinkova-Kostova, A.T.; Holtzclaw, W.D.; Kang, M.I.; Kobayashi, A.; Yamamoto, M.; Kensler, T.W.; Talalay, P. Protection against electrophile and oxidant stress by induction of the phase 2 response: Fate of cysteines of the Keap1 sensor modified by inducers. Proc. Natl. Acad. Sci. USA 2004, 101, 2040–2045. [Google Scholar] [CrossRef]
  49. Gamet-Payrastre, L.; Li, P.; Lumeau, S.; Cassar, G.; Dupont, M.A.; Chevolleau, S.; Gasc, N.; Tulliez, J.; Terce, F. Sulforaphane, a naturally occurring isothiocyanate, induces cell cycle arrest and apoptosis in HT29 human colon cancer cells. Cancer Res. 2000, 60, 1426–1433. [Google Scholar]
  50. Heiss, E.; Herhaus, C.; Klimo, K.; Bartsch, H.; Gerhauser, C. Nuclear factor kappa B is a molecular target for sulforaphane-mediated anti-inflammatory mechanisms. J. Biol. Chem. 2001, 276, 32008–32015. [Google Scholar] [CrossRef]
  51. Dickinson, S.E.; Melton, T.F.; Olson, E.R.; Zhang, J.; Saboda, K.; Bowden, G.T. Inhibition of activator protein-1 by sulforaphane involves interaction with cysteine in the cFos DNA-binding domain: Implications for chemoprevention of UVB-induced skin cancer. Cancer Res. 2009, 69, 7103–7110. [Google Scholar] [CrossRef] [PubMed]
  52. Liby, K.T.; Royce, D.B.; Risingsong, R.; Williams, C.R.; Maitra, A.; Hruban, R.H.; Sporn, M.B. Synthetic triterpenoids prolong survival in a transgenic mouse model of pancreatic cancer. Cancer Prev. Res. 2010, 3, 1427–1434. [Google Scholar] [CrossRef] [PubMed]
  53. Kim, E.H.; Deng, C.; Sporn, M.B.; Royce, D.B.; Risingsong, R.; Williams, C.R.; Liby, K.T. CDDO-methyl ester delays breast cancer development in BRCA1-mutated mice. Cancer Prev. Res. 2012, 5, 89–97. [Google Scholar] [CrossRef] [PubMed]
  54. Boyanapalli, S.S.; Paredes-Gonzalez, X.; Fuentes, F.; Zhang, C.; Guo, Y.; Pung, D.; Saw, C.L.; Kong, A.N. Nrf2 knockout attenuates the anti-inflammatory effects of phenethyl isothiocyanate and curcumin. Chem. Res. Toxicol. 2014, 27, 2036–2043. [Google Scholar] [CrossRef] [PubMed]
  55. Kobayashi, E.H.; Suzuki, T.; Funayama, R.; Nagashima, T.; Hayashi, M.; Sekine, H.; Tanaka, N.; Moriguchi, T.; Motohashi, H.; Nakayama, K.; et al. Nrf2 suppresses macrophage inflammatory response by blocking proinflammatory cytokine transcription. Nat. Commun. 2016, 7, 11624. [Google Scholar] [CrossRef]
  56. Kong, X.; Thimmulappa, R.; Craciun, F.; Harvey, C.; Singh, A.; Kombairaju, P.; Reddy, S.P.; Remick, D.; Biswal, S. Enhancing Nrf2 pathway by disruption of Keap1 in myeloid leukocytes protects against sepsis. Am. J. Respir. Crit. Care Med. 2011, 184, 928–938. [Google Scholar] [CrossRef]
  57. Chen, X.L.; Dodd, G.; Thomas, S.; Zhang, X.; Wasserman, M.A.; Rovin, B.H.; Kunsch, C. Activation of Nrf2/ARE pathway protects endothelial cells from oxidant injury and inhibits inflammatory gene expression. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, H1862–H1870. [Google Scholar] [CrossRef]
  58. Kim, J.E.; You, D.J.; Lee, C.; Ahn, C.; Seong, J.Y.; Hwang, J.I. Suppression of NF-kappaB signaling by KEAP1 regulation of IKKbeta activity through autophagic degradation and inhibition of phosphorylation. Cell Signal. 2010, 22, 1645–1654. [Google Scholar] [CrossRef]
  59. Gao, W.; Guo, L.; Yang, Y.; Wang, Y.; Xia, S.; Gong, H.; Zhang, B.K.; Yan, M. Dissecting the Crosstalk Between Nrf2 and NF-kappaB Response Pathways in Drug-Induced Toxicity. Front. Cell Dev. Biol. 2021, 9, 809952. [Google Scholar]
  60. Kim, Y.C.; Masutani, H.; Yamaguchi, Y.; Itoh, K.; Yamamoto, M.; Yodoi, J. Hemin-induced activation of the thioredoxin gene by Nrf2. A differential regulation of the antioxidant responsive element by a switch of its binding factors. J. Biol. Chem. 2001, 276, 18399–18406. [Google Scholar] [CrossRef]
  61. Hanada, N.; Takahata, T.; Zhou, Q.; Ye, X.; Sun, R.; Itoh, J.; Ishiguro, A.; Kijima, H.; Mimura, J.; Itoh, K.; et al. Methylation of the KEAP1 gene promoter region in human colorectal cancer. BMC Cancer 2012, 12, 66. [Google Scholar] [CrossRef] [PubMed]
  62. DeNicola, G.M.; Karreth, F.A.; Humpton, T.J.; Gopinathan, A.; Wei, C.; Frese, K.; Mangal, D.; Yu, K.H.; Yeo, C.J.; Calhoun, E.S.; et al. Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 2011, 475, 106–109. [Google Scholar] [CrossRef] [PubMed]
  63. Rojo, A.I.; Rada, P.; Mendiola, M.; Ortega-Molina, A.; Wojdyla, K.; Rogowska-Wrzesinska, A.; Hardisson, D.; Serrano, M.; Cuadrado, A. The PTEN/NRF2 axis promotes human carcinogenesis. Antioxid. Redox Signal. 2014, 21, 2498–2514. [Google Scholar] [CrossRef]
  64. Mitsuishi, Y.; Taguchi, K.; Kawatani, Y.; Shibata, T.; Nukiwa, T.; Aburatani, H.; Yamamoto, M.; Motohashi, H. Nrf2 redirects glucose and glutamine into anabolic pathways in metabolic reprogramming. Cancer Cell 2012, 22, 66–79. [Google Scholar] [CrossRef] [PubMed]
  65. Kitteringham, N.R.; Abdullah, A.; Walsh, J.; Randle, L.; Jenkins, R.E.; Sison, R.; Goldring, C.E.; Powell, H.; Sanderson, C.; Williams, S.; et al. Proteomic analysis of Nrf2 deficient transgenic mice reveals cellular defence and lipid metabolism as primary Nrf2-dependent pathways in the liver. J. Proteomics 2010, 73, 1612–1631. [Google Scholar] [CrossRef]
  66. Malhotra, D.; Portales-Casamar, E.; Singh, A.; Srivastava, S.; Arenillas, D.; Happel, C.; Shyr, C.; Wakabayashi, N.; Kensler, T.W.; Wasserman, W.W.; 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, 5718–5734. [Google Scholar] [CrossRef]
  67. Reddy, N.M.; Kleeberger, S.R.; Bream, J.H.; Fallon, P.G.; Kensler, T.W.; Yamamoto, M.; Reddy, S.P. Genetic disruption of the Nrf2 compromises cell-cycle progression by impairing GSH-induced redox signaling. Oncogene 2008, 27, 5821–5832. [Google Scholar] [CrossRef]
  68. Chang, C.W.; Chen, Y.S.; Tsay, Y.G.; Han, C.L.; Chen, Y.J.; Yang, C.C.; Hung, K.F.; Lin, C.H.; Huang, T.Y.; Kao, S.Y.; et al. ROS-independent ER stress-mediated NRF2 activation promotes warburg effect to maintain stemness-associated properties of cancer-initiating cells. Cell Death Dis. 2018, 9, 194. [Google Scholar] [CrossRef]
  69. Chang, C.W.; Chen, Y.S.; Chou, S.H.; Han, C.L.; Chen, Y.J.; Yang, C.C.; Huang, C.Y.; Lo, J.F. Distinct subpopulations of head and neck cancer cells with different levels of intracellular reactive oxygen species exhibit diverse stemness, proliferation, and chemosensitivity. Cancer Res. 2014, 74, 6291–6305. [Google Scholar] [CrossRef]
  70. Kumar, H.; Kumar, R.M.; Bhattacharjee, D.; Somanna, P.; Jain, V. Role of Nrf2 Signaling Cascade in Breast Cancer: Strategies and Treatment. Front. Pharmacol. 2022, 13, 720076. [Google Scholar] [CrossRef]
  71. Ryoo, I.G.; Lee, S.H.; Kwak, M.K. Redox Modulating NRF2: A Potential Mediator of Cancer Stem Cell Resistance. Oxid. Med. Cell Longev. 2016, 2016, 2428153. [Google Scholar] [CrossRef] [PubMed]
  72. Singh, A.; Boldin-Adamsky, S.; Thimmulappa, R.K.; Rath, S.K.; Ashush, H.; Coulter, J.; Blackford, A.; Goodman, S.N.; Bunz, F.; Watson, W.H.; et al. RNAi-mediated silencing of nuclear factor erythroid-2-related factor 2 gene expression in non-small cell lung cancer inhibits tumor growth and increases efficacy of chemotherapy. Cancer Res. 2008, 68, 7975–7984. [Google Scholar] [CrossRef]
  73. Yasuda, T.; Ishimoto, T.; Baba, H. Conflicting metabolic alterations in cancer stem cells and regulation by the stromal niche. Regen. Ther. 2021, 17, 8–12. [Google Scholar] [CrossRef]
  74. Gao, L.; Morine, Y.; Yamada, S.; Saito, Y.; Ikemoto, T.; Tokuda, K.; Takasu, C.; Miyazaki, K.; Shimada, M. Nrf2 signaling promotes cancer stemness, migration, and expression of ABC transporter genes in sorafenib-resistant hepatocellular carcinoma cells. PLoS ONE 2021, 16, e0256755. [Google Scholar] [CrossRef] [PubMed]
  75. Kahroba, H.; Shirmohamadi, M.; Hejazi, M.S.; Samadi, N. The Role of Nrf2 signaling in cancer stem cells: From stemness and self-renewal to tumorigenesis and chemoresistance. Life Sci. 2019, 239, 116986. [Google Scholar] [CrossRef] [PubMed]
  76. Ridge, S.M.; Sullivan, F.J.; Glynn, S.A. Mesenchymal stem cells: Key players in cancer progression. Mol. Cancer 2017, 16, 31. [Google Scholar] [CrossRef]
  77. Mohammadzadeh-Vardin, M.; Habibi Roudkenar, M.; Jahanian-Najafabadi, A. Adenovirus-Mediated Over-Expression of Nrf2 Within Mesenchymal Stem Cells (MSCs) Protected Rats Against Acute Kidney Injury. Adv. Pharm. Bull. 2015, 5, 201–208. [Google Scholar] [CrossRef]
  78. Yuan, Z.; Zhang, J.; Huang, Y.; Zhang, Y.; Liu, W.; Wang, G.; Zhang, Q.; Wang, G.; Yang, Y.; Li, H.; et al. NRF2 overexpression in mesenchymal stem cells induces stem-cell marker expression and enhances osteoblastic differentiation. Biochem. Biophys. Res. Commun. 2017, 491, 228–235. [Google Scholar] [CrossRef]
  79. Kitamura, H.; Motohashi, H. NRF2 addiction in cancer cells. Cancer Sci. 2018, 109, 900–911. [Google Scholar] [CrossRef]
  80. Okazaki, K.; Papagiannakopoulos, T.; Motohashi, H. Metabolic features of cancer cells in NRF2 addiction status. Biophys. Rev. 2020, 12, 435–441. [Google Scholar] [CrossRef]
  81. 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, 149–160. [Google Scholar] [CrossRef] [PubMed]
  82. Bussolati, B.; Mason, J.C. Dual role of VEGF-induced heme-oxygenase-1 in angiogenesis. Antioxid. Redox Signal. 2006, 8, 1153–1163. [Google Scholar] [CrossRef] [PubMed]
  83. Rushworth, S.A.; MacEwan, D.J. HO-1 underlies resistance of AML cells to TNF-induced apoptosis. Blood 2008, 111, 3793–3801. [Google Scholar] [CrossRef] [PubMed]
  84. Niture, S.K.; Jaiswal, A.K. Nrf2 protein up-regulates antiapoptotic protein Bcl-2 and prevents cellular apoptosis. J. Biol. Chem. 2012, 287, 9873–9886. [Google Scholar] [CrossRef]
  85. Elsby, R.; Kitteringham, N.R.; Goldring, C.E.; Lovatt, C.A.; Chamberlain, M.; Henderson, C.J.; Wolf, C.R.; Park, B.K. Increased constitutive c-Jun N-terminal kinase signaling in mice lacking glutathione S-transferase Pi. J. Biol. Chem. 2003, 278, 22243–22249. [Google Scholar] [CrossRef]
  86. Jain, A.; Lamark, T.; Sjottem, E.; Larsen, K.B.; Awuh, J.A.; Overvatn, A.; McMahon, M.; Hayes, J.D.; Johansen, T. p62/SQSTM1 is a target gene for transcription factor NRF2 and creates a positive feedback loop by inducing antioxidant response element-driven gene transcription. J. Biol. Chem. 2010, 285, 22576–22591. [Google Scholar] [CrossRef]
  87. Komatsu, M.; Kurokawa, H.; Waguri, S.; Taguchi, K.; Kobayashi, A.; Ichimura, Y.; Sou, Y.S.; Ueno, I.; Sakamoto, A.; Tong, K.I.; et al. The selective autophagy substrate p62 activates the stress responsive transcription factor Nrf2 through inactivation of Keap1. Nat. Cell Biol. 2010, 12, 213–223. [Google Scholar] [CrossRef]
  88. Towers, C.G.; Fitzwalter, B.E.; Regan, D.; Goodspeed, A.; Morgan, M.J.; Liu, C.W.; Gustafson, D.L.; Thorburn, A. Cancer Cells Upregulate NRF2 Signaling to Adapt to Autophagy Inhibition. Dev. Cell 2019, 50, 690–703.e6. [Google Scholar] [CrossRef]
  89. Taguchi, K.; Hirano, I.; Itoh, T.; Tanaka, M.; Miyajima, A.; Suzuki, A.; Motohashi, H.; Yamamoto, M. Nrf2 enhances cholangiocyte expansion in Pten-deficient livers. Mol. Cell Biol. 2014, 34, 900–913. [Google Scholar] [CrossRef]
  90. Shirasaki, K.; Taguchi, K.; Unno, M.; Motohashi, H.; Yamamoto, M. NF-E2-related factor 2 promotes compensatory liver hypertrophy after portal vein branch ligation in mice. Hepatology 2014, 59, 2371–2382. [Google Scholar] [CrossRef]
  91. Suzuki, T.; Seki, S.; Hiramoto, K.; Naganuma, E.; Kobayashi, E.H.; Yamaoka, A.; Baird, L.; Takahashi, N.; Sato, H.; Yamamoto, M. Hyperactivation of Nrf2 in early tubular development induces nephrogenic diabetes insipidus. Nat. Commun. 2017, 8, 14577. [Google Scholar] [CrossRef] [PubMed]
  92. Murakami, S.; Suzuki, T.; Harigae, H.; Romeo, P.H.; Yamamoto, M.; Motohashi, H. NRF2 Activation Impairs Quiescence and Bone Marrow Reconstitution Capacity of Hematopoietic Stem Cells. Mol. Cell Biol. 2017, 37, e00086-17. [Google Scholar] [CrossRef]
  93. Taguchi, K.; Maher, J.M.; Suzuki, T.; Kawatani, Y.; Motohashi, H.; Yamamoto, M. Genetic analysis of cytoprotective functions supported by graded expression of Keap1. Mol. Cell Biol. 2010, 30, 3016–3026. [Google Scholar] [CrossRef]
  94. Jeong, Y.; Hoang, N.T.; Lovejoy, A.; Stehr, H.; Newman, A.M.; Gentles, A.J.; Kong, W.; Truong, D.; Martin, S.; Chaudhuri, A.; et al. Role of KEAP1/NRF2 and TP53 Mutations in Lung Squamous Cell Carcinoma Development and Radiation Resistance. Cancer Discov. 2017, 7, 86–101. [Google Scholar] [CrossRef]
  95. Bai, X.; Chen, Y.; Hou, X.; Huang, M.; Jin, J. Emerging role of NRF2 in chemoresistance by regulating drug-metabolizing enzymes and efflux transporters. Drug Metab. Rev. 2016, 48, 541–567. [Google Scholar] [CrossRef] [PubMed]
  96. Ryoo, I.G.; Kim, G.; Choi, B.H.; Lee, S.H.; Kwak, M.K. Involvement of NRF2 Signaling in Doxorubicin Resistance of Cancer Stem Cell-Enriched Colonospheres. Biomol. Ther. 2016, 24, 482–488. [Google Scholar] [CrossRef]
  97. Sasaki, H.; Shitara, M.; Yokota, K.; Hikosaka, Y.; Moriyama, S.; Yano, M.; Fujii, Y. MRP3 gene expression correlates with NRF2 mutations in lung squamous cell carcinomas. Mol. Med. Rep. 2012, 6, 705–708. [Google Scholar] [CrossRef] [PubMed]
  98. Gao, A.M.; Ke, Z.P.; Wang, J.N.; Yang, J.Y.; Chen, S.Y.; Chen, H. Apigenin sensitizes doxorubicin-resistant hepatocellular carcinoma BEL-7402/ADM cells to doxorubicin via inhibiting PI3K/Akt/Nrf2 pathway. Carcinogenesis 2013, 34, 1806–1814. [Google Scholar] [CrossRef]
  99. Ahmed, H.M. Ethnomedicinal, Phytochemical and Pharmacological Investigations of Perilla frutescens (L.) Britt. Molecules 2018, 24, 102. [Google Scholar] [CrossRef]
  100. Ji, W.W.; Wang, S.Y.; Ma, Z.Q.; Li, R.P.; Li, S.S.; Xue, J.S.; Li, W.; Niu, X.X.; Yan, L.; Zhang, X.; et al. Effects of perillaldehyde on alternations in serum cytokines and depressive-like behavior in mice after lipopolysaccharide administration. Pharmacol. Biochem. Behav. 2014, 116, 1–8. [Google Scholar] [CrossRef]
  101. Song, Y.; Sun, R.; Ji, Z.; Li, X.; Fu, Q.; Ma, S. Perilla aldehyde attenuates CUMS-induced depressive-like behaviors via regulating TXNIP/TRX/NLRP3 pathway in rats. Life Sci. 2018, 206, 117–124. [Google Scholar] [CrossRef] [PubMed]
  102. Uemura, T.; Yashiro, T.; Oda, R.; Shioya, N.; Nakajima, T.; Hachisu, M.; Kobayashi, S.; Nishiyama, C.; Arimura, G.I. Intestinal Anti-Inflammatory Activity of Perillaldehyde. J. Agric. Food Chem. 2018, 66, 3443–3448. [Google Scholar] [CrossRef] [PubMed]
  103. Fuyuno, Y.; Uchi, H.; Yasumatsu, M.; Morino-Koga, S.; Tanaka, Y.; Mitoma, C.; Furue, M. Perillaldehyde Inhibits AHR Signaling and Activates NRF2 Antioxidant Pathway in Human Keratinocytes. Oxid. Med. Cell Longev. 2018, 2018, 9524657. [Google Scholar] [CrossRef]
  104. Cabello, C.M.; Bair, W.B., 3rd; Lamore, S.D.; Ley, S.; Bause, A.S.; Azimian, S.; Wondrak, G.T. The cinnamon-derived Michael acceptor cinnamic aldehyde impairs melanoma cell proliferation, invasiveness, and tumor growth. Free Radic. Biol. Med. 2009, 46, 220–231. [Google Scholar] [CrossRef]
  105. Zhao, J.; Zhang, X.; Dong, L.; Wen, Y.; Zheng, X.; Zhang, C.; Chen, R.; Zhang, Y.; Li, Y.; He, T.; et al. Cinnamaldehyde inhibits inflammation and brain damage in a mouse model of permanent cerebral ischaemia. Br. J. Pharmacol. 2015, 172, 5009–5023. [Google Scholar] [CrossRef]
  106. Kim, T.W. Cinnamaldehyde induces autophagy-mediated cell death through ER stress and epigenetic modification in gastric cancer cells. Acta Pharmacol. Sin. 2022, 43, 712–723. [Google Scholar] [CrossRef] [PubMed]
  107. Uchi, H.; Yasumatsu, M.; Morino-Koga, S.; Mitoma, C.; Furue, M. Inhibition of aryl hydrocarbon receptor signaling and induction of NRF2-mediated antioxidant activity by cinnamaldehyde in human keratinocytes. J. Dermatol. Sci. 2017, 85, 36–43. [Google Scholar] [CrossRef]
  108. Wang, Y.; Wu, H.; Dong, N.; Su, X.; Duan, M.; Wei, Y.; Wei, J.; Liu, G.; Peng, Q.; Zhao, Y. Sulforaphane induces S-phase arrest and apoptosis via p53-dependent manner in gastric cancer cells. Sci. Rep. 2021, 11, 2504. [Google Scholar] [CrossRef] [PubMed]
  109. Zhang, Y.; Lu, Q.; Li, N.; Xu, M.; Miyamoto, T.; Liu, J. Sulforaphane suppresses metastasis of triple-negative breast cancer cells by targeting the RAF/MEK/ERK pathway. NPJ Breast Cancer 2022, 8, 40. [Google Scholar] [CrossRef]
  110. Silva-Palacios, A.; Ostolga-Chavarria, M.; Sanchez-Garibay, C.; Rojas-Morales, P.; Galvan-Arzate, S.; Buelna-Chontal, M.; Pavon, N.; Pedraza-Chaverri, J.; Konigsberg, M.; Zazueta, C. Sulforaphane protects from myocardial ischemia-reperfusion damage through the balanced activation of Nrf2/AhR. Free Radic. Biol. Med. 2019, 143, 331–340. [Google Scholar] [CrossRef]
  111. Cao, S.; Hu, S.; Jiang, P.; Zhang, Z.; Li, L.; Wu, Q. Effects of sulforaphane on breast cancer based on metabolome and microbiome. Food Sci. Nutr. 2023, 11, 2277–2287. [Google Scholar] [CrossRef] [PubMed]
  112. Yeager, R.L.; Reisman, S.A.; Aleksunes, L.M.; Klaassen, C.D. Introducing the "TCDD-inducible AhR-Nrf2 gene battery. Toxicol. Sci. 2009, 111, 238–246. [Google Scholar] [CrossRef]
  113. Wuputra, K.; Tsai, M.H.; Kato, K.; Ku, C.C.; Pan, J.B.; Yang, Y.H.; Saito, S.; Wu, C.C.; Lin, Y.C.; Cheng, K.H.; et al. Jdp2 is a spatiotemporal transcriptional activator of the AhR via the Nrf2 gene battery. Inflamm. Regen. 2023, 43, 42. [Google Scholar] [CrossRef] [PubMed]
  114. Sultana, S.; Elengickal, A.; Bensreti, H.; Belin de Chantemèle, E.; McGee-Lawrence, M.E.; Hamrick, M.W. The kynurenine pathway in HIV, frailty and inflammaging. Front. Immunol. 2023, 14, 1244622. [Google Scholar] [CrossRef] [PubMed]
  115. Chaudhry, K.A.; Bianchi-Smiraglia, A. The aryl hydrocarbon receptor as a tumor modulator: Mechanisms to therapy. Front. Oncol. 2024, 14, 1375905. [Google Scholar] [CrossRef] [PubMed]
  116. Zhuang, H.; Li, B.; Xie, T.; Xu, C.; Ren, X.; Jiang, F.; Lei, T.; Zhou, P. Indole-3-aldehyde alleviates chondrocytes inflammation through the AhR-NF-kappaB signalling pathway. Int Immunopharmacol. 2022, 113, 109314. [Google Scholar] [CrossRef]
  117. Wang, M.; Guo, J.; Hart, A.L.; Li, J.V. Indole-3-Aldehyde Reduces Inflammatory Responses and Restores Intestinal Epithelial Barrier Function Partially via Aryl Hydrocarbon Receptor (AhR) in Experimental Colitis Models. J Inflamm Res. 2023, 16, 5845–5864. [Google Scholar] [CrossRef]
  118. Renga, G.; Nunzi, E.; Pariano, M.; Puccetti, M.; Bellet, M.M.; Pieraccini, G.; D’Onofrio, F.; Santarelli, I.; Stincardini, C.; Aversa, F.; et al. Optimizing therapeutic outcomes of immune checkpoint blockade by a microbial tryptophan metabolite. J. Immunother. Cancer 2022, 10, e003725. [Google Scholar] [CrossRef]
  119. Kober, C.; Roewe, J.; Schmees, N.; Roese, L.; Roehn, U.; Bader, B.; Stoeckigt, D.; Prinz, F.; Gorjanacz, M.; Roider, H.G.; et al. Targeting the aryl hydrocarbon receptor (AhR) with BAY 2416964: A selective small molecule inhibitor for cancer immunotherapy. J. Immunother. Cancer 2023, 11, e007495. [Google Scholar] [CrossRef]
  120. Ratain, M.J.; Eisen, T.; Stadler, W.M.; Flaherty, K.T.; Kaye, S.B.; Rosner, G.L.; Gore, M.; Desai, A.A.; Patnaik, A.; Xiong, H.Q.; et al. Phase II placebo-controlled randomized discontinuation trial of sorafenib in patients with metastatic renal cell carcinoma. J Clin Oncol 2006, 24, 2505–2512. [Google Scholar] [CrossRef]
  121. Itkin, B.; Breen, A.; Turyanska, L.; Sandes, E.O.; Bradshaw, T.D.; Loaiza-Perez, A.I. New Treatments in Renal Cancer: The AhR Ligands. Int J Mol Sci 2020, 21, 3551. [Google Scholar] [CrossRef] [PubMed]
  122. Mattina, J.; Carlisle, B.; Hachem, Y.; Fergusson, D.; Kimmelman, J. Inefficiencies and Patient Burdens in the Development of the Targeted Cancer Drug Sorafenib: A Systematic Review. PLoS Biol. 2017, 15, e2000487. [Google Scholar] [CrossRef] [PubMed]
  123. Silverberg, J.I.; Eichenfield, L.F.; Hebert, A.A.; Simpson, E.L.; Stein Gold, L.; Bissonnette, R.; Papp, K.A.; Browning, J.; Kwong, P.; Korman, N.J.; et al. Tapinarof cream 1% once daily: Significant efficacy in the treatment of moderate to severe atopic dermatitis in adults and children down to 2 years of age in the pivotal phase 3 ADORING trials. J. Am. Acad. Dermatol. 2024, 91, 457–465. [Google Scholar] [CrossRef] [PubMed]
  124. Pilotto, F.; Chellapandi, D.M.; Puccio, H. Omaveloxolone: A groundbreaking milestone as the first FDA-approved drug for Friedreich ataxia. Trends Mol. Med. 2024, 30, 117–125. [Google Scholar] [CrossRef]
  125. Cassidy, P.B.; Liu, T.; Florell, S.R.; Honeggar, M.; Leachman, S.A.; Boucher, K.M.; Grossman, D. A Phase II Randomized Placebo-Controlled Trial of Oral N-acetylcysteine for Protection of Melanocytic Nevi against UV-Induced Oxidative Stress In Vivo. Cancer Prev. Res. 2017, 10, 36–44. [Google Scholar] [CrossRef]
  126. Sari, T.T.; Maritha, I.D.; Putri, I.A. N-acetylcysteine vs placebo as adjunctive treatment in paediatric leukaemia: A single blind, randomized controlled trial. Sri Lanka J. Child Heal. 2023, 52, 299–306. [Google Scholar] [CrossRef]
  127. Rajewska-Tabor, J.; Sosinska-Zawierucha, P.; Pyda, M.; Lesiak, M.; Breborowicz, A. Protective role of N-acetylcysteine and Sulodexide on endothelial cells exposed on patients’ serum after SARS-CoV-2 infection. Front Cell Infect. Microbiol. 2023, 13, 1268016. [Google Scholar] [CrossRef]
  128. Gupta, S.K.; Kamendulis, L.M.; Clauss, M.A.; Liu, Z. A randomized, placebo-controlled pilot trial of N-acetylcysteine on oxidative stress and endothelial function in HIV-infected older adults receiving antiretroviral treatment. AIDS 2016, 30, 2389–2391. [Google Scholar] [CrossRef] [PubMed]
  129. Boutin, J.A.; Kennaway, D.J.; Jockers, R. Melatonin: Facts, Extrapolations and Clinical Trials. Biomolecules 2023, 13, 943. [Google Scholar] [CrossRef]
  130. Aisa-Álvarez, A.; Pérez-Torres, I.; Guarner-Lans, V.; Manzano-Pech, L.; Cruz-Soto, R.; Márquez-Velasco, R.; Casarez-Alvarado, S.; Franco-Granillo, J.; Núñez-Martínez, M.E.; Soto, M.E. Randomized Clinical Trial of Antioxidant Therapy Patients with Septic Shock and Organ Dysfunction in the ICU: SOFA Score Reduction by Improvement of the Enzymatic and Non-Enzymatic Antioxidant System. Cells 2023, 12, 1330. [Google Scholar] [CrossRef]
  131. Ding, S.; Hu, A.; Hu, Y.; Ma, J.; Weng, P.; Dai, J. Anti-hepatoma cells function of luteolin through inducing apoptosis and cell cycle arrest. Tumor. Biol. 2013, 35, 3053–3060. [Google Scholar] [CrossRef] [PubMed]
  132. Young, S.K.; Baird, T.D.; Wek, R.C. Translation Regulation of the Glutamyl-prolyl-tRNA Synthetase Gene EPRS through Bypass of Upstream Open Reading Frames with Noncanonical Initiation Codons. J. Biol. Chem. 2016, 291, 10824–10835. [Google Scholar] [CrossRef] [PubMed]
  133. Fan, X.-X.; Leung, E.L.-H.; Xie, Y.; Liu, Z.Q.; Zheng, Y.F.; Yao, X.J.; Lu, L.L.; Wu, J.L.; He, J.-X.; Yuan, Z.-W.; et al. Suppression of Lipogenesis via Reactive Oxygen Species–AMPK Signaling for Treating Malignant and Proliferative Diseases. Antioxidants Redox Signal. 2018, 28, 339–357. [Google Scholar] [CrossRef]
  134. Chen, X.X.; Chen, Y.X.; Bi, H.X.; Zhao, X.; Zhang, L.F.; Liu, J.Y.; Shi, Y.Q. Efficacy and safety of triple therapy containing berberine hydrochloride, amoxicillin, and rabeprazole in the eradication of Helicobacter pylori. J. Dig. Dis. 2022, 23, 568–576. [Google Scholar] [CrossRef]
  135. Zhang, J.; Han, C.; Lu, W.Q.; Wang, N.; Wu, S.R.; Wang, Y.X.; Ma, J.P.; Wang, J.H.; Hao, C.; Yuan, D.H.; et al. A randomized, multicenter and noninferiority study of amoxicillin plus berberine vs tetracycline plus furazolidone in quadruple therapy for Helicobacter pylori rescue treatment. J. Dig. Dis. 2020, 21, 256–263. [Google Scholar] [CrossRef]
  136. Chen, Y.-X.; Gao, Q.-Y.; Zou, T.-H.; Wang, B.-M.; Liu, S.-D.; Sheng, J.-Q.; Ren, J.-L.; Zou, X.-P.; Liu, Z.-J.; Song, Y.-Y.; et al. Berberine versus placebo for the prevention of recurrence of colorectal adenoma: A multicentre, double-blinded, randomised controlled study. Lancet Gastroenterol. Hepatol. 2020, 5, 267–275. [Google Scholar] [CrossRef]
  137. Hall, M.J. Updates in chemoprevention research for hereditary gastrointestinal and polyposis syndromes. Curr. Treat. Options Gastro. 2021, 19, 30–46. [Google Scholar] [CrossRef]
  138. Parikh, R.A.; Taylor, J.A.; Chen, Q.; Woolbright, B.L.; Chen, P.; Wulff-Burchfield, E.M.; Holzbeierlein, J.; Jensen, R.A.; Drisko, J.A. IV vitamin C with chemotherapy for cisplatin ineligible bladder cancer patients (CI-MIBC) first-stage analysis NCT04046094. J. Clin. Oncol. 2022, 40, e16540. [Google Scholar] [CrossRef]
  139. Nielsen, T.K.; Højgaard, M.; Andersen, J.T.; Jørgensen, N.R.; Zerahn, B.; Kristensen, B.; Henriksen, T.; Lykkesfeldt, J.; Mikines, K.J.; Poulsen, H.E. Weekly ascorbic acid infusion in castration-resistant prostate cancer patients: A single-arm phase II trial. Transl. Androl. Urol. 2017, 6, 517–528. [Google Scholar] [CrossRef]
  140. Nielsen, T.K.; Højgaard, M.; Andersen, J.T.; Poulsen, H.E.; Lykkesfeldt, J.; Mikines, K.J. Elimination of Ascorbic Acid After High-Dose Infusion in Prostate Cancer Patients: A Pharmacokinetic Evaluation. Basic Clin. Pharmacol. Toxicol. 2014, 116, 343–348. [Google Scholar] [CrossRef]
  141. Welsh, J.L.; Wagner, B.A.; Van’t Erve, T.J.; Zehr, P.S.; Berg, D.J.; Halfdanarson, T.R.; Yee, N.S.; Bodeker, K.L.; Du, J.; Roberts, L.J., II; et al. Pharmacological ascorbate with gemcitabine for the control of metastatic and node-positive pancreatic cancer (PACMAN): Results from a phase I clinical trial. Cancer Chemother. Pharmacol. 2013, 71, 765–775. [Google Scholar] [CrossRef] [PubMed]
  142. Monti, D.A.; Mitchell, E.; Bazzan, A.J.; Littman, S.; Zabrecky, G.; Yeo, C.J.; Pillai, M.V.; Newberg, A.B.; Deshmukh, S.; Levine, M. Phase I Evaluation of Intravenous Ascorbic Acid in Combination with Gemcitabine and Erlotinib in Patients with Metastatic Pancreatic Cancer. PLoS ONE 2012, 7, e29794. [Google Scholar] [CrossRef] [PubMed]
  143. Ou, J.; Zhu, X.; Chen, P.; Du, Y.; Lu, Y.; Peng, X.; Bao, S.; Wang, J.; Zhang, X.; Zhang, T.; et al. A randomized phase II trial of best supportive care with or without hyperthermia and vitamin C for heavily pretreated, advanced, refractory non-small-cell lung cancer. J. Adv. Res. 2020, 24, 175–182. [Google Scholar] [CrossRef]
  144. ClinicalTrials.Gov Identifier: NCT01754987. A Study of Vitamin C in the Treatment of Liver Cancer to Determine If It Is Safe and Effective. (2012). Available online: https://clinicaltrials.gov/ct2/show/NCT01754987 (accessed on 25 September 2024).
  145. Villagran, M.; Ferreira, J.; Martorell, M.; Mardones, L. The Role of Vitamin C in Cancer Prevention and Therapy: A Literature Review. Antioxidants 2021, 10, 1894. [Google Scholar] [CrossRef]
  146. Freitas, B.d.J.e.S.d.A.; Lloret, G.R.; Visacri, M.B.; Tuan, B.T.; Amaral, L.S.; Baldini, D.; de Sousa, V.M.; de Castro, L.L.; Aguiar, J.R.S.; Pincinato, E.d.C.; et al. High 15-F2t-Isoprostane Levels in Patients with a Previous History of Nonmelanoma Skin Cancer: The Effects of Supplementary Antioxidant Therapy. BioMed Res. Int. 2015, 2015, 1–8. [Google Scholar] [CrossRef] [PubMed]
  147. Skorupan, N.; Ahmad, M.I.; Steinberg, S.M.; Trepel, J.B.; Cridebring, D.; Han, H.; Von Hoff, D.D.; Alewine, C. A phase II trial of the super-enhancer inhibitor Minnelide™ in advanced refractory adenosquamous carcinoma of the pancreas. Futur. Oncol. 2022, 18, 2475–2481. [Google Scholar] [CrossRef]
  148. Zhang, C.Y.; Yang, M. Anti-oxidative stress treatment and current clinical trials. World J. Hepatol. 2024, 16, 294–299. [Google Scholar] [CrossRef]
  149. Luo, M.; Zhou, L.; Huang, Z.; Li, B.; Nice, E.C.; Xu, J.; Huang, C. Antioxidant Therapy in Cancer: Rationale and Progress. Antioxidants 2022, 11, 1128. [Google Scholar] [CrossRef]
  150. Lin, L.; Wu, Q.; Lu, F.; Lei, J.; Zhou, Y.; Liu, Y.; Zhu, N.; Yu, Y.; Ning, Z.; She, T.; et al. Nrf2 signaling pathway: Current status and potential therapeutic targetable role in human cancers. Front. Oncol. 2023, 13, 1184079. [Google Scholar] [CrossRef]
  151. Hecht, F.; Zocchi, M.; Alimohammadi, F.; Harris, I.S. Regulation of antioxidants in cancer. Mol. Cell 2024, 84, 23–33. [Google Scholar] [CrossRef]
  152. Singh, S.V.; Srivastava, S.K.; Choi, S.; Lew, K.L.; Antosiewicz, J.; Xiao, D.; Zeng, Y.; Watkins, S.C.; Johnson, C.S.; Trump, D.L.; et al. Sulforaphane-induced cell death in human prostate cancer cells is initiated by reactive oxygen species. J. Biol. Chem. 2005, 280, 19911–19924. [Google Scholar] [CrossRef] [PubMed]
  153. Atwell, L.L.; Zhang, Z.; Mori, M.; Farris, P.; Vetto, J.T.; Naik, A.M.; Oh, K.Y.; Thuillier, P.; Ho, E.; Shannon, J. Sulforaphane Bioavailability and Chemopreventive Activity in Women Scheduled for Breast Biopsy. Cancer Prev. Res. 2015, 8, 1184–1191. [Google Scholar] [CrossRef] [PubMed]
  154. Ge, M.; Zhang, L.; Cao, L.; Xie, C.; Li, X.; Li, Y.; Meng, Y.; Chen, Y.; Wang, X.; Chen, J.; et al. Sulforaphane inhibits gastric cancer stem cells via suppressing sonic hedgehog pathway. Int. J. Food Sci. Nutr. 2019, 70, 570–578. [Google Scholar] [CrossRef] [PubMed]
  155. Li, S.; Khoi, P.N.; Yin, H.; Sah, D.K.; Kim, N.H.; Lian, S.; Jung, Y.D. Sulforaphane Suppresses the Nicotine-Induced Expression of the Matrix Metalloproteinase-9 via Inhibiting ROS-Mediated AP-1 and NF-kappaB Signaling in Human Gastric Cancer Cells. Int. J. Mol. Sci. 2022, 23, 5172. [Google Scholar] [CrossRef] [PubMed]
  156. Honma, M.; Yamada, M.; Yasui, M.; Horibata, K.; Sugiyama, K.I.; Masumura, K. In vivo and in vitro mutagenicity of perillaldehyde and cinnamaldehyde. Genes Environ. 2021, 43, 30. [Google Scholar] [CrossRef]
  157. Puschhof, J.; Pleguezuelos-Manzano, C.; Martinez-Silgado, A.; Akkerman, N.; Saftien, A.; Boot, C.; de Waal, A.; Beumer, J.; Dutta, D.; Heo, I.; et al. Intestinal organoid cocultures with microbes. Nat. Protoc. 2021, 16, 4633–4649. [Google Scholar] [CrossRef]
  158. Murai, K.; Dentro, S.; Ong, S.H.; Sood, R.; Fernandez-Antoran, D.; Herms, A.; Kostiou, V.; Abnizova, I.; Hall, B.A.; Gerstung, M.; et al. p53 mutation in normal esophagus promotes multiple stages of carcinogenesis but is constrained by clonal competition. Nat. Commun. 2022, 13, 6206. [Google Scholar] [CrossRef]
  159. Sabapathy, K.; Lane, D.P. Therapeutic targeting of p53: All mutants are equal, but some mutants are more equal than others. Nat. Rev. Clin. Oncol. 2018, 15, 13–30. [Google Scholar] [CrossRef]
  160. Liu, K.; Chen, W.; Lei, S.; Xiong, L.; Zhao, H.; Liang, D.; Lei, Z.; Zhou, N.; Yao, H.; Liang, Y. Wild-type and mutant p53 differentially modulate miR-124/iASPP feedback following pohotodynamic therapy in human colon cancer cell line. Cell Death Dis. 2017, 8, e3096. [Google Scholar] [CrossRef]
  161. Lisek, K.; Campaner, E.; Ciani, Y.; Walerych, D.; Del Sal, G. Mutant p53 tunes the NRF2-dependent antioxidant response to support survival of cancer cells. Oncotarget 2018, 9, 20508–20523. [Google Scholar] [CrossRef]
  162. Walerych, D.; Lisek, K.; Sommaggio, R.; Piazza, S.; Ciani, Y.; Dalla, E.; Rajkowska, K.; Gaweda-Walerych, K.; Ingallina, E.; Tonelli, C.; et al. Proteasome machinery is instrumental in a common gain-of-function program of the p53 missense mutants in cancer. Nat. Cell Biol. 2016, 18, 897–909. [Google Scholar] [CrossRef]
  163. Tung, M.C.; Lin, P.L.; Wang, Y.C.; He, T.Y.; Lee, M.C.; Yeh, S.D.; Chen, C.Y.; Lee, H. Mutant p53 confers chemoresistance in non-small cell lung cancer by upregulating Nrf2. Oncotarget 2015, 6, 41692–41705. [Google Scholar] [CrossRef] [PubMed]
  164. Rushworth, S.A.; Zaitseva, L.; Murray, M.Y.; Shah, N.M.; Bowles, K.M.; MacEwan, D.J. The high Nrf2 expression in human acute myeloid leukemia is driven by NF-kappaB and underlies its chemo-resistance. Blood 2012, 120, 5188–5198. [Google Scholar] [CrossRef] [PubMed]
  165. Asher, G.; Lotem, J.; Kama, R.; Sachs, L.; Shaul, Y. NQO1 stabilizes p53 through a distinct pathway. Proc. Natl. Acad. Sci. USA 2002, 99, 3099–3104. [Google Scholar] [CrossRef] [PubMed]
  166. Asher, G.; Lotem, J.; Tsvetkov, P.; Reiss, V.; Sachs, L.; Shaul, Y. P53 hot-spot mutants are resistant to ubiquitin-independent degradation by increased binding to NAD(P)H:quinone oxidoreductase 1. Proc. Natl. Acad. Sci. USA 2003, 100, 15065–15070. [Google Scholar] [CrossRef] [PubMed]
  167. Kalo, E.; Kogan-Sakin, I.; Solomon, H.; Bar-Nathan, E.; Shay, M.; Shetzer, Y.; Dekel, E.; Goldfinger, N.; Buganim, Y.; Stambolsky, P.; et al. Mutant p53R273H attenuates the expression of phase 2 detoxifying enzymes and promotes the survival of cells with high levels of reactive oxygen species. J. Cell Sci. 2012, 125, 5578–5586. [Google Scholar]
  168. Gomes, A.S.; Ramos, H.; Soares, J.; Saraiva, L. p53 and glucose metabolism: An orchestra to be directed in cancer therapy. Pharmacol. Res. 2018, 131, 75–86. [Google Scholar] [CrossRef]
  169. Bykov, V.J.N.; Eriksson, S.E.; Bianchi, J.; Wiman, K.G. Targeting mutant p53 for efficient cancer therapy. Nat. Rev. Cancer 2018, 18, 89–102. [Google Scholar] [CrossRef]
  170. Peng, X.; Zhang, M.Q.; Conserva, F.; Hosny, G.; Selivanova, G.; Bykov, V.J.; Arner, E.S.; Wiman, K.G. APR-246/PRIMA-1MET inhibits thioredoxin reductase 1 and converts the enzyme to a dedicated NADPH oxidase. Cell Death Dis. 2013, 4, e881. [Google Scholar] [CrossRef]
  171. Haffo, L.; Lu, J.; Bykov, V.J.N.; Martin, S.S.; Ren, X.; Coppo, L.; Wiman, K.G.; Holmgren, A. Inhibition of the glutaredoxin and thioredoxin systems and ribonucleotide reductase by mutant p53-targeting compound APR-246. Sci. Rep. 2018, 8, 12671. [Google Scholar] [CrossRef]
  172. Tessoulin, B.; Descamps, G.; Moreau, P.; Maiga, S.; Lode, L.; Godon, C.; Marionneau-Lambot, S.; Oullier, T.; Le Gouill, S.; Amiot, M.; et al. PRIMA-1Met induces myeloma cell death independent of p53 by impairing the GSH/ROS balance. Blood 2014, 124, 1626–1636. [Google Scholar] [CrossRef] [PubMed]
  173. Mohell, N.; Alfredsson, J.; Fransson, A.; Uustalu, M.; Bystrom, S.; Gullbo, J.; Hallberg, A.; Bykov, V.J.; Bjorklund, U.; Wiman, K.G. APR-246 overcomes resistance to cisplatin and doxorubicin in ovarian cancer cells. Cell Death Dis. 2015, 6, e1794. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, Z.; Hu, H.; Heitink, L.; Rogers, K.; You, Y.; Tan, T.; Suen, C.L.W.; Garnham, A.; Chen, H.; Lieschke, E.; et al. The anti-cancer agent APR-246 can activate several programmed cell death processes to kill malignant cells. Cell Death Differ. 2023, 30, 1033–1046. [Google Scholar] [CrossRef] [PubMed]
  175. Faraonio, R.; Vergara, P.; Di Marzo, D.; Pierantoni, M.G.; Napolitano, M.; Russo, T.; Cimino, F. p53 suppresses the Nrf2-dependent transcription of antioxidant response genes. J. Biol. Chem. 2006, 281, 39776–39784. [Google Scholar] [CrossRef]
  176. Liang, Y.; Liu, J.; Feng, Z. The regulation of cellular metabolism by tumor suppressor p53. Cell Biosci. 2013, 3, 9. [Google Scholar] [CrossRef]
  177. Li, T.; Kon, N.; Jiang, L.; Tan, M.; Ludwig, T.; Zhao, Y.; Baer, R.; Gu, W. Tumor suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell 2012, 149, 1269–1283. [Google Scholar] [CrossRef]
  178. Yang, H.; Xiang, S.; Kazi, A.; Sebti, S.M. The GTPase KRAS suppresses the p53 tumor suppressor by activating the NRF2-regulated antioxidant defense system in cancer cells. J. Biol. Chem. 2020, 295, 3055–3063. [Google Scholar] [CrossRef]
  179. Kung, C.P.; Weber, J.D. It’s Getting Complicated-A Fresh Look at p53-MDM2-ARF Triangle in Tumorigenesis and Cancer Therapy. Front. Cell Dev. Biol. 2022, 10, 818744. [Google Scholar] [CrossRef] [PubMed]
  180. Xu, Y.; Jin, C.; Liu, Z.; Pan, J.; Li, H.; Zhang, Z.; Bi, S.; Yokoyama, K.K. Cloning and characterization of the mouse JDP2 gene promoter reveal negative regulation by p53. Biochem. Biophys. Res. Commun. 2014, 450, 1531–1536. [Google Scholar] [CrossRef]
  181. Tanigawa, S.; Lee, C.H.; Lin, C.S.; Ku, C.C.; Hasegawa, H.; Qin, S.; Kawahara, A.; Korenori, Y.; Miyamori, K.; Noguchi, M.; et al. Jun dimerization protein 2 is a critical component of the Nrf2/MafK complex regulating the response to ROS homeostasis. Cell Death Dis. 2013, 4, e921. [Google Scholar] [CrossRef]
  182. Jin, C.; Kato, K.; Chimura, T.; Yamasaki, T.; Nakade, K.; Murata, T.; Li, H.; Pan, J.; Zhao, M.; Sun, K.; et al. Regulation of histone acetylation and nucleosome assembly by transcription factor JDP2. Nat. Struct. Mol. Biol. 2006, 13, 331–338. [Google Scholar] [CrossRef] [PubMed]
  183. Pan, J.; Nakade, K.; Huang, Y.C.; Zhu, Z.W.; Masuzaki, S.; Hasegawa, H.; Murata, T.; Yoshiki, A.; Yamaguchi, N.; Lee, C.H.; et al. Suppression of cell-cycle progression by Jun dimerization protein-2 (JDP2) involves downregulation of cyclin-A2. Oncogene 2010, 29, 6245–6256. [Google Scholar] [CrossRef] [PubMed]
  184. Dinkova-Kostova, A.T.; Copple, I.M. Advances ad challenges in therapeutic targeting of NRF2. Trends Pharmocol Sci. 2023, 44, 137–149. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic model of ROS balance induced by phase I enzyme ligands and phase II enzyme ligands. Phase I enzyme ligands or phase II enzyme metabolites are able to induce the AhR activation, to generate ROS production. By contrast, the phase II enzyme ligands induce Nrf2 activation to decrease the level of AhR and ROS production, which results in the phase II enzyme metabolites. Thus, ROS balance is controlled by AhR and Nrf2. Cyp1A1 (p450 family); cytochrome p450 family Cyp1A1, NOX; nitrogen oxides, COX; cyclooxygenase, AKR; aldo-keto reductase, NQO1; NAD(P)H dehydrogenase (quinone) family, GSTA1/2; glutathione S transferase alpha 1 and 2, UGT1A; UDP glucuronosyltransferase family 1 member A complex locus. Black arrow indicated the decreased level of AhR and white arrow indicated the increased level of Nrf2 expression.
Figure 1. Schematic model of ROS balance induced by phase I enzyme ligands and phase II enzyme ligands. Phase I enzyme ligands or phase II enzyme metabolites are able to induce the AhR activation, to generate ROS production. By contrast, the phase II enzyme ligands induce Nrf2 activation to decrease the level of AhR and ROS production, which results in the phase II enzyme metabolites. Thus, ROS balance is controlled by AhR and Nrf2. Cyp1A1 (p450 family); cytochrome p450 family Cyp1A1, NOX; nitrogen oxides, COX; cyclooxygenase, AKR; aldo-keto reductase, NQO1; NAD(P)H dehydrogenase (quinone) family, GSTA1/2; glutathione S transferase alpha 1 and 2, UGT1A; UDP glucuronosyltransferase family 1 member A complex locus. Black arrow indicated the decreased level of AhR and white arrow indicated the increased level of Nrf2 expression.
Cells 13 01648 g001
Figure 2. Schematic representation of the actin mechanism of the NRF2 signaling pathway. Under basal conditions, a KEAP1/CUL3/RBX1 complex binds to NRF2 and promotes ubiquitination to commit proteasomal degradation. Upon activation of NRF2 by electrophiles or reactive oxygen species (ROS) and other ligands, NRF2 enters the nucleus to associate with sMAF, p21Cip1, JDP2, binds to ARE cis-element, and modulates the expression of several gene families such as anti-oxidation, detoxification, anti-inflammation, reprogramming and cell fate decision, and cell cycle. GCLC (glutamate-cysteine ligase catalytic subunit), GCLM (glutamate-cysteine ligase modifier subunit), GPX4 (glutathione peroxidase 4), SLC7A11 (solute carrier family 7 member 11), TXN1 (thioredoxin1), SDO1 (super oxide dismutase type 1), HO-1 (heme oxygenase 1), NQO1(NAD(P)H quinone dehydrogenase 1), UGT1A1 (UDP glucuronosyltransferase family 1 member A1), SULTA1 (Sulfotransferase family 1A member 1), ABCB6 (ATP-binding cassette superfamily B member 6), IL (interleukin), HIF 1α (hypoxia inducible factor 1α), BMP6 (bone morphogenic protein 6), NOTCH 1 (neurogenic locus notch homolog protein 1), PPARγ (peroxisome proliferator-activated receptor gamma), and MyoD1(myogenic differentiation 1).The circle indicated the transcriptional event in the nucleus.
Figure 2. Schematic representation of the actin mechanism of the NRF2 signaling pathway. Under basal conditions, a KEAP1/CUL3/RBX1 complex binds to NRF2 and promotes ubiquitination to commit proteasomal degradation. Upon activation of NRF2 by electrophiles or reactive oxygen species (ROS) and other ligands, NRF2 enters the nucleus to associate with sMAF, p21Cip1, JDP2, binds to ARE cis-element, and modulates the expression of several gene families such as anti-oxidation, detoxification, anti-inflammation, reprogramming and cell fate decision, and cell cycle. GCLC (glutamate-cysteine ligase catalytic subunit), GCLM (glutamate-cysteine ligase modifier subunit), GPX4 (glutathione peroxidase 4), SLC7A11 (solute carrier family 7 member 11), TXN1 (thioredoxin1), SDO1 (super oxide dismutase type 1), HO-1 (heme oxygenase 1), NQO1(NAD(P)H quinone dehydrogenase 1), UGT1A1 (UDP glucuronosyltransferase family 1 member A1), SULTA1 (Sulfotransferase family 1A member 1), ABCB6 (ATP-binding cassette superfamily B member 6), IL (interleukin), HIF 1α (hypoxia inducible factor 1α), BMP6 (bone morphogenic protein 6), NOTCH 1 (neurogenic locus notch homolog protein 1), PPARγ (peroxisome proliferator-activated receptor gamma), and MyoD1(myogenic differentiation 1).The circle indicated the transcriptional event in the nucleus.
Cells 13 01648 g002
Figure 3. Divergent functions of PER, CIN, and SNF to control gene expressions of AHR and NRF2. The heterogeneous function of phase II enzyme ligands to affect the ROS level and Nrf2 gene expression is illustrated. The p53 mutation status and the normal and cancer cells context are critical for the bifunctional effects of antioxidation drugs to control the balance of ROS. Thus, the endogenous ROS in the cells we targeted should be examined. Higher ROS levels decreased the expression levels of Nrf2, p53, and Jdp2, and lower ROS decreased the generation of ROS and increased p53 and Jdp2 levels. The ROS level was controlled by the AhR expression, and antioxidation was controlled by Nrf2-mediated responses. PER; perillaldehyde, CIN; cinnamon aldehyde, SFN; sulforaphane, ARE; antioxidative response element. The white arrows indicated the increased levels of genes or functions, and the black arrows indicated the decreased levels of genes or functions.
Figure 3. Divergent functions of PER, CIN, and SNF to control gene expressions of AHR and NRF2. The heterogeneous function of phase II enzyme ligands to affect the ROS level and Nrf2 gene expression is illustrated. The p53 mutation status and the normal and cancer cells context are critical for the bifunctional effects of antioxidation drugs to control the balance of ROS. Thus, the endogenous ROS in the cells we targeted should be examined. Higher ROS levels decreased the expression levels of Nrf2, p53, and Jdp2, and lower ROS decreased the generation of ROS and increased p53 and Jdp2 levels. The ROS level was controlled by the AhR expression, and antioxidation was controlled by Nrf2-mediated responses. PER; perillaldehyde, CIN; cinnamon aldehyde, SFN; sulforaphane, ARE; antioxidative response element. The white arrows indicated the increased levels of genes or functions, and the black arrows indicated the decreased levels of genes or functions.
Cells 13 01648 g003
Figure 4. IPA analysis of the signaling networks between AhR signaling and NRF2 signaling. The canonical pathways of aryl hydrocarbon receptor (AhR) signaling (blue arrow), NRF2-mediated oxidative stress response pathway (orange arrow), and KEAP1-NEF2L2 (NRF2) pathway (yellow arrow) are selected from the Ingenuity Pathway Analysis (IPA) system (https://www.nihlibrary.nih.gov/resources/tools/ingenuity-pathways-analysis-ipa accessed on 30 August 2024). For the construction of signaling networks, AhR signaling is considered as the major axis, and the other two pathways are indicated with “CP: KEAP1-NEF2L2 Pathway” and “CP: NRF2-mediated Oxidative Stress Response”, respectively. The web-based IPA bioinformatic software is applied to overlay with each signaling pathway and to indicate the crosstalk molecules by the cross line between two signaling pathways.
Figure 4. IPA analysis of the signaling networks between AhR signaling and NRF2 signaling. The canonical pathways of aryl hydrocarbon receptor (AhR) signaling (blue arrow), NRF2-mediated oxidative stress response pathway (orange arrow), and KEAP1-NEF2L2 (NRF2) pathway (yellow arrow) are selected from the Ingenuity Pathway Analysis (IPA) system (https://www.nihlibrary.nih.gov/resources/tools/ingenuity-pathways-analysis-ipa accessed on 30 August 2024). For the construction of signaling networks, AhR signaling is considered as the major axis, and the other two pathways are indicated with “CP: KEAP1-NEF2L2 Pathway” and “CP: NRF2-mediated Oxidative Stress Response”, respectively. The web-based IPA bioinformatic software is applied to overlay with each signaling pathway and to indicate the crosstalk molecules by the cross line between two signaling pathways.
Cells 13 01648 g004
Table 2. Current summary of preclinical and clinical trials of AHR and NRF2 targeted drugs against cancers and diseases.
Table 2. Current summary of preclinical and clinical trials of AHR and NRF2 targeted drugs against cancers and diseases.
DrugsMode of FunctionDiseases & CancerClinical TrialPerformance
[AHR Target]
IK-175Inhibitor
  • Urothelial Cancer
  • Advanced solid tumor
NCT104200963[34]
BAY-2416964Antagonist
  • Head neck cancer
  • Lung cancer
  • Colon cancer
NCT04069026[119]
NCT04069036[119]
SorafenibAntagonist
  • Liver cancer
  • Ovarian cancer
  • Renal cell cancer
[120,121,122]
AFP464
(Amino flavone)
Antagonist
  • Solid tumors
5F203
(Benzothiazole)
Antagonist
  • Renal Cancer
JTE061
Tapinarof
Agonist
  • Atopic dermatitis
  • Plaque psoriasis
NCT05680740[115,123]
NCT05142774[115,123]
[NRF2 Target]
OmaveloxoloneAntagonist
  • The first FDA-approved drug for Friedreich’s ataxia (FA)
NCT02255436[124]
N-acethycysteineAgonist
  • Melanoma
NCT01612221[125]
  • Leukemia
NCT05611086[126]
  • SARS-CoV-2-infection
NCT04792021[127]
  • HIV infection
NCT01962961[128]
MelatoninAgonist
  • Necrotizing enterocolitis
NCT05033639[129]
  • Multiple sclerosis
NCT02463318
Melatonin + Vitamin C and E, N-Acethycysteine Agonist
  • Septic shock
NCT03557229[130]
Curcumin paclitaxelAgonist
  • Chemotherapy peripheral neuropathy
NCT05966441
LuteolinInhibitor
  • Tongue carcinoma
NCT03288298[131]
HalofuginoneInhibitor
  • Unspecified adult solid tumor
NCT00027677[132]
  • AIDS-related Kaposi sarcoma
NCT00064142[132]
BerberineInhibitor
  • Colorectal adenoma
NCT03281096
  • G-R nonsmall lung adenocarcinoma
NCT03486496[133]
  • Gastric Cancer (Heliobacter pylori)
NCT04697186[134]
  • Gastric Cancer (Heliobacter pylori)
NCT03609892[135]
  • Colorectal adenoma
NCT02226185[136]
  • Colorectal adenoma
NCT03333265[137]
Ascorbic acidInhibitor
  • Breast Cancer
NCT03175341
  • Bladder Cancer
NCT04046094[138]
Ascorbic acid
+ mFOL FOX6
Inhibitor
  • Gastric Cancer
NCT03015675
Vitamin CInhibitor
  • Prostatic neoplasm
NCT01080352[139,140]
Gemcitabine + Ascorbic acidInhibitor
  • Pancreatic neoplasm
NCT01049880[141,142]
Vitamin C + SupplementInhibitor
  • Non-small cell lung cancer
NCT02655913[143]
Ascorbic acid + SorafenibInhibitor
  • Metastatic Hepatocarcinoma
NCT01754987[144]
Vitamin C, E and ZincInhibitor
  • Skin neoplasms
NCT02248584[145,146]
Biopsy, Osimextinb Triptolide analogInhibitor
  • Advanced lung non small cell carcinoma
NCT05166616
MinnelideTM capulesInhibitor
  • Gastric, breast, pancreatic, prostate colorectal, stomach cancer
NCT03129139
MinnelideInhibitor
  • Adeno squamous II carcinoma of the pancreas
NCT04896073[147]
  • Pancreatic Cancer
NCT03117920
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lin, Y.-C.; Ku, C.-C.; Wuputra, K.; Wu, D.-C.; Yokoyama, K.K. Vulnerability of Antioxidant Drug Therapies on Targeting the Nrf2-Trp53-Jdp2 Axis in Controlling Tumorigenesis. Cells 2024, 13, 1648. https://doi.org/10.3390/cells13191648

AMA Style

Lin Y-C, Ku C-C, Wuputra K, Wu D-C, Yokoyama KK. Vulnerability of Antioxidant Drug Therapies on Targeting the Nrf2-Trp53-Jdp2 Axis in Controlling Tumorigenesis. Cells. 2024; 13(19):1648. https://doi.org/10.3390/cells13191648

Chicago/Turabian Style

Lin, Ying-Chu, Chia-Chen Ku, Kenly Wuputra, Deng-Chyang Wu, and Kazunari K. Yokoyama. 2024. "Vulnerability of Antioxidant Drug Therapies on Targeting the Nrf2-Trp53-Jdp2 Axis in Controlling Tumorigenesis" Cells 13, no. 19: 1648. https://doi.org/10.3390/cells13191648

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

Article Metrics

Back to TopTop