Next Article in Journal
Cancer-Associated Fibroblasts as a Common Orchestrator of Therapy Resistance in Lung and Pancreatic Cancer
Next Article in Special Issue
Plasmin and Plasminogen System in the Tumor Microenvironment: Implications for Cancer Diagnosis, Prognosis, and Therapy
Previous Article in Journal
Fatigue in Cancer Patients in Palliative Care—A Review on Pharmacological Interventions
Previous Article in Special Issue
The Role of Hypoxia in Glioblastoma Radiotherapy Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Stress in the Tumor Microenvironment and Its Relevance to Cancer Immunotherapy

1
Georgia Cancer Center, Medical College of Georgia, Augusta University, Augusta, GA 30912, USA
2
The Graduate School, Augusta University, Augusta, GA 30912, USA
3
The Center for Undergraduate Research and Scholarship, Augusta University, Augusta, GA 30912, USA
4
Department of Biochemistry and Molecular Biology, Medical College of Georgia, Augusta University, Augusta, GA 30912, USA
5
Department of Medicine, Medical College of Georgia, Augusta University, Augusta, GA 30912, USA
*
Author to whom correspondence should be addressed.
Cancers 2021, 13(5), 986; https://doi.org/10.3390/cancers13050986
Submission received: 27 January 2021 / Revised: 20 February 2021 / Accepted: 23 February 2021 / Published: 27 February 2021

Abstract

:

Simple Summary

Cancer cells are consistently under oxidative stress, as reflected by elevated basal level of reactive oxygen species (ROS), due to increased metabolism driven by aberrant cell growth. This feature has been exploited to develop therapeutic strategies that control tumor growth by modulating the oxidative stress in tumor cells. This review provides an overview of recent advances in cancer therapies targeting tumor oxidative stress, and highlights the emerging evidence implicating the effectiveness of cancer immunotherapies in intensifying tumor oxidative stress. The promises and challenges of combining ROS-inducing agents with cancer immunotherapy are also discussed.

Abstract

It has been well-established that cancer cells are under constant oxidative stress, as reflected by elevated basal level of reactive oxygen species (ROS), due to increased metabolism driven by aberrant cell growth. Cancer cells can adapt to maintain redox homeostasis through a variety of mechanisms. The prevalent perception about ROS is that they are one of the key drivers promoting tumor initiation, progression, metastasis, and drug resistance. Based on this notion, numerous antioxidants that aim to mitigate tumor oxidative stress have been tested for cancer prevention or treatment, although the effectiveness of this strategy has yet to be established. In recent years, it has been increasingly appreciated that ROS have a complex, multifaceted role in the tumor microenvironment (TME), and that tumor redox can be targeted to amplify oxidative stress inside the tumor to cause tumor destruction. Accumulating evidence indicates that cancer immunotherapies can alter tumor redox to intensify tumor oxidative stress, resulting in ROS-dependent tumor rejection. Herein we review the recent progresses regarding the impact of ROS on cancer cells and various immune cells in the TME, and discuss the emerging ROS-modulating strategies that can be used in combination with cancer immunotherapies to achieve enhanced antitumor effects.

1. Introduction

Reactive oxygen species (ROS) are a group of highly reactive oxygen-containing molecules, including free radicals such as hydroxyl (HO), superoxide (O2), peroxides (RO) and oxides of nitrogen (NO) and the non-radical hydrogen peroxide (H2O2). ROS are physiologically generated as a byproduct of cellular respiration and aerobic metabolism, pathologically elevated in diseases like inflammation and cancer, and exogenously formulated after exposure to xenobiotics such as chemotherapy, radiotherapy, or UV. At low to medium levels, ROS can act as cellular signaling messengers, involved in regulating a variety of cellular functions including gene expression, cell proliferation and differentiation, and immunity against diseases. At high levels, ROS cause oxidative damage to DNA, proteins, and lipids, and become detrimental to cells. Due to the multifaceted role of ROS in cell survival and function, the cellular levels of ROS have to be tightly controlled to maintain the redox homeostasis, i.e., the balance between ROS production and scavenging, through multi-layer mechanisms. Oxidative stress occurs when this balance is disrupted in cells. The ontogeny, regulation, and biological function of oxidative stress in cancer biology have been extensively reviewed by others [1,2,3,4]. In this review, we mainly discuss the impact of oxidative stress on the tumor microenvironment (TME), including cancer cells and various immune cells. By focusing on how the interplays between cancer cells and immune cells influence the redox status of both populations, we highlight the therapeutic potential of rational combination of ROS-modulating agents with cancer immunotherapies.

2. The Impact of Oxidative Stress on Cancer Cells

It has been well-established that cancer cells are under higher degree of basal level oxidative stress than normal cells, reflected by an increased presence of ROS. Mitochondria are the major cellular source of ROS production. Mitochondria produce ROS during respiration as a natural by-product of electron transport chain (ETC) activity. Incomplete electron transfer and leakage of electrons through ETC complexes I, II, and III results in superoxide production [5]. Membrane-bound NADPH oxidases (NOXs) are another important source of ROS. NOXs are a family of hetero-oligomeric enzymes that catalyze the production of superoxide from O2 and NADPH. In most mammals, there are seven NOX isoforms: NOX1, NOX2, NOX3, NOX4, NOX5, dual oxidase (DUOX) 1, and DUOX2 [3,6]. Deregulated ROS generation in cancer cells may occur due to cell-intrinsic events such as oncogene activation, tumor suppressor gene inactivation, increased metabolism, and adaptation to hypoxia (i.e., low oxygen levels), or exogenous insults such as chemotherapy and ionizing radiation [2,3,7,8,9].

2.1. ROS in Tumor Initiation, Progression, and Survival

Mildly increased levels of ROS are known to contribute to tumor progression by promoting cell transformation [10], proliferation [11], and survival [12,13,14]. It has been well-documented that growth factor signaling and oncogenic mutations can result in increased ROS production, which is tightly associated with the incidence of various cancers. For example, platelet-derived growth factor (PDGF), epidermal growth factor (EGF), tumor necrosis factor α (TNFα), interleukin-1 (IL-1), transforming growth factor β (TGFβ), etc. can stimulate ROS production and promote tumor progression [15,16,17,18,19]. Oncogenic mutations in RAS have been shown to cause increased generation of superoxide [20,21,22,23]. The oncogene-induced ROS can hyperactivate two important pathways: PI3K/Akt/mTOR and MAPK/ERK signaling cascades [20,24,25]. The PI3K/Akt/mTOR pathway critically regulates cell survival. ROS can activate this pathway by oxidizing and inactivating its negative phosphatase regulators, including phosphatase and tensin homolog (PTEN), protein-tyrosine phosphatase 1B (PTP1B), and protein phosphatase 2 (PP2A), or by the direct oxidation of kinases [26,27]. Many solid tumors, including glioblastoma, melanoma, prostate, and breast cancer, are frequently marked by inactivation of PTEN [28,29,30], suggesting that ROS-induced hyperactivation of the PI3K/Akt survival pathway is critical to the development of these cancers. MAPK signaling pathways are involved in cell growth, differentiation, and survival. Similar to PI3K/Akt/mTOR pathway hyperactivation, ROS induce MAPK/ERK-mediated proliferative signaling through oxidizing and inactivating MAPK phosphatases [31,32]. It is also worth mentioning that ROS can induce nuclear translocation of NF-κB through oxidation and degradation of IκB, the phosphatase inhibitor of NF-κB [33,34,35]. NF-κB is a transcription factor that regulates the genes responsible for inflammation, cell proliferation, differentiation, and survival and is known to promote tumorigenesis, angiogenesis, and metastasis [36,37]. Altogether these studies underscore the correlation between the aberrant cell signaling events and the deregulated ROS generation in cancers [38].

2.2. ROS in Tumor Angiogenesis, Metastasis, and Chemoresistance

Increased ROS also facilitate cancer cell angiogenesis [39], metastasis [40,41], and chemoresistance [12]. To meet the increased metabolic needs of proliferating cancer cells, new blood vessels are established to enhance oxygen and nutrient supplies. It is well-known that ROS promote blood vessel formation and angiogenesis [39,42,43]. Tumor hypoxia, a condition in which tumor cells are deprived of oxygen, occurs when tumor growth outpaces blood supply. Hypoxia stimulates the production of mitochondrial ROS (mROS) via the transfer of electrons from ubisemiquinone to molecular oxygen at the Qo site of complex III of the mitochondrial electron transport chain [44,45]. Increased mROS induce and stabilize hypoxia-inducible factor-1 (HIF1a) [44,46,47], a transcription factor that enhances the survival and progression of tumors by upregulating genes regulating tumor angiogenesis, metabolism, metastasis, and chemoresistance [48,49].
Metastasis involves the spread of cancer cells from the primary tumor to the surrounding tissues and to distant organs. Epithelial to mesenchymal transition (EMT) is the process of epithelial cell transition into mesenchymal cell, which is the major cause of tumor metastasis. It has been shown that ROS promote EMT by inducing the expression and activity of certain matrix metalloproteinases (MMPs) that mediate proteolytic degradation of extracellular matrix (ECM) components [50,51]. Cancer stem cells (CSCs) represent a less differentiated but highly tumorigenic subpopulation of cancer cells that contribute to chemoresistance [52]. CSCs are marked by a heightened antioxidant capacity, which allows them to self-renew, differentiate, and importantly, to resist ROS-mediated oxidative damage and cell death induced by radiation or chemotherapy [53,54].
Excessive induction of ROS above a certain threshold can be lethal to the cancer cells [1,2,3]. Cancer cells have an increased antioxidant capacity, mediated by enzymatic and nonenzymatic antioxidants, to adapt to their high oxidative stress status. The main endogenous antioxidant enzymes include superoxide dismutase (SOD), catalase, glutathione peroxidase, glutathione reductase, thioredoxins, peroxiredoxins, etc. The natural nonenzymatic antioxidants include glutathione (GSH), carotenoids, vitamins, etc. One of the vital transcription factors that regulate redox homeostasis in cancer cells is nuclear factor erythroid 2–related factor 2 (NRF2) [55,56]. High levels of ROS prevent the proteasomal degradation of NRF2, thus promoting its nuclear translocation and initiation of the transcription of a multitude of antioxidant genes, including GSH peroxidases (GPXs), and GSH S-transferases (GSTs) [57], glutathione reductase, thioredoxin, thioredoxin reductase, peroxiredoxin and sulfiredoxin. Of note, NRF2 regulates the expression of glutamate-cysteine ligase catalytic (GCLC) and modifier (GCLM) subunits, which combine to form a heterodimer to catalyze the rate-limiting step in GSH biosynthesis. In addition, NRF2 regulates NADPH regeneration enzymes and NAD(P)H:quinone oxidoreductase 1 (NQO1), which inhibits the formation of free radicals by the redox-cycling of quinones [58,59]. NRF2 is therefore considered to be a stress alleviator, supporting cancer cell survival, growth, and escape from the deleterious effects of elevated ROS by maintaining a high but balanced redox status within TME. Moreover, it has been reported that mutations in genes encoding the NRF2 transcription factor and its negative regulator (KEAP1) are frequently detected in cancer [59,60,61]. These mutations may lead to aberrant NRF2 activation, which is associated with poor prognosis and correlates with chemoresistance and tumor recurrence [62,63]. Some recent studies reported that cancer cells can produce neuroglobin (NGB), a monomeric globin, in response to oxidative stress [64]. NGB can act as an oxidative stress sensor and compensatory protein that intersects with the NRF2 pathway to enable tumor cells to resist oxidative stress and acquire chemoresistance [65,66]. Due to its essential role in tumorigenesis, cancer cell proliferation and drug resistance, NRF2 represents a plausible target for anticancer therapy [61,63].

3. The Impact of Oxidative Stress on Immune Cells in the TME

The TME is a dynamic environment in which tumor cells reside and interact with the surrounding vasculature, various immune cells, fibroblasts, and ECM. On the one hand, immunosurveillance mediated by T cells and natural killer (NK) cells can detect and attack transformed cells. ROS are important signal mediators involved in the activation of T cells and NK cells. ROS are also used by neutrophils and macrophages to destroy cancer cells. On the other hand, cancer cells possess the ability to induce tumor-promoting immune cells, including regulatory T cells (Tregs), myeloid-derived suppressor cells (MDSCs), tumor-associated macrophages (TAMs), and tumor-associated neutrophils (TANs). The presence of increased ROS and various types of myeloid cells in the TME is characteristic of chronic inflammation, which is intimately intertwined with cancer development and progression [67]. The crosstalk between inflammatory and oxidative stress mediators may form a positive feed-back loop termed “oxinflammation”, shaping the outcome of antitumor immune responses [68]. ROS in the TME, along with other mechanisms, are used by cancer cells and immunosuppressive cells to create immune tolerance to tumors [69,70,71,72,73,74,75,76]. Here, we focus on the impact of ROS on several types of immune cells with relevance to cancer immunotherapy.

3.1. The Impact of ROS on T Cells and NK Cells

T cell and NK cell activation leads to an increase in ROS production. It has been well-documented that a mild ROS elevation is required for proper T cell activation and differentiation. ROS act as a secondary messenger participating in the activation of nuclear factor of activated T cells (NFAT) and inhibition of negative regulatory phosphatases to ensure appropriate signaling [77,78]. ROS also contribute to activation induced cell death for T cells to maintain immune homeostasis. Kappler and Marrack reported that activation-induced ROS upregulate Fas expression and downregulate antiapoptotic Bcl2 expression to facilitate T cell apoptosis [79]. The levels of ROS in T and NK cells need to be delicately controlled to avoid the detrimental effects of high levels of ROS. Excessive ROS in T and NK cells can decrease TCRζ- and CD16ζ-chain levels, block NF-kB activation, resulting in deficient IFN-γ, TNF-α, and IL-2 production [80,81,82]. The antioxidative GSH pathway plays a critical role in controlling the redox status in T cells. Mak et al. demonstrated that GSH deficiency in T cells has compromised activation of mammalian target of rapamycin-1 (mTOR) and reduces expression of NFAT and Myc transcription factors, resulting in impaired metabolic integration and reprogramming during inflammatory T cell responses [83]. Tumor-specific T cells treated with the antioxidant N-acetyl cysteine (NAC) during activation significantly reduce DNA damage and cell death, and show improved persistence and antitumor effects upon adoptive transfer into tumor-bearing mice [84,85,86,87]. T cells engineered to co-express a chimeric antigen receptor (CAR) and catalase not only reduce oxidative stress in themselves and exert superior antitumor activity, but also protect bystander NK cells from ROS-mediated repression [88]. Likewise, NK cells primed by IL-15 acquire resistance against oxidative stress through the thioredoxin system, and can aid in protecting other lymphocytes from ROS within the TME [89].

3.2. Oxidative Stress and Antigen Presentation

Appropriate levels of ROS are needed for the proper function of antigen-presenting cells. It has been reported that NOX2-mediated phagosomal ROS production in macrophages and dendritic cells (DCs) regulates antigen cross-presentation [90]. Extracellular ROS can also modify the immunogenicity of antigenic peptides, altering T cell priming [91,92]. It has been well-established that induction of oxidative stress in the endoplasmic reticulum (ER) can cause immunogenic death of cancer cells [93,94,95]. Immunogenic cell death (ICD) leads to exhibition and secretion of alarmins, i.e., damage-associated molecular patterns (DAMPs), including adenosine triphosphate (ATP), ER protein calreticulin (CRT) and nuclear heat-shock protein high mobility group box 1 (HMGB1). These DAMPs interact with their receptors (CD91 for CRT, TLR4 for HMGB1, P2RX7 for ATP) on DCs, leading to DC activation, antigen cross-presentation, and ultimately antitumor CD8+ T cell responses. It has been shown that scavenging ROS by antioxidants such as GSH and NAC diminishes ICD [96], whereas strategies that amplify ROS in the ER enhance ICD and augment antitumor immunity [97]. However, there is also evidence that ROS can oxidize the danger signal HMGB1 released from dying cells and thereby neutralize its alarmin activity [98]. A recent study reported that targeted scavenging of extracellular ROS using a tumor ECM targeting nanomaterial can maintain the stimulatory activity of HMGB1 and restore ICD-induced antitumor immunity [99]. These studies suggest that the level and duration of ROS may determine whether or not ICD can occur and lead to effective antitumor immunity.

3.3. Oxidative Stress and Immunosuppressor Cells

ROS are not only involved in the induction of Tregs [100], but are also used by Tregs to suppress other immune cells [101,102,103]. Increased numbers of Tregs are often present at tumor sites, indicating that Tregs can persist in this environment despite increased oxidative stress in the TME. Previous studies have shown that Tregs exhibit resistance to oxidative stress, a phenomenon that may be attributed to their increased antioxidative capacity [104,105]. This is further supported by a report showing that GSH-deficiency in Tregs leads to increased serine metabolism, mTOR activation, and proliferation but downregulated FoxP3, resulting in diminished Treg suppressive function in vitro and in vivo [106]. Intriguingly, Tregs sensitive to oxidative stress have recently been described. Maj et al. reported that tumor-infiltrating Tregs tend to undergo apoptosis due to a weak NRF2-associated antioxidant system and resultant vulnerability to oxidative stress in the TME. The apoptotic Tregs convert a large amount of ATP to immunosuppressive adenosine, which antagonizes spontaneous and immune checkpoint blockade (ICB)-induced antitumor T cell immunity [107].
Myeloid cells, including neutrophils, macrophages and MDSCs, are known to produce high amounts of ROS. ROS released by phagocytic cells, mainly neutrophils and macrophages, contribute to tumor killing after chemoimmunotherapy in animal models [108]. However, neutrophils can also use ROS to suppress T cells [109,110]. MDSCs mediate immune suppression via production of ROS and reactive nitrogen species (RNS) [111]. MDSC-derived ROS and RNS reduce T cell responses by inhibiting T cell receptor (TCR) recognition of its ligand, the MHC-peptide complex on target cells [109,112,113]. It is worth mentioning that high levels of ROS in TME promote the maintenance of MDSCs in an immature and immunosuppressive state, while lacking NOX2 [114] or scavenging H2O2 with catalase [115] promotes immature myeloid cell differentiation into macrophages, resulting in loss of immune suppressive activity of MDSCs. MDSCs are resistant to increased oxidative stress due to an upregulated NRF2-mediated antioxidative system [116], allowing them to exert immunosuppression upon other immune cells through ROS [112,114,117,118,119,120], while protecting themselves from the detrimental effects of ROS.

4. Cancer Therapies Targeting Tumor Redox

At different levels, oxidative stress can exert either pro-tumor or anti-tumor effects, presenting as a double-edged sword to cancer cells. Two opposite strategies have been attempted to modulate tumor redox as a way to prevent or treat cancer. One approach is to reduce the tumor-promoting effects of ROS by attenuating oxidative stress using antioxidants. The other approach is to augment cancer cell death by intensifying the levels of ROS in cancer cells. Here we briefly summarize the current status of the two strategies in cancer therapy. More detailed reviews on this subject can be found elsewhere [7,8,9,121,122,123,124,125].

4.1. The Use of Antioxidants for Cancer Prevention

Given the role of ROS in promoting tumorigenesis, angiogenesis, and metastasis, various antioxidants have been tested as chemopreventive agents based on the rationale that ROS scavenging can reduce the incidence of cancer and/or delay cancer progression [126]. Gao et al. demonstrated that administration of the antioxidant NAC inhibits tumor incidence in mice by suppressing HIF1a-driven tumor growth [127]. Along the same line, other studies showed that overexpression or targeted delivery of SOD, catalase, or glutathione peroxidase can inhibit tumor growth [128,129,130,131,132]. Although encouraging results were observed in some pre-clinical studies, several large-scale clinical trials with dietary antioxidant supplementation such as vitamin A, vitamin E, and β-carotene failed to demonstrate measurable antitumor benefits [133,134]. Paradoxically, in some cases, antioxidant supplementation has been linked with increased rates of certain cancers [135,136,137]. Possible reasons behind the unexpected failure of the antioxidant approach include inefficient scavenging of tumor-promoting ROS in the relevant cellular compartment such as mitochondria, and/or interference with the antitumor roles of ROS in cancer cells [138,139,140,141,142].

4.2. The Use of Pro-Oxidants in Cancer Therapy

Although cancer cells can activate their antioxidant systems to allow them to thrive in the face of increased oxidative stress, they also become more sensitive to further redox disruption. The vulnerability of cancer cells to redox imbalance becomes the Achilles’ heel for cancer. Breaking redox homeostasis in cancer cells can be achieved either by intensifying ROS production or decreasing ROS scavenging through suppressing the antioxidant systems. It is now clear that numerous chemotherapeutic agents exert tumor killing effects through the production of free radicals that cause irreversible cell injury [123,124,143]. Cisplatin, a widely used platinum-based chemotherapy, is known to induce tumor cell apoptosis through generating high levels of cellular superoxide, an effect that can be abolished by the superoxide scavenger Tiron or the antioxidant NAC [144]. 5-fluorouracil (5-FU), an antimetabolite used to treat colon, head, and neck cancers, and other solid tumors, induces tumor cell apoptosis via induction of mitochondrial ROS, and this effect can be blocked by the addition of mitoQ, a mitochondrial-selective antioxidant [145]. Doxorubicin, an anthracycline and topoisomerase inhibitor, induces cancer cell apoptosis as well as cardiotoxicity via direct oxidative DNA damage and indirect induction of H2O2 [146,147]. Chemotherapeutic agents such as taxanes (paclitaxel and docetaxel) and vinca alkaloids (vincristine and vinblastine) promote the release of cytochrome c from the mitochondria and interfere with the electron transport chain, resulting in the production of superoxide radicals and inducing cell death [148,149,150]. ROS induction also contributes to arsenic trioxide’s potent inhibitory effect on acute promyelocytic leukemia [151]. Buthionine sulfoximine (BSO), an inhibitor of glutamate-cysteine ligase (GCL), the enzyme required for GSH synthesis, exhibits anticancer activity by depleting GSH [152,153]. Some tyrosine kinase inhibitors (TKI) widely used as targeted therapy for cancer, including erlotinib, imatinib, and dasatinib, have been found to induce oxidative stress, which contributes to cancer cell apoptosis, TKI resistance, and cardiac toxicity [154,155]. In addition to these well-established drugs, increasing number of novel compounds, such as beta-phenylethyl isothiocyanate (PEITC) [156], leinamycin [157], and lanperisone (LP) [158], that can act as pro-oxidants or antioxidant inhibitors, have been developed and tested for their anticancer effects. It should be noted that the application of oxidative stress-inducing drugs for cancer treatment also faces many challenges. For example, the use of pro-oxidants may encounter limited tumor-selectivity, dose-limiting toxicity, acquired resistance, and difficulty in effective drug delivery. Besides chemotherapeutic agents, ionizing radiation can also trigger tumor cell apoptosis via ROS induction and the release of mitochondrial cytochrome c [159,160]. Comprehensive summarization of the progress and status of oxidative stress-inducing cancer therapy can be found in other reviews [7,8,9,123,124,125,143].

5. The Impact of Cancer Immunotherapies on Oxidative Stress in the TME

In recent years, immune-based therapies, exemplified by ICB therapy and CAR-T cell therapy, have increasingly become a viable treatment option for patients with cancer. Durable and curative outcomes have been observed in a fraction of patients with certain types of cancer after receiving immunotherapies. For example, ICB with aPD1 (nivolumab) and aCTLA4 (ipilimumab) antibodies led to durable responses in ~20% of patients with metastatic melanoma [161], and complete responses were achieved in nearly 80% patients with advanced B-cell acute lymphoblastic leukemia (B-ALL) who received CD19-targeting CAR-T cell therapy [162]. ICB treatment leads to better T cell activation and function by blocking the inhibitory signals transmitted by co-inhibitory molecules such as PD1 and CTLA4. CAR-T cells mediate antitumor effects by specifically recognizing the target molecules on cancer cells and subsequently destroying the cells through cytotoxic granules such as perforin and granzymes, along with inflammatory cytokines such as IFN-γ and TNF-α. Despite their distinct mechanisms of action, emerging evidence from preclinical studies indicates that ICB therapy and CAR-T therapy can both modulate oxidative stress in the TME, and that alteration of tumor oxidative stress contributes to the efficacy of immunotherapy [163,164].
Wang et al. reported that tumor stromal cells, such as fibroblasts, can facilitate tumor chemoresistance by modulating ROS in the TME [165]. GSH and cysteine released by fibroblasts can be used by ovarian cancer cells to diminish nuclear accumulation of platinum, resulting in resistance to platinum-based chemotherapy. The authors showed that tumor-infiltrating CD8 T cells can abolish fibroblast-mediated chemoresistance. Mechanistically, CD8+ T cell-derived IFN-γ reduces extracellular source of GSH and cysteine by upregulating gamma-glutamyl transferases, which break down extracellular GSH, and meanwhile repressing the expression of the cystine and glutamate antiporter system xc- (xCT), which imports extracellular cystine to facilitate GSH synthesis [165]. The same research team further demonstrated that the combination of cyst(e)inase, an engineered enzyme which degrades both cystine and cysteine, and ICB therapy synergistically impairs cystine uptake via xCT in tumor cells, resulting in GSH deficiency, ROS accumulation, lipid peroxidation, and ferroptosis of cancer cells in preclinical models [164]. These studies imply that ROS-driven tumor ferroptosis is an exploitable anti-tumor mechanism, and targeting this pathway in the context of immunotherapy represents a promising therapeutic strategy.
Using mouse tumor models, we reported that adoptive T cell therapy (ACT) can profoundly alter tumor metabolism, resulting in GSH depletion and consequential ROS accumulation in tumor cells [163]. We found that T cell-derived TNFα can synergize with chemotherapy to intensify oxidative stress in cancer cells in a NOX-dependent manner. Reduction of oxidative stress, by preventing TNFα-signaling in tumor cells or scavenging ROS with NAC, antagonizes the therapeutic effects of ACT. Depletion of GSH is one of the mechanisms by which many anticancer drugs elicit ROS-induced tumor cell death [153,156,166]. Our study provides evidence that GSH depletion can be achieved by T cell-based immunotherapy. Unlike most small compound inhibitors, T cell-mediated GSH depletion does not impair the function of the rate-limiting GSH-synthesizing enzyme GCL. However, tumor-specific CD4+ effector T cells can simultaneously disrupt multiple metabolic pathways to cause deficits in several intermediate metabolites involved in GSH synthesis, including homocysteine, cystathionine, and glycine [163]. The overall collapse of the redox-related pathways driven by T cells may block potential compensatory mechanisms, thereby overcoming tumor resistance. These findings imply that the ability of T cells to tilt tumor redox balance toward oxidative destruction is integral to the efficacy of ACT.
The impact of therapeutic antibodies on tumor oxidative stress is examined in a study in which mice bearing implanted lung adenocarcinoma tumors were treated with a cocktail of immunomodulators (anti-PD1, anti-CTLA-4, anti-CD137, and anti-CD19 monoclonal antibodies). Treatment-induced reduction in tumor burden is associated with decreased tumor proliferation but increased oxidative stress, apoptosis, autophagy, and T cell infiltration. The data suggest that treatment with therapeutic antibodies may induce oxidative stress that drives cell cycle arrest and tumor cell death [167].
Taken together, accumulating studies start to reveal the cellular and molecular mechanisms by which the dynamic interplay among cancer cells, immune cells, and stromal cells in the TME alters the redox status of each cell population. It is increasingly clear that antitumor T cells possess the ability to induce oxidative stress in tumor cells, and meanwhile they are susceptible to suppression imposed by ROS derived from the surrounding immunosuppressive cells such as Tregs and MDSCs (Figure 1). Therefore, therapeutic strategies should be directed to amplify T cell-induced oxidative stress in cancer cells while relieving effector T cells from the elevated oxidative stress in the TME.

6. Emerging ROS-Modulating Agents with the Potential to Enhance the Efficacy of Cancer Immunotherapy

So far, extensive efforts have been focused on developing small molecule compounds or biologics to target certain redox pathways in cancer cells. Although promising results have been observed in some cases, the use of these pro-oxidants for cancer treatment often encounters challenges related to tumor selectivity, toxicity to normal tissues, and development of chemoresistance [7,8,123,124]. We postulate that these issues can be addressed by combining pro-oxidants with immunotherapy in a synergistic manner to achieve durable antitumor effects while minimizing unwanted side-effects. Indeed, the recent findings that increased tumor oxidative stress correlates with the efficacy of ICB and ACT in preclinical models imply that pro-oxidants can be employed to intensify tumor oxidative stress so as to sensitize tumor cells to T cell-based immunotherapy [163,164]. Given that T cells are also sensitive to oxidative stress, it is unlikely any type of pro-oxidant is suitable for combination with immunotherapy. We consider that an immunotherapy-compatible pro-oxidant should meet the following criteria: (1) No obvious toxicity to tumor-reactive T cells at the doses needed to induce oxidative stress in tumor cells; (2) Easy administration to tumor-bearing hosts; (3) Good safety profiles that allow rapid translational studies. Many compounds may satisfy these criteria; here, we only highlight several representative agents which have shown the promise of being able to enhance the efficacy of T cell-based immunotherapy.

6.1. High Dose Ascorbate (Vitamin C)

Ascorbate, aka ascorbic acid (AA or vitamin C), at physiological dose functions as an antioxidant. However, mounting evidence indicates that ascorbate used at pharmacological doses (millimolar range) can act as a pro-oxidant that induces extracellular hydrogen peroxide (H2O2), which can freely diffuse into cells to cause damages in DNA, lipids, and proteins [168,169,170]. Since the uptake of oral ascorbate in humans is tightly controlled by the gut and kidney filtration, pharmacologic concentrations of ascorbate cannot be obtained by oral administration. Intravenous administration of ascorbate bypasses the tight control of the gut and renal excretion, resulting in high levels of ascorbate in plasma. Ascorbate undergoes autoxidation to generate a high flux of extracellular H2O2. It has been shown that high-dose ascorbate can be tumoricidal in vitro and can inhibit tumor growth in a variety of preclinical models [171,172,173,174,175,176]. Although early randomized clinical trials concluded that oral administration of high-dose ascorbate to patients with advanced cancers does not afford any therapeutic benefits [177,178], this conclusion was later challenged based on the discovery that parenteral (i.v. or i.p.) injection, not oral administration, of ascorbate is required to achieve plasma concentration high enough (20 mM) to damage cancer cells [175,176,179,180,181]. Currently, there are more than 30 completed, recruiting and active clinical trials investigating the usefulness of high-dose ascorbate in cancer treatment, either as monotherapy or in combination with other chemotherapeutic agents (https://clinicaltrials.gov, accessed on 11 February 2021). Completed clinical trials demonstrated that high-dose intravenous ascorbate is well tolerated in cancer patients with normal renal function, and in some cases can alleviate the severity of side-effects caused by chemotherapy [171,175].
The mechanisms underlying the preferential toxicity of ascorbate toward cancer cells over normal cells are not fully understood. One apparent explanation is that the antioxidant systems in cancer cells, which are already overstretched due to increased basal level of ROS, are overwhelmed by the influx of ascorbate-induced H2O2, while normal cells still have the capacity to mitigate the threat of the H2O2 burst. Additional mechanisms for ascorbate’s tumoritropic toxicity have also been described [176]. It has been shown that high dose ascorbate can selectively kill human colorectal cancers (CRCs) carrying KRAS or BRAF mutations [173]. This effect is due to increased uptake of the oxidized form of ascorbate, dehydroascorbate (DHA), via the glucose transporter GLUT1. Intracellular DHA is reduced to ascorbate at the expense of intracellular GSH, causing increased oxidative stress in cancer cells. Accumulated ROS inactivate glyceraldehyde 3-phosphate dehydrogenase (GAPDH), an enzyme critically involved in regulating glycolysis, causing an energetic crisis and cell death in highly glycolytic KRAS or BRAF mutant CRC cells but not normal cells. However, another study reported that the selective tumor toxicity by ascorbate is not dependent on DHA uptake. Instead, ascorbate’s toxicity on non-small-cell lung cancer (NSCLC) and glioblastoma (GBM) cells is dependent on the intracellular reactions of H2O2 and redox-active labile iron [182]. Cancer cells have increased basal levels of O2⋅− and H2O2 [183,184] and increased labile iron [185,186,187]. The increased labile iron in cancer cells leads to increased oxidation of ascorbate to generate more H2O2 capable of further exacerbating the differences in labile iron in cancer versus normal cells. This self-amplifying labile iron-H2O2 cycle results in increased Fenton chemistry to generate hydroxyl radicals (•OH) that cause irreversible oxidative damages to cancer cells [182].
It should be noted that ascorbate may also mediate tumoricidal effects through ROS-independent mechanisms. It has been shown that ascorbate is a cofactor for the Ten-Eleven Translocation (TET) enzymes, which mediate DNA demethylation by converting 5-methylcytosine (5 mC) to 5-hydroxymethylcytosine (5 hmC) and other oxidized methylcytosines. Shenoy et al. reported that ascorbate treatment of diffuse large B-cell (DLBCL) and peripheral T-cell (PTCL) lymphomas increases TET activities, which lead to increased demethylation in cancer cells [188]. This epigenetic effect of ascorbate results in reactivation of SMAD1, a tumor suppressor gene, which sensitizes cancer cells to chemotherapy. Along the same line, a subsequent study showed that high dose ascorbate reduces methylation and restores genome-wide 5 hmC levels in clear cell renal cell carcinoma (ccRCC) cell lines via TET activation [189]. Pharmacologic dose ascorbate treatment leads to increased intratumoral 5 hmC and reduced growth of ccRCC in vitro and in vivo. Furthermore, ascorbate treatment has been shown to mimic TET2 activities and suppress human leukemic colony formation and leukemia progression of primary human leukemia PDXs [190]. Of note, the TET-inducing effect of ascorbate is independent of hydrogen peroxide. These data indicate that in addition to its pro-oxidative effect, ascorbate-mediated epigenetic regulation may also contribute to tumor suppression.
It is important to note that two recent reports demonstrated that high dose ascorbate synergizes with anti-PD1 ICB therapy in mouse tumor models [191,192]. The two studies showed that administration of high dose ascorbate augments the efficacy of anti-PD1 therapy against several types of cancer in immunocompetent mice. The beneficial effects of ascorbate are associated with enhanced tumor infiltration by CD8+ T cells, granzyme B production by CD8+ T and NK cells, and IL-12 production by antigen-presenting cells. Interestingly, the immunopotentiating effects of high dose ascorbate appear to be independent of its pro-oxidant property. Instead, increased levels of 5 hmC are observed in both cancer and CD8+ T cells, suggesting the involvement of ascorbate-induced TET activities. Importantly, these studies demonstrated that high dose ascorbate does not harm effector T cells, but rather enhances T cell functionality through TET-mediated epigenetic modifications. It is worth noting that in these studies the assumption that the immunopotentiating effects of ascorbate are ROS-independent is based on the observation that provision of antioxidant NAC does not diminish the beneficial effect of ascorbate [192]. Future studies should employ more mechanism-based genetic and/or epigenetic approaches to further determine whether the pro-oxidant effects of ascorbate act in parallel to its epigenetic-modification effects. The possible mechanisms of action of ascorbate and potential combination with immunotherapy are illustrated in Figure 2A.

6.2. Non-Steroid Anti-Inflammatory Drugs (NSAIDs)

NSAIDs commonly used for relief of pain, fever, and inflammation act by inhibiting cyclooxygenases (COXs), COX1 and COX2, to suppress prostaglandin synthesis [193]. Some FDA-approved NSAIDs have also been shown to inhibit tumorigenesis in multiple rodent models, and epidemiological studies reported reduced incidence of various cancers in humans, especially colorectal cancer [194,195,196]. Their mechanisms in anticancer activities are not fully understood, but both COX-dependent and -independent pathways play a role. COX2-derived prostaglandin E2 (PGE2) can bind to its receptors on cancer cells and promote tumor cell proliferation, migration, angiogenesis, and chemoresistance [197,198]. Although the anticancer effects of many NSAIDs are attributable to inhibition of the COX2/PGE2 axis, additional mechanisms of action of NSAIDs have been characterized. NSAIDs can inhibit β-catenin transcriptional activity in cancer cells, resulting in reduced tumor growth [199,200]. In addition, suppression of tumor cell growth by some NSAIDs correlates with inhibition of cGMP degrading phosphodiesterase (PDE) activity [201,202]. Furthermore, it has been well-established that NSAIDs can induce oxidative and ER stresses that cause cancer cell apoptosis [203,204,205,206,207,208,209]. A number of commonly used NSAIDs, including sulindac, celecoxib, indomethacin, etc., have been found to induce ROS in various cancer cell lines and can inhibit tumor cell growth independent of COX2 inhibition [210,211,212,213,214]. It has been shown that NSAID treatment destabilizes the redox balance and antioxidant defense mechanisms of the thioredoxin and glutathione systems, resulting in GSH depletion and increased ROS production in tumor cells [215,216,217]. These events lead to a decline in mitochondrial membrane potential, release of cytochrome c, degradation of pro-survival molecules BCL-XL and BCL-2, and activation of the caspase cascade that leads to cancer cell apoptosis [209].
Published studies indicate that the antineoplastic activities of NSAIDs can also provoke antitumor immune responses. Inhibition of PGE2, a potent immunosuppressive factor enriched in the TME, leads to improved antitumor immunity [198,218]. Inhibition of phosphodiesterase 5 (PDE5) activity can abrogate MDSC-mediated immune suppression [219,220]. Suppression of β-catenin can turn immunologic “cold” tumors into “hot” tumors by activating dendritic cells and recruiting T cells into tumors [221,222]. Moreover, NSAID-induced ER stress was reported to correlate with ICD and tumor immunosurveillance [223]. Moreover, tumor antigen-specific T cells stimulated in the presence of a NSAID indomethacin have been shown to acquire stem-like property and can mediate strong antitumor effects upon adoptive transfer in mouse models [224]. These data collectively suggest that certain NSAIDs are compatible with T cell-based immunotherapy. This is supported by the results from an elegant study in which administration of NSAIDs, including celecoxib and aspirin, augments the efficacy of anti-PD1 therapy in multiple mouse tumor models [225]. The beneficial effects of NSAIDs are largely attributed to inhibition of COX2 and PGE2 in this study. It remains to be determined whether the pro-oxidant effect of NSAIDs can enhance immunotherapy for cancers that do not rely heavily on the COX2-PGE2 axis for survival and progression. The development of non-COX inhibitory NSAIDs can also avoid GI and cardiovascular toxicities associated with long-term NSAID administration, reducing potential complications when used in combination with immunotherapy. The possible mechanisms of action of NSAIDs and potential combination with immunotherapy are illustrated in Figure 2B.

6.3. xCT Inhibitors and Cyst(e)inase

Cancer cells have a higher demand for GSH to counter balance increased levels of ROS to maintain redox homeostasis for survival and proliferation. Cysteine is the rate-limiting amino acid necessary for GSH biosynthesis. However, under the condition of increased oxidative stress, the production of cysteine in cancer cells is often insufficient to meet the requirements of GSH synthesis, causing cancer cells to rely on uptake of extracellular source of cysteine in its disulfide form cystine via the cystine-glutamate antiporter xCT [226]. Various approaches have been developed to target this pathway as an anticancer strategy [226,227,228,229,230,231]. xCT inhibitors, such as sulfasalazine [232,233,234] and erastin [235,236,237,238,239], can reduce GSH and increase ROS in cancer cells, propagating iron-dependent lipid peroxidation that leads to tumor ferroptosis. Since inhibition of xCT alone may force cancer cells to import cysteine via other amino acid transporters such as alanine/serine/cysteine/threonine transporters (ASCT1 and ASCT2), xCT inhibitors are often used in combination with other chemotherapeutic agents or radiotherapy to overcome drug resistance. An engineered protein cyst(e)inase has been developed to enzymatically degrade both cysteine and cysteine [166]. Administration of cyst(e)inase mediates sustained depletion of the extracellular cysteine and cystine pool in mice and non-human primates. Cyst(e)inase selectively causes cell cycle arrest and death in cancer cells due to depletion of intracellular GSH and ensuing elevated ROS. Cyst(e)inase suppresses the growth of multiple types of cancer in mice, including prostate, breast, leukemia, and pancreatic cancer, yet no apparent toxicities are observed in mice after prolonged treatment [166,240]. These data implicate cyst(e)inase as a safe and effective therapeutic modality for inactivating antioxidant cellular responses in a wide range of malignancies.
Accumulating evidence suggests that xCT inhibitors and cyst(e)inase can be used in combination with immunotherapy to achieve a synergistic antitumor effect. Using genetic approaches, Arensman et al. demonstrated that while xCT is essential for tumor cell growth, it is dispensable for T cell proliferation in vivo and for the generation of primary and memory immune responses to tumors [241]. This study also showed that anti-CTLA4 therapy is more effective in treating xCT-deficient tumors in mouse models, providing proof of concept that administration of xCT inhibitors may augment the antitumor efficacy of ICB. Recent studies from Weiping Zou’s group demonstrated that antitumor CD8+ T cells can drive tumor ferroptosis through IFN-γ-mediated inhibition of xCT, and the combination of cyst(e)inase and ICB therapy synergistically suppresses tumor growth in preclinical models [164,242]. The possible mechanisms of action of cyst(e)inase/xCT inhibitors and potential combination with immunotherapy are illustrated in Figure 2C.

6.4. ROS-Responsive Prodrugs

The feature that cancer cells have elevated levels of ROS compared to normal cells has been employed to develop a class of prodrugs that only become cytotoxic in the presence of ROS. Prodrugs that are specifically activated by ROS in tumor cells have the potential to improve tumor selectivity and reduce toxicity to normal tissues. Therapeutic molecules, including chemotherapeutics and anti-PDL1 antibody, can be delivered to and released within tumor cells or TME by ROS-responsive prodrugs or nanoparticles [243,244,245,246,247,248], resulting in significant inhibition of tumor cell growth both in vitro and in vivo. In recent years, several research groups have developed novel prodrugs that act as DNA cross-linking or alkylating agents upon activation by ROS. Leinamycin (LNM) is a potent antitumor antibiotic produced by Streptomyces atroolivaceus S-140. LNM E1 as a prodrug can be oxidatively activated by cellular ROS to generate an intermediate with DNA alkylating activity, exhibiting potent cytotoxicity to prostate cancer cell lines with increased levels of ROS [157]. Peng’s group has developed a series of aromatic nitrogen mustards that are released from prodrugs upon a specific reaction between boronates and H2O2 in cancer cells [249,250]. These agents show potent DNA cross-linking abilities when coupled with H2O2, whereas little DNA cross-linking is detected in the absence of H2O2. These prodrugs can selectively kill leukemia and breast cancer cells, which have inherently high levels of ROS [251]. Interestingly, these prodrugs are not toxic to normal lymphocytes at the doses needed to kill cancer cells [252], suggesting their potential usage in combination with T cell-based immunotherapy. Prodrugs activated via ferrocene-mediated oxidation have also been developed to improve the selectivity of anticancer drugs [253,254,255,256]. These prodrugs show selective toxicity to a variety of cancer cell lines in vitro and in vivo, but remain weakly toxic to nonmalignant cells. Importantly, a recent study demonstrated that one such ROS-responsive prodrug, N-(3-(piperidin-1-ylmethyl)benzyl)-4-(ferrocenylcarbamatmethyl)phenyl boronic acid pinacol ester (PipFcB), can sensitize human lymphoma cell lines and primary chronic lymphocytic leukemia cells to CD19CAR-T cells [257]. It is noteworthy that exposure of CAR-T cells to PipFcB does not influence T cell exhaustion, viability, or T cell subpopulations. The T cell-friendly feature of the prodrugs, together with the findings by our group and others that antitumor T cells intensify ROS accumulation in tumor cells, suggest potential synergistic anticancer effects when combining ROS-responsive prodrugs with T cell-based immunotherapies. One possible scenario is that antitumor T cells arising after ICB or CAR-T therapy cause ROS accumulation in cancer cells, which activates prodrugs to release alkylating intermediates, which in turn further amplifies ROS in tumor (Figure 2D). This mutually reinforcing and self-amplifying ROS-inducing loop may lead to specific and complete tumor rejection with minimal toxicities to normal tissues.

7. Conclusions and Perspectives

An increasing number of studies indicate that ROS accumulated in tumor cells after immunotherapies are not merely metabolic byproducts but actively contribute to the treatment efficacy. Since these studies were mostly conducted in preclinical models, it remains to be determined in clinical samples whether various forms of cancer immunotherapy, including cancer vaccines, ICB and CAR-T therapy, lead to increased oxidative stress in cancer cells, and whether the levels of tumor oxidative stress correlate with the treatment outcomes. Rational combination of ROS-modulating agents and cancer immunotherapy is emerging as a promising treatment strategy (Figure 2). Given the availability of a multitude of pro-oxidants developed in recent decades, it is possible to identify and utilize a number of novel T cell-compatible agents, such as cyst(e)inase and ROS-responsive prodrugs, to enhance the efficacy of ICB or CAR-T therapy. There are also existing drugs (some are FDA-approved), such as ascorbate and NSAIDs, that can be repurposed as pro-oxidants and used in combination with immunotherapy. Future studies should address the sequence and timing of pro-oxidant administration in relation to immunotherapy, and determine the efficacy and toxicity of the combination therapy in preclinical models with the goal of translating this strategy for the betterment of cancer treatment.

Author Contributions

N.S.A. wrote/edited the manuscript, C.B., T.K., and Z.-C.D. reviewed/edited the manuscript, G.Z. wrote/edited the manuscript and prepared the figures. All authors have read and agreed to the published version of the manuscript.

Funding

G.Z. is supported in part by funds from the National Institutes of Health (1R01CA215523 and 1R01CA238514).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schumacker, P.T. Reactive oxygen species in cancer cells: Live by the sword, die by the sword. Cancer Cell 2006, 10, 175–176. [Google Scholar] [CrossRef] [Green Version]
  2. Liou, G.Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef] [Green Version]
  3. Chio, I.I.C.; Tuveson, D.A. ROS in Cancer: The Burning Question. Trends Mol. Med. 2017, 23, 411–429. [Google Scholar] [CrossRef] [Green Version]
  4. Reczek, C.R.; Chandel, N.S. The Two Faces of Reactive Oxygen Species in Cancer. Annu. Rev. Cancer Biol. 2017, 1, 79–98. [Google Scholar] [CrossRef]
  5. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [Green Version]
  6. Bae, Y.S.; Oh, H.; Rhee, S.G.; Yoo, Y.D. Regulation of reactive oxygen species generation in cell signaling. Mol. Cells 2011, 32, 491–509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Trachootham, D.; Alexandre, J.; Huang, P. Targeting cancer cells by ROS-mediated mechanisms: A radical therapeutic approach? Nat. Rev. Drug Discov. 2009, 8, 579–591. [Google Scholar] [CrossRef] [PubMed]
  8. 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]
  9. Nogueira, V.; Hay, N. Molecular pathways: Reactive oxygen species homeostasis in cancer cells and implications for cancer therapy. Clin. Cancer Res. 2013, 19, 4309–4314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Behrend, L.; Henderson, G.; Zwacka, R.M. Reactive oxygen species in oncogenic transformation. Biochem. Soc. Trans. 2003, 31, 1441–1444. [Google Scholar] [CrossRef]
  11. Hu, Y.; Rosen, D.G.; Zhou, Y.; Feng, L.; Yang, G.; Liu, J.; Huang, P. Mitochondrial manganese-superoxide dismutase expression in ovarian cancer: Role in cell proliferation and response to oxidative stress. J. Biol. Chem. 2005, 280, 39485–39492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Pervaiz, S.; Clement, M.V. Tumor intracellular redox status and drug resistance--serendipity or a causal relationship? Curr. Pharm. Des. 2004, 10, 1969–1977. [Google Scholar] [CrossRef]
  13. Clerkin, J.S.; Naughton, R.; Quiney, C.; Cotter, T.G. Mechanisms of ROS modulated cell survival during carcinogenesis. Cancer Lett. 2008, 266, 30–36. [Google Scholar] [CrossRef]
  14. Trachootham, D.; Lu, W.; Ogasawara, M.A.; Nilsa, R.D.; Huang, P. Redox regulation of cell survival. Antioxid. Redox Signal. 2008, 10, 1343–1374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Meier, B.; Radeke, H.H.; Selle, S.; Younes, M.; Sies, H.; Resch, K.; Habermehl, G.G. Human fibroblasts release reactive oxygen species in response to interleukin-1 or tumour necrosis factor-alpha. Biochem. J. 1989, 263, 539–545. [Google Scholar] [CrossRef]
  16. Ohba, M.; Shibanuma, M.; Kuroki, T.; Nose, K. Production of hydrogen peroxide by transforming growth factor-beta 1 and its involvement in induction of egr-1 in mouse osteoblastic cells. J. Cell Biol. 1994, 126, 1079–1088. [Google Scholar] [CrossRef]
  17. Lo, Y.Y.; Cruz, T.F. Involvement of reactive oxygen species in cytokine and growth factor induction of c-fos expression in chondrocytes. J. Biol. Chem. 1995, 270, 11727–11730. [Google Scholar] [CrossRef] [Green Version]
  18. Sundaresan, M.; Yu, Z.X.; Ferrans, V.J.; Irani, K.; Finkel, T. Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science 1995, 270, 296–299. [Google Scholar] [CrossRef] [Green Version]
  19. Roy, D.; Sarkar, S.; Felty, Q. Levels of IL-1 beta control stimulatory/inhibitory growth of cancer cells. Front. Biosci. 2006, 11, 889–898. [Google Scholar] [CrossRef] [Green Version]
  20. Irani, K.; Xia, Y.; Zweier, J.L.; Sollott, S.J.; Der, C.J.; Fearon, E.R.; Sundaresan, M.; Finkel, T.; GoldschmidtClermont, P.J. Mitogenic signaling mediated by oxidants in ras-transformed fibroblasts. Science 1997, 275, 1649–1652. [Google Scholar] [CrossRef]
  21. Mitsushita, J.; Lambeth, J.D.; Kamata, T. The superoxide-generating oxidase Nox1 is functionally required for Ras oncogene transformation. Cancer Res. 2004, 64, 3580–3585. [Google Scholar] [CrossRef] [Green Version]
  22. Weinberg, F.; Hamanaka, R.; Wheaton, W.W.; Weinberg, S.; Joseph, J.; Lopez, M.; Kalyanaraman, B.; Mutlu, G.M.; Budinger, G.R.; Chandel, N.S. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl. Acad. Sci. USA 2010, 107, 8788–8793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ogrunc, M.; Di Micco, R.; Liontos, M.; Bombardelli, L.; Mione, M.; Fumagalli, M.; Gorgoulis, V.G.; d’Adda di Fagagna, F. Oncogene-induced reactive oxygen species fuel hyperproliferation and DNA damage response activation. Cell Death Differ. 2014, 21, 998–1012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bae, Y.S.; Kang, S.W.; Seo, M.S.; Baines, I.C.; Tekle, E.; Chock, P.B.; Rhee, S.G. Epidermal growth factor (EGF)-induced generation of hydrogen peroxide—Role in EGF receptor-mediated tyrosine phosphorylation. J. Biol. Chem. 1997, 272, 217–221. [Google Scholar] [CrossRef] [Green Version]
  25. Sirokmány, G.; Pató, A.; Zana, M.; Donkó, Á.; Bíró, A.; Nagy, P.; Geiszt, M. Epidermal growth factor-induced hydrogen peroxide production is mediated by dual oxidase 1. Free Radic. Biol. Med. 2016, 97, 204–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Salmeen, A.; Andersen, J.N.; Myers, M.P.; Meng, T.C.; Hinks, J.A.; Tonks, N.K.; Barford, D. Redox regulation of protein tyrosine phosphatase 1B involves a sulphenyl-amide intermediate. Nature 2003, 423, 769–773. [Google Scholar] [CrossRef] [PubMed]
  27. Koundouros, N.; Poulogiannis, G. Phosphoinositide 3-Kinase/Akt Signaling and Redox Metabolism in Cancer. Front. Oncol. 2018, 8, 160. [Google Scholar] [CrossRef] [PubMed]
  28. Li, J.; Yen, C.; Liaw, D.; Podsypanina, K.; Bose, S.; Wang, S.I.; Puc, J.; Miliaresis, C.; Rodgers, L.; McCombie, R.; et al. PTEN, a Putative Protein Tyrosine Phosphatase Gene Mutated in Human Brain, Breast, and Prostate Cancer. Science 1997, 275, 1943. [Google Scholar] [CrossRef]
  29. Wu, H.; Goel, V.; Haluska, F.G. PTEN signaling pathways in melanoma. Oncogene 2003, 22, 3113–3122. [Google Scholar] [CrossRef] [Green Version]
  30. Verrastro, I.; Tveen-Jensen, K.; Woscholski, R.; Spickett, C.M.; Pitt, A.R. Reversible oxidation of phosphatase and tensin homolog (PTEN) alters its interactions with signaling and regulatory proteins. Free Radic. Biol. Med. 2016, 90, 24–34. [Google Scholar] [CrossRef] [Green Version]
  31. Seth, D.; Rudolph, J. Redox Regulation of MAP Kinase Phosphatase 3. Biochemistry 2006, 45, 8476–8487. [Google Scholar] [CrossRef] [PubMed]
  32. Son, Y.; Cheong, Y.-K.; Kim, N.-H.; Chung, H.-T.; Kang, D.; Pae, H.-O. Mitogen-Activated Protein Kinases and Reactive Oxygen Species: How Can ROS Activate MAPK Pathways? J. Signal. Transduct. 2011, 2011, 792639. [Google Scholar] [CrossRef]
  33. Zhang, J.; Johnston, G.; Stebler, B.; Keller, E.T. Hydrogen Peroxide Activates NFκB and the Interleukin-6 Promoter Through NFκB-Inducing Kinase. Antioxid. Redox Signal. 2001, 3, 493–504. [Google Scholar] [CrossRef]
  34. Morgan, M.J.; Liu, Z.-g. Crosstalk of reactive oxygen species and NF-κB signaling. Cell Res. 2011, 21, 103–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lingappan, K. NF-κB in Oxidative Stress. Curr. Opin. Toxicol. 2018, 7, 81–86. [Google Scholar] [CrossRef] [PubMed]
  36. Xia, Y.; Shen, S.; Verma, I.M. NF-κB, an active player in human cancers. Cancer Immunol. Res. 2014, 2, 823–830. [Google Scholar] [CrossRef] [Green Version]
  37. Li, F.; Zhang, J.; Arfuso, F.; Chinnathambi, A.; Zayed, M.E.; Alharbi, S.A.; Kumar, A.P.; Ahn, K.S.; Sethi, G. NF-kappaB in cancer therapy. Arch. Toxicol. 2015, 89, 711–731. [Google Scholar] [CrossRef]
  38. Panieri, E.; Santoro, M.M. ROS homeostasis and metabolism: A dangerous liason in cancer cells. Cell Death Dis. 2016, 7, e2253. [Google Scholar] [CrossRef]
  39. Ushio-Fukai, M.; Nakamura, Y. Reactive oxygen species and angiogenesis: NADPH oxidase as target for cancer therapy. Cancer Lett. 2008, 266, 37–52. [Google Scholar] [CrossRef] [Green Version]
  40. Wu, W.S. The signaling mechanism of ROS in tumor progression. Cancer Metastasis Rev. 2006, 25, 695–705. [Google Scholar] [CrossRef]
  41. Nishikawa, M. Reactive oxygen species in tumor metastasis. Cancer Lett. 2008, 266, 53–59. [Google Scholar] [CrossRef]
  42. Xia, C.; Meng, Q.; Liu, L.-Z.; Rojanasakul, Y.; Wang, X.-R.; Jiang, B.-H. Reactive Oxygen Species Regulate Angiogenesis and Tumor Growth through Vascular Endothelial Growth Factor. Cancer Res. 2007, 67, 10823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Komatsu, D.; Kato, M.; Nakayama, J.; Miyagawa, S.; Kamata, T. NADPH oxidase 1 plays a critical mediating role in oncogenic Ras-induced vascular endothelial growth factor expression. Oncogene 2008, 27, 4724–4732. [Google Scholar] [CrossRef] [Green Version]
  44. Chandel, N.S.; McClintock, D.S.; Feliciano, C.E.; Wood, T.M.; Melendez, J.A.; Rodriguez, A.M.; Schumacker, P.T. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: A mechanism of O2 sensing. J. Biol. Chem. 2000, 275, 25130–25138. [Google Scholar] [CrossRef] [Green Version]
  45. Guzy, R.D.; Hoyos, B.; Robin, E.; Chen, H.; Liu, L.; Mansfield, K.D.; Simon, M.C.; Hammerling, U.; Schumacker, P.T. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab. 2005, 1, 401–408. [Google Scholar] [CrossRef] [Green Version]
  46. Chandel, N.S.; Maltepe, E.; Goldwasser, E.; Mathieu, C.E.; Simon, M.C.; Schumacker, P.T. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci. USA 1998, 95, 11715–11720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Bell, E.L.; Klimova, T.A.; Eisenbart, J.; Schumacker, P.T.; Chandel, N.S. Mitochondrial reactive oxygen species trigger hypoxia-inducible factor-dependent extension of the replicative life span during hypoxia. Mol. Cell Biol. 2007, 27, 5737–5745. [Google Scholar] [CrossRef] [Green Version]
  48. Semenza, G.L. Hypoxia-inducible factors: Coupling glucose metabolism and redox regulation with induction of the breast cancer stem cell phenotype. EMBO J. 2017, 36, 252–259. [Google Scholar] [CrossRef]
  49. Al Tameemi, W.; Dale, T.P.; Al-Jumaily, R.M.K.; Forsyth, N.R. Hypoxia-Modified Cancer Cell Metabolism. Front. Cell Dev. Biol. 2019, 7, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Binker, M.G.; Binker-Cosen, A.A.; Richards, D.; Oliver, B.; Cosen-Binker, L.I. EGF promotes invasion by PANC-1 cells through Rac1/ROS-dependent secretion and activation of MMP-2. Biochem. Biophys. Res. Commun. 2009, 379, 445–450. [Google Scholar] [CrossRef]
  51. Kar, S.; Subbaram, S.; Carrico, P.M.; Melendez, J.A. Redox-control of matrix metalloproteinase-1: A critical link between free radicals, matrix remodeling and degenerative disease. Respir. Physiol. Neurobiol. 2010, 174, 299–306. [Google Scholar] [CrossRef] [Green Version]
  52. Meacham, C.E.; Morrison, S.J. Tumour heterogeneity and cancer cell plasticity. Nature 2013, 501, 328–337. [Google Scholar] [CrossRef] [Green Version]
  53. Diehn, M.; Cho, R.W.; Lobo, N.A.; Kalisky, T.; Dorie, M.J.; Kulp, A.N.; Qian, D.; Lam, J.S.; Ailles, L.E.; Wong, M.; et al. Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature 2009, 458, 780–783. [Google Scholar] [CrossRef]
  54. Ryoo, I.-g.; Lee, S.-h.; Kwak, M.-K. Redox Modulating NRF2: A Potential Mediator of Cancer Stem Cell Resistance. Oxidative Med. Cell. Longev. 2016, 2016, 2428153. [Google Scholar] [CrossRef] [Green Version]
  55. Rojo de la Vega, M.; Chapman, E.; Zhang, D.D. NRF2 and the Hallmarks of Cancer. Cancer Cell 2018, 34, 21–43. [Google Scholar] [CrossRef]
  56. Zimta, A.-A.; Cenariu, D.; Irimie, A.; Magdo, L.; Nabavi, S.M.; Atanasov, A.G.; Berindan-Neagoe, I. The Role of Nrf2 Activity in Cancer Development and Progression. Cancers 2019, 11, 1755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Reddy, N.M.; Kleeberger, S.R.; Yamamoto, M.; Kensler, T.W.; Scollick, C.; Biswal, S.; Reddy, S.P. Genetic dissection of the Nrf2-dependent redox signaling-regulated transcriptional programs of cell proliferation and cytoprotection. Physiol. Genom. 2007, 32, 74–81. [Google Scholar] [CrossRef] [Green Version]
  58. Nioi, P.; Hayes, J.D. Contribution of NAD(P)H:quinone oxidoreductase 1 to protection against carcinogenesis, and regulation of its gene by the Nrf2 basic-region leucine zipper and the arylhydrocarbon receptor basic helix-loop-helix transcription factors. Mutat. Res. 2004, 555, 149–171. [Google Scholar] [CrossRef]
  59. Hayes, J.D.; McMahon, M. NRF2 and KEAP1 mutations: Permanent activation of an adaptive response in cancer. Trends Biochem. Sci. 2009, 34, 176–188. [Google Scholar] [CrossRef]
  60. Kerins, M.J.; Ooi, A. A catalogue of somatic NRF2 gain-of-function mutations in cancer. Sci. Rep. 2018, 8, 12846. [Google Scholar] [CrossRef]
  61. 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]
  62. Hayes, J.D.; McMahon, M. The double-edged sword of Nrf2: Subversion of redox homeostasis during the evolution of cancer. Mol. Cell 2006, 21, 732–734. [Google Scholar] [CrossRef]
  63. Wu, S.; Lu, H.; Bai, Y. Nrf2 in cancers: A double-edged sword. Cancer Med. 2019, 8, 2252–2267. [Google Scholar] [CrossRef] [PubMed]
  64. Fiocchetti, M.; Fernandez, V.S.; Montalesi, E.; Marino, M. Neuroglobin: A Novel Player in the Oxidative Stress Response of Cancer Cells. Oxid Med. Cell Longev 2019, 2019, 6315034. [Google Scholar] [CrossRef] [Green Version]
  65. Solar Fernandez, V.; Cipolletti, M.; Ascenzi, P.; Marino, M.; Fiocchetti, M. Neuroglobin As Key Mediator in the 17beta-Estradiol-Induced Antioxidant Cell Response to Oxidative Stress. Antioxid Redox Signal. 2020, 32, 217–227. [Google Scholar] [CrossRef]
  66. Fiocchetti, M.; Solar Fernandez, V.; Segatto, M.; Leone, S.; Cercola, P.; Massari, A.; Cavaliere, F.; Marino, M. Extracellular Neuroglobin as a Stress-Induced Factor Activating Pre-Adaptation Mechanisms against Oxidative Stress and Chemotherapy-Induced Cell Death in Breast Cancer. Cancers 2020, 12, 2451. [Google Scholar] [CrossRef] [PubMed]
  67. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef]
  68. Valacchi, G.; Virgili, F.; Cervellati, C.; Pecorelli, A. OxInflammation: From Subclinical Condition to Pathological Biomarker. Front. Physiol. 2018, 9, 858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Nathan, C.; Cunningham-Bussel, A. Beyond oxidative stress: An immunologist’s guide to reactive oxygen species. Nat. Rev. Immunol. 2013, 13, 349–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Chen, X.; Song, M.; Zhang, B.; Zhang, Y. Reactive Oxygen Species Regulate T Cell Immune Response in the Tumor Microenvironment. Oxid Med. Cell Longev 2016, 2016, 1580967. [Google Scholar] [CrossRef] [Green Version]
  71. Franchina, D.G.; Dostert, C.; Brenner, D. Reactive Oxygen Species: Involvement in T Cell Signaling and Metabolism. Trends Immunol. 2018, 39, 489–502. [Google Scholar] [CrossRef] [PubMed]
  72. Kong, H.; Chandel, N.S. Regulation of redox balance in cancer and T cells. J. Biol. Chem. 2018, 293, 7499–7507. [Google Scholar] [CrossRef]
  73. Kotsafti, A.; Scarpa, M.; Castagliuolo, I.; Scarpa, M. Reactive Oxygen Species and Antitumor Immunity-From Surveillance to Evasion. Cancers 2020, 12, 1748. [Google Scholar] [CrossRef]
  74. Mullen, L.; Mengozzi, M.; Hanschmann, E.M.; Alberts, B.; Ghezzi, P. How the redox state regulates immunity. Free Radic. Biol. Med. 2020, 157, 3–14. [Google Scholar] [CrossRef] [PubMed]
  75. Muri, J.; Kopf, M. Redox regulation of immunometabolism. Nat. Rev. Immunol. 2020. [Google Scholar] [CrossRef] [PubMed]
  76. Van Loenhout, J.; Peeters, M.; Bogaerts, A.; Smits, E.; Deben, C. Oxidative Stress-Inducing Anticancer Therapies: Taking a Closer Look at Their Immunomodulating Effects. Antioxidants 2020, 9, 1188. [Google Scholar] [CrossRef]
  77. Reth, M. Hydrogen peroxide as second messenger in lymphocyte activation. Nat. Immunol. 2002, 3, 1129–1134. [Google Scholar] [CrossRef]
  78. Sena, L.A.; Li, S.; Jairaman, A.; Prakriya, M.; Ezponda, T.; Hildeman, D.A.; Wang, C.R.; Schumacker, P.T.; Licht, J.D.; Perlman, H.; et al. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity 2013, 38, 225–236. [Google Scholar] [CrossRef] [Green Version]
  79. Hildeman, D.A.; Mitchell, T.; Teague, T.K.; Henson, P.; Day, B.J.; Kappler, J.; Marrack, P.C. Reactive oxygen species regulate activation-induced T cell apoptosis. Immunity 1999, 10, 735–744. [Google Scholar] [CrossRef] [Green Version]
  80. Kono, K.; Salazar-Onfray, F.; Petersson, M.; Hansson, J.; Masucci, G.; Wasserman, K.; Nakazawa, T.; Anderson, P.; Kiessling, R. Hydrogen peroxide secreted by tumor-derived macrophages down-modulates signal-transducing zeta molecules and inhibits tumor-specific T cell-and natural killer cell-mediated cytotoxicity. Eur. J. Immunol. 1996, 26, 1308–1313. [Google Scholar] [CrossRef]
  81. Nakamura, K.; Matsunaga, K. Susceptibility of natural killer (NK) cells to reactive oxygen species (ROS) and their restoration by the mimics of superoxide dismutase (SOD). Cancer Biother. Radiopharm. 1998, 13, 275–290. [Google Scholar] [CrossRef]
  82. Malmberg, K.J.; Arulampalam, V.; Ichihara, F.; Petersson, M.; Seki, K.; Andersson, T.; Lenkei, R.; Masucci, G.; Pettersson, S.; Kiessling, R. Inhibition of activated/memory (CD45RO(+)) T cells by oxidative stress associated with block of NF-kappaB activation. J. Immunol. 2001, 167, 2595–2601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Mak, T.W.; Grusdat, M.; Duncan, G.S.; Dostert, C.; Nonnenmacher, Y.; Cox, M.; Binsfeld, C.; Hao, Z.; Brustle, A.; Itsumi, M.; et al. Glutathione Primes T Cell Metabolism for Inflammation. Immunity 2017, 46, 1089–1090. [Google Scholar] [CrossRef]
  84. Kesarwani, P.; Al-Khami, A.A.; Scurti, G.; Thyagarajan, K.; Kaur, N.; Husain, S.; Fang, Q.; Naga, O.S.; Simms, P.; Beeson, G.; et al. Promoting thiol expression increases the durability of antitumor T-cell functions. Cancer Res. 2014, 74, 6036–6047. [Google Scholar] [CrossRef] [Green Version]
  85. Scheffel, M.J.; Scurti, G.; Simms, P.; Garrett-Mayer, E.; Mehrotra, S.; Nishimura, M.I.; Voelkel-Johnson, C. Efficacy of Adoptive T-cell Therapy Is Improved by Treatment with the Antioxidant N-Acetyl Cysteine, Which Limits Activation-Induced T-cell Death. Cancer Res. 2016, 76, 6006–6016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Pilipow, K.; Scamardella, E.; Puccio, S.; Gautam, S.; De Paoli, F.; Mazza, E.M.; De Simone, G.; Polletti, S.; Buccilli, M.; Zanon, V.; et al. Antioxidant metabolism regulates CD8+ T memory stem cell formation and antitumor immunity. JCI Insight 2018, 3. [Google Scholar] [CrossRef] [Green Version]
  87. Scheffel, M.J.; Scurti, G.; Wyatt, M.M.; Garrett-Mayer, E.; Paulos, C.M.; Nishimura, M.I.; Voelkel-Johnson, C. N-acetyl cysteine protects anti-melanoma cytotoxic T cells from exhaustion induced by rapid expansion via the downmodulation of Foxo1 in an Akt-dependent manner. Cancer Immunol. Immunother. 2018, 67, 691–702. [Google Scholar] [CrossRef]
  88. Ligtenberg, M.A.; Mougiakakos, D.; Mukhopadhyay, M.; Witt, K.; Lladser, A.; Chmielewski, M.; Riet, T.; Abken, H.; Kiessling, R. Coexpressed Catalase Protects Chimeric Antigen Receptor-Redirected T Cells as well as Bystander Cells from Oxidative Stress-Induced Loss of Antitumor Activity. J. Immunol. 2016, 196, 759–766. [Google Scholar] [CrossRef] [Green Version]
  89. Yang, Y.; Neo, S.Y.; Chen, Z.; Cui, W.; Chen, Y.; Guo, M.; Wang, Y.; Xu, H.; Kurzay, A.; Alici, E.; et al. Thioredoxin activity confers resistance against oxidative stress in tumor-infiltrating NK cells. J. Clin. Invest. 2020, 130, 5508–5522. [Google Scholar] [CrossRef] [PubMed]
  90. Mantegazza, A.R.; Savina, A.; Vermeulen, M.; Perez, L.; Geffner, J.; Hermine, O.; Rosenzweig, S.D.; Faure, F.; Amigorena, S. NADPH oxidase controls phagosomal pH and antigen cross-presentation in human dendritic cells. Blood 2008, 112, 4712–4722. [Google Scholar] [CrossRef] [Green Version]
  91. Weiskopf, D.; Schwanninger, A.; Weinberger, B.; Almanzar, G.; Parson, W.; Buus, S.; Lindner, H.; Grubeck-Loebenstein, B. Oxidative stress can alter the antigenicity of immunodominant peptides. J. Leukoc. Biol. 2010, 87, 165–172. [Google Scholar] [CrossRef]
  92. Trujillo, J.A.; Croft, N.P.; Dudek, N.L.; Channappanavar, R.; Theodossis, A.; Webb, A.I.; Dunstone, M.A.; Illing, P.T.; Butler, N.S.; Fett, C.; et al. The cellular redox environment alters antigen presentation. J. Biol. Chem. 2014, 289, 27979–27991. [Google Scholar] [CrossRef] [Green Version]
  93. Malhotra, J.D.; Kaufman, R.J. Endoplasmic reticulum stress and oxidative stress: A vicious cycle or a double-edged sword? Antioxid Redox Signal. 2007, 9, 2277–2293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kepp, O.; Menger, L.; Vacchelli, E.; Locher, C.; Adjemian, S.; Yamazaki, T.; Martins, I.; Sukkurwala, A.Q.; Michaud, M.; Senovilla, L.; et al. Crosstalk between ER stress and immunogenic cell death. Cytokine Growth Factor Rev. 2013, 24, 311–318. [Google Scholar] [CrossRef] [PubMed]
  95. Cao, S.S.; Kaufman, R.J. Endoplasmic reticulum stress and oxidative stress in cell fate decision and human disease. Antioxid Redox Signal. 2014, 21, 396–413. [Google Scholar] [CrossRef] [PubMed]
  96. Panaretakis, T.; Kepp, O.; Brockmeier, U.; Tesniere, A.; Bjorklund, A.C.; Chapman, D.C.; Durchschlag, M.; Joza, N.; Pierron, G.; van Endert, P.; et al. Mechanisms of pre-apoptotic calreticulin exposure in immunogenic cell death. EMBO J. 2009, 28, 578–590. [Google Scholar] [CrossRef]
  97. Deng, H.; Zhou, Z.; Yang, W.; Lin, L.S.; Wang, S.; Niu, G.; Song, J.; Chen, X. Endoplasmic Reticulum Targeting to Amplify Immunogenic Cell Death for Cancer Immunotherapy. Nano Lett. 2020, 20, 1928–1933. [Google Scholar] [CrossRef]
  98. Kazama, H.; Ricci, J.E.; Herndon, J.M.; Hoppe, G.; Green, D.R.; Ferguson, T.A. Induction of immunological tolerance by apoptotic cells requires caspase-dependent oxidation of high-mobility group box-1 protein. Immunity 2008, 29, 21–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Deng, H.; Yang, W.; Zhou, Z.; Tian, R.; Lin, L.; Ma, Y.; Song, J.; Chen, X. Targeted scavenging of extracellular ROS relieves suppressive immunogenic cell death. Nat. Commun. 2020, 11, 4951. [Google Scholar] [CrossRef] [PubMed]
  100. Kraaij, M.D.; Savage, N.D.; van der Kooij, S.W.; Koekkoek, K.; Wang, J.; van den Berg, J.M.; Ottenhoff, T.H.; Kuijpers, T.W.; Holmdahl, R.; van Kooten, C.; et al. Induction of regulatory T cells by macrophages is dependent on production of reactive oxygen species. Proc. Natl. Acad. Sci. USA 2010, 107, 17686–17691. [Google Scholar] [CrossRef] [Green Version]
  101. Yan, Z.; Garg, S.K.; Banerjee, R. Regulatory T cells interfere with glutathione metabolism in dendritic cells and T cells. J. Biol. Chem. 2010, 285, 41525–41532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Efimova, O.; Szankasi, P.; Kelley, T.W. Ncf1 (p47phox) is essential for direct regulatory T cell mediated suppression of CD4+ effector T cells. PLoS ONE 2011, 6, e16013. [Google Scholar] [CrossRef] [Green Version]
  103. Wen, Z.; Shimojima, Y.; Shirai, T.; Li, Y.; Ju, J.; Yang, Z.; Tian, L.; Goronzy, J.J.; Weyand, C.M. NADPH oxidase deficiency underlies dysfunction of aged CD8+ Tregs. J. Clin. Invest. 2016, 126, 1953–1967. [Google Scholar] [CrossRef] [PubMed]
  104. Mougiakakos, D.; Johansson, C.C.; Kiessling, R. Naturally occurring regulatory T cells show reduced sensitivity toward oxidative stress-induced cell death. Blood 2009, 113, 3542–3545. [Google Scholar] [CrossRef] [Green Version]
  105. Mougiakakos, D.; Johansson, C.C.; Jitschin, R.; Bottcher, M.; Kiessling, R. Increased thioredoxin-1 production in human naturally occurring regulatory T cells confers enhanced tolerance to oxidative stress. Blood 2011, 117, 857–861. [Google Scholar] [CrossRef] [Green Version]
  106. Kurniawan, H.; Franchina, D.G.; Guerra, L.; Bonetti, L.; Baguet, L.S.; Grusdat, M.; Schlicker, L.; Hunewald, O.; Dostert, C.; Merz, M.P.; et al. Glutathione Restricts Serine Metabolism to Preserve Regulatory T Cell Function. Cell Metab 2020, 31, 920–936. [Google Scholar] [CrossRef]
  107. Maj, T.; Wang, W.; Crespo, J.; Zhang, H.; Wang, W.; Wei, S.; Zhao, L.; Vatan, L.; Shao, I.; Szeliga, W.; et al. Oxidative stress controls regulatory T cell apoptosis and suppressor activity and PD-L1-blockade resistance in tumor. Nat. Immunol. 2017, 18, 1332–1341. [Google Scholar] [CrossRef] [PubMed]
  108. Iida, N.; Dzutsev, A.; Stewart, C.A.; Smith, L.; Bouladoux, N.; Weingarten, R.A.; Molina, D.A.; Salcedo, R.; Back, T.; Cramer, S.; et al. Commensal bacteria control cancer response to therapy by modulating the tumor microenvironment. Science 2013, 342, 967–970. [Google Scholar] [CrossRef]
  109. Schmielau, J.; Finn, O.J. Activated granulocytes and granulocyte-derived hydrogen peroxide are the underlying mechanism of suppression of t-cell function in advanced cancer patients. Cancer Res. 2001, 61, 4756–4760. [Google Scholar] [PubMed]
  110. Aarts, C.E.M.; Hiemstra, I.H.; Beguin, E.P.; Hoogendijk, A.J.; Bouchmal, S.; van Houdt, M.; Tool, A.T.J.; Mul, E.; Jansen, M.H.; Janssen, H.; et al. Activated neutrophils exert myeloid-derived suppressor cell activity damaging T cells beyond repair. Blood Adv. 2019, 3, 3562–3574. [Google Scholar] [CrossRef] [Green Version]
  111. Otsuji, M.; Kimura, Y.; Aoe, T.; Okamoto, Y.; Saito, T. Oxidative stress by tumor-derived macrophages suppresses the expression of CD3 zeta chain of T-cell receptor complex and antigen-specific T-cell responses. Proc. Natl. Acad. Sci. USA 1996, 93, 13119–13124. [Google Scholar] [CrossRef] [Green Version]
  112. Nagaraj, S.; Gupta, K.; Pisarev, V.; Kinarsky, L.; Sherman, S.; Kang, L.; Herber, D.L.; Schneck, J.; Gabrilovich, D.I. Altered recognition of antigen is a mechanism of CD8(+) T cell tolerance in cancer. Nat. Med. 2007, 13, 828–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Liu, Y.; Wei, J.; Guo, G.; Zhou, J. Norepinephrine-induced myeloid-derived suppressor cells block T-cell responses via generation of reactive oxygen species. Immunopharmacol. Immunotoxicol. 2015, 37, 359–365. [Google Scholar] [CrossRef] [PubMed]
  114. Corzo, C.A.; Cotter, M.J.; Cheng, P.Y.; Cheng, F.D.; Kusmartsev, S.; Sotomayor, E.; Padhya, T.; McCaffrey, T.V.; McCaffrey, J.C.; Gabrilovich, D.I. Mechanism Regulating Reactive Oxygen Species in Tumor-Induced Myeloid-Derived Suppressor Cells. J. Immunol. 2009, 182, 5693–5701. [Google Scholar] [CrossRef]
  115. Kusmartsev, S.; Gabrilovich, D.I. Inhibition of myeloid cell differentiation in cancer: The role of reactive oxygen species. J. Leukoc Biol. 2003, 74, 186–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Beury, D.W.; Carter, K.A.; Nelson, C.; Sinha, P.; Hanson, E.; Nyandjo, M.; Fitzgerald, P.J.; Majeed, A.; Wali, N.; Ostrand-Rosenberg, S. Myeloid-Derived Suppressor Cell Survival and Function Are Regulated by the Transcription Factor Nrf2. J. Immunol. 2016, 196, 3470–3478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Kusmartsev, S.; Nefedova, Y.; Yoder, D.; Gabrilovich, D.I. Antigen-specific inhibition of CD8+ T cell response by immature myeloid cells in cancer is mediated by reactive oxygen species. J. Immunol. 2004, 172, 989–999. [Google Scholar] [CrossRef] [Green Version]
  118. Lu, T.; Gabrilovich, D.I. Molecular pathways: Tumor-infiltrating myeloid cells and reactive oxygen species in regulation of tumor microenvironment. Clin. Cancer Res. 2012, 18, 4877–4882. [Google Scholar] [CrossRef] [Green Version]
  119. Ohl, K.; Tenbrock, K. Reactive Oxygen Species as Regulators of MDSC-Mediated Immune Suppression. Front. Immunol. 2018, 9, 2499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Ostrand-Rosenberg, S.; Beury, D.W.; Parker, K.H.; Horn, L.A. Survival of the fittest: How myeloid-derived suppressor cells survive in the inhospitable tumor microenvironment. Cancer Immunol. Immunother. 2020, 69, 215–221. [Google Scholar] [CrossRef]
  121. Wang, J.; Yi, J. Cancer cell killing via ROS: To increase or decrease, that is the question. Cancer Biol. Ther. 2008, 7, 1875–1884. [Google Scholar] [CrossRef]
  122. Galadari, S.; Rahman, A.; Pallichankandy, S.; Thayyullathil, F. Reactive oxygen species and cancer paradox: To promote or to suppress? Free Radic. Biol. Med. 2017, 104, 144–164. [Google Scholar] [CrossRef]
  123. Raza, M.H.; Siraj, S.; Arshad, A.; Waheed, U.; Aldakheel, F.; Alduraywish, S.; Arshad, M. ROS-modulated therapeutic approaches in cancer treatment. J. Cancer Res. Clin. Oncol. 2017, 143, 1789–1809. [Google Scholar] [CrossRef] [PubMed]
  124. Kim, S.J.; Kim, H.S.; Seo, Y.R. Understanding of ROS-Inducing Strategy in Anticancer Therapy. Oxid Med. Cell Longev 2019, 2019, 5381692. [Google Scholar] [CrossRef] [PubMed]
  125. 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]
  126. Khan, N.; Afaq, F.; Mukhtar, H. Cancer chemoprevention through dietary antioxidants: Progress and promise. Antioxid Redox Signal. 2008, 10, 475–510. [Google Scholar] [CrossRef] [PubMed]
  127. Gao, P.; Zhang, H.; Dinavahi, R.; Li, F.; Xiang, Y.; Raman, V.; Bhujwalla, Z.M.; Felsher, D.W.; Cheng, L.; Pevsner, J.; et al. HIF-dependent antitumorigenic effect of antioxidants in vivo. Cancer Cell 2007, 12, 230–238. [Google Scholar] [CrossRef] [Green Version]
  128. Nishikawa, M.; Hyoudou, K.; Kobayashi, Y.; Umeyama, Y.; Takakura, Y.; Hashida, M. Inhibition of metastatic tumor growth by targeted delivery of antioxidant enzymes. J. Control. Release 2005, 109, 101–107. [Google Scholar] [CrossRef] [PubMed]
  129. Venkataraman, S.; Jiang, X.; Weydert, C.; Zhang, Y.; Zhang, H.J.; Goswami, P.C.; Ritchie, J.M.; Oberley, L.W.; Buettner, G.R. Manganese superoxide dismutase overexpression inhibits the growth of androgen-independent prostate cancer cells. Oncogene 2005, 24, 77–89. [Google Scholar] [CrossRef] [Green Version]
  130. Liu, J.; Du, J.; Zhang, Y.; Sun, W.; Smith, B.J.; Oberley, L.W.; Cullen, J.J. Suppression of the malignant phenotype in pancreatic cancer by overexpression of phospholipid hydroperoxide glutathione peroxidase. Hum. Gene Ther. 2006, 17, 105–116. [Google Scholar] [CrossRef]
  131. Nelson, S.K.; Bose, S.K.; Grunwald, G.K.; Myhill, P.; McCord, J.M. The induction of human superoxide dismutase and catalase in vivo: A fundamentally new approach to antioxidant therapy. Free Radic. Biol. Med. 2006, 40, 341–347. [Google Scholar] [CrossRef] [PubMed]
  132. Teoh-Fitzgerald, M.L.; Fitzgerald, M.P.; Zhong, W.; Askeland, R.W.; Domann, F.E. Epigenetic reprogramming governs EcSOD expression during human mammary epithelial cell differentiation, tumorigenesis and metastasis. Oncogene 2014, 33, 358–368. [Google Scholar] [CrossRef] [Green Version]
  133. The Effect of Vitamin E and Beta Carotene on the Incidence of Lung Cancer and Other Cancers in Male Smokers. N. Engl. J. Med. 1994, 330, 1029–1035. [CrossRef]
  134. Omenn, G.S.; Goodman, G.E.; Thornquist, M.D.; Balmes, J.; Cullen, M.R.; Glass, A.; Keogh, J.P.; Meyskens, F.L.; Valanis, B.; Williams, J.H.; et al. Effects of a combination of beta carotene and vitamin A on lung cancer and cardiovascular disease. N. Engl. J. Med. 1996, 334, 1150–1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. D’Andrea, G.M. Use of antioxidants during chemotherapy and radiotherapy should be avoided. CA Cancer J. Clin. 2005, 55, 319–321. [Google Scholar] [CrossRef]
  136. Seifried, H.E.; Anderson, D.E.; Fisher, E.I.; Milner, J.A. A review of the interaction among dietary antioxidants and reactive oxygen species. J. Nutr. Biochem. 2007, 18, 567–579. [Google Scholar] [CrossRef]
  137. Klein, E.A.; Thompson, I.M., Jr.; Tangen, C.M.; Crowley, J.J.; Lucia, M.S.; Goodman, P.J.; Minasian, L.M.; Ford, L.G.; Parnes, H.L.; Gaziano, J.M.; et al. Vitamin E and the risk of prostate cancer: The Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 2011, 306, 1549–1556. [Google Scholar] [CrossRef]
  138. Chandel, N.S.; Tuveson, D.A. The promise and perils of antioxidants for cancer patients. N. Engl. J. Med. 2014, 371, 177–178. [Google Scholar] [CrossRef]
  139. Le Gal, K.; Ibrahim, M.X.; Wiel, C.; Sayin, V.I.; Akula, M.K.; Karlsson, C.; Dalin, M.G.; Akyurek, L.M.; Lindahl, P.; Nilsson, J.; et al. Antioxidants can increase melanoma metastasis in mice. Sci. Transl. Med. 2015, 7, 308re308. [Google Scholar] [CrossRef] [PubMed]
  140. Peiris-Pages, M.; Martinez-Outschoorn, U.E.; Sotgia, F.; Lisanti, M.P. Metastasis and Oxidative Stress: Are Antioxidants a Metabolic Driver of Progression? Cell Metab. 2015, 22, 956–958. [Google Scholar] [CrossRef] [Green Version]
  141. Piskounova, E.; Agathocleous, M.; Murphy, M.M.; Hu, Z.; Huddlestun, S.E.; Zhao, Z.; Leitch, A.M.; Johnson, T.M.; DeBerardinis, R.J.; Morrison, S.J. Oxidative stress inhibits distant metastasis by human melanoma cells. Nature 2015, 527, 186–191. [Google Scholar] [CrossRef] [Green Version]
  142. Wiel, C.; Le Gal, K.; Ibrahim, M.X.; Jahangir, C.A.; Kashif, M.; Yao, H.; Ziegler, D.V.; Xu, X.; Ghosh, T.; Mondal, T.; et al. BACH1 Stabilization by Antioxidants Stimulates Lung Cancer Metastasis. Cell 2019, 178, 330–345. [Google Scholar] [CrossRef]
  143. Firczuk, M.; Bajor, M.; Graczyk-Jarzynka, A.; Fidyt, K.; Goral, A.; Zagozdzon, R. Harnessing altered oxidative metabolism in cancer by augmented prooxidant therapy. Cancer Lett. 2020, 471, 1–11. [Google Scholar] [CrossRef]
  144. Berndtsson, M.; Hagg, M.; Panaretakis, T.; Havelka, A.M.; Shoshan, M.C.; Linder, S. Acute apoptosis by cisplatin requires induction of reactive oxygen species but is not associated with damage to nuclear DNA. Int. J. Cancer 2007, 120, 175–180. [Google Scholar] [CrossRef] [PubMed]
  145. Hwang, P.M.; Bunz, F.; Yu, J.; Rago, C.; Chan, T.A.; Murphy, M.P.; Kelso, G.F.; Smith, R.A.; Kinzler, K.W.; Vogelstein, B. Ferredoxin reductase affects p53-dependent, 5-fluorouracil-induced apoptosis in colorectal cancer cells. Nat. Med. 2001, 7, 1111–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Muller, I.; Niethammer, D.; Bruchelt, G. Anthracycline-derived chemotherapeutics in apoptosis and free radical cytotoxicity (Review). Int. J. Mol. Med. 1998, 1, 491–494. [Google Scholar] [CrossRef]
  147. De Boo, S.; Kopecka, J.; Brusa, D.; Gazzano, E.; Matera, L.; Ghigo, D.; Bosia, A.; Riganti, C. iNOS activity is necessary for the cytotoxic and immunogenic effects of doxorubicin in human colon cancer cells. Mol. Cancer 2009, 8, 108. [Google Scholar] [CrossRef] [Green Version]
  148. Varbiro, G.; Veres, B.; Gallyas, F., Jr.; Sumegi, B. Direct effect of Taxol on free radical formation and mitochondrial permeability transition. Free Radic. Biol. Med. 2001, 31, 548–558. [Google Scholar] [CrossRef]
  149. Ramanathan, B.; Jan, K.Y.; Chen, C.H.; Hour, T.C.; Yu, H.J.; Pu, Y.S. Resistance to paclitaxel is proportional to cellular total antioxidant capacity. Cancer Res. 2005, 65, 8455–8460. [Google Scholar] [CrossRef] [Green Version]
  150. Chiu, W.H.; Luo, S.J.; Chen, C.L.; Cheng, J.H.; Hsieh, C.Y.; Wang, C.Y.; Huang, W.C.; Su, W.C.; Lin, C.F. Vinca alkaloids cause aberrant ROS-mediated JNK activation, Mcl-1 downregulation, DNA damage, mitochondrial dysfunction, and apoptosis in lung adenocarcinoma cells. Biochem. Pharmacol. 2012, 83, 1159–1171. [Google Scholar] [CrossRef]
  151. Chou, W.C.; Jie, C.; Kenedy, A.A.; Jones, R.J.; Trush, M.A.; Dang, C.V. Role of NADPH oxidase in arsenic-induced reactive oxygen species formation and cytotoxicity in myeloid leukemia cells. Proc. Natl. Acad. Sci. USA 2004, 101, 4578–4583. [Google Scholar] [CrossRef] [Green Version]
  152. Griffith, O.W.; Meister, A. Potent and specific inhibition of glutathione synthesis by buthionine sulfoximine (S-n-butyl homocysteine sulfoximine). J. Biol. Chem. 1979, 254, 7558–7560. [Google Scholar] [CrossRef]
  153. Harris, I.S.; Treloar, A.E.; Inoue, S.; Sasaki, M.; Gorrini, C.; Lee, K.C.; Yung, K.Y.; Brenner, D.; Knobbe-Thomsen, C.B.; Cox, M.A.; et al. Glutathione and Thioredoxin Antioxidant Pathways Synergize to Drive Cancer Initiation and Progression (vol 27, pg 211, 2015). Cancer Cell 2015, 27, 314. [Google Scholar] [CrossRef]
  154. Teppo, H.R.; Soini, Y.; Karihtala, P. Reactive Oxygen Species-Mediated Mechanisms of Action of Targeted Cancer Therapy. Oxid Med. Cell Longev. 2017, 2017, 1485283. [Google Scholar] [CrossRef] [PubMed]
  155. Rodriguez-Hernandez, M.A.; de la Cruz-Ojeda, P.; Lopez-Grueso, M.J.; Navarro-Villaran, E.; Requejo-Aguilar, R.; Castejon-Vega, B.; Negrete, M.; Gallego, P.; Vega-Ochoa, A.; Victor, V.M.; et al. Integrated molecular signaling involving mitochondrial dysfunction and alteration of cell metabolism induced by tyrosine kinase inhibitors in cancer. Redox. Biol. 2020, 36, 101510. [Google Scholar] [CrossRef] [PubMed]
  156. Trachootham, D.; Zhou, Y.; Zhang, H.; Demizu, Y.; Chen, Z.; Pelicano, H.; Chiao, P.J.; Achanta, G.; Arlinghaus, R.B.; Liu, J.S.; et al. Selective killing of oncogenically transformed cells through a ROS-mediated mechanism by beta-phenylethyl isothiocyanate. Cancer Cell 2006, 10, 241–252. [Google Scholar] [CrossRef] [Green Version]
  157. Huang, S.X.; Yun, B.S.; Ma, M.; Basu, H.S.; Church, D.R.; Ingenhorst, G.; Huang, Y.; Yang, D.; Lohman, J.R.; Tang, G.L.; et al. Leinamycin E1 acting as an anticancer prodrug activated by reactive oxygen species. Proc. Natl Acad Sci USA 2015, 112, 8278–8283. [Google Scholar] [CrossRef] [Green Version]
  158. Shaw, A.T.; Winslow, M.M.; Magendantz, M.; Ouyang, C.; Dowdle, J.; Subramanian, A.; Lewis, T.A.; Maglathin, R.L.; Tolliday, N.; Jacks, T. Selective killing of K-ras mutant cancer cells by small molecule inducers of oxidative stress. Proc. Natl. Acad. Sci. USA 2011, 108, 8773–8778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Ogura, A.; Oowada, S.; Kon, Y.; Hirayama, A.; Yasui, H.; Meike, S.; Kobayashi, S.; Kuwabara, M.; Inanami, O. Redox regulation in radiation-induced cytochrome c release from mitochondria of human lung carcinoma A549 cells. Cancer Lett. 2009, 277, 64–71. [Google Scholar] [CrossRef] [Green Version]
  160. Maier, P.; Hartmann, L.; Wenz, F.; Herskind, C. Cellular Pathways in Response to Ionizing Radiation and Their Targetability for Tumor Radiosensitization. Int. J. Mol. Sci. 2016, 17, 102. [Google Scholar] [CrossRef] [Green Version]
  161. Larkin, J.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Rutkowski, P.; Lao, C.D.; Cowey, C.L.; Schadendorf, D.; Wagstaff, J.; Dummer, R.; et al. Five-Year Survival with Combined Nivolumab and Ipilimumab in Advanced Melanoma. N. Engl. J. Med. 2019, 381, 1535–1546. [Google Scholar] [CrossRef] [Green Version]
  162. Brown, C.E.; Mackall, C.L. CAR T cell therapy: Inroads to response and resistance. Nat. Rev. Immunol 2019, 19, 73–74. [Google Scholar] [CrossRef] [PubMed]
  163. Habtetsion, T.; Ding, Z.C.; Pi, W.; Li, T.; Lu, C.; Chen, T.; Xi, C.; Spartz, H.; Liu, K.; Hao, Z.; et al. Alteration of Tumor Metabolism by CD4+ T Cells Leads to TNF-alpha-Dependent Intensification of Oxidative Stress and Tumor Cell Death. Cell Metab. 2018, 28, 228–242. [Google Scholar] [CrossRef] [Green Version]
  164. Wang, W.; Green, M.; Choi, J.E.; Gijon, M.; Kennedy, P.D.; Johnson, J.K.; Liao, P.; Lang, X.; Kryczek, I.; Sell, A.; et al. CD8(+) T cells regulate tumour ferroptosis during cancer immunotherapy. Nature 2019, 569, 270–274. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, W.; Kryczek, I.; Dostal, L.; Lin, H.; Tan, L.; Zhao, L.; Lu, F.; Wei, S.; Maj, T.; Peng, D.; et al. Effector T Cells Abrogate Stroma-Mediated Chemoresistance in Ovarian Cancer. Cell 2016, 165, 1092–1105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Cramer, S.L.; Saha, A.; Liu, J.; Tadi, S.; Tiziani, S.; Yan, W.; Triplett, K.; Lamb, C.; Alters, S.E.; Rowlinson, S.; et al. Systemic depletion of L-cyst(e)ine with cyst(e)inase increases reactive oxygen species and suppresses tumor growth. Nat. Med. 2017, 23, 120–127. [Google Scholar] [CrossRef]
  167. Tang, J.; Ramis-Cabrer, D.; Wang, X.; Barreiro, E. Immunotherapy with Monoclonal Antibodies in Lung Cancer of Mice: Oxidative Stress and Other Biological Events. Cancers 2019, 11, 1301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Chen, Q.; Espey, M.G.; Krishna, M.C.; Mitchell, J.B.; Corpe, C.P.; Buettner, G.R.; Shacter, E.; Levine, M. Pharmacologic ascorbic acid concentrations selectively kill cancer cells: Action as a pro-drug to deliver hydrogen peroxide to tissues. Proc. Natl. Acad. Sci. USA 2005, 102, 13604–13609. [Google Scholar] [CrossRef] [Green Version]
  169. Chen, Q.; Espey, M.G.; Sun, A.Y.; Lee, J.H.; Krishna, M.C.; Shacter, E.; Choyke, P.L.; Pooput, C.; Kirk, K.L.; Buettner, G.R.; et al. Ascorbate in pharmacologic concentrations selectively generates ascorbate radical and hydrogen peroxide in extracellular fluid in vivo. Proc. Natl. Acad. Sci. USA 2007, 104, 8749–8754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Chen, Q.; Espey, M.G.; Sun, A.Y.; Pooput, C.; Kirk, K.L.; Krishna, M.C.; Khosh, D.B.; Drisko, J.; Levine, M. Pharmacologic doses of ascorbate act as a prooxidant and decrease growth of aggressive tumor xenografts in mice. Proc. Natl. Acad. Sci. USA 2008, 105, 11105–11109. [Google Scholar] [CrossRef] [Green Version]
  171. Du, J.; Cullen, J.J.; Buettner, G.R. Ascorbic acid: Chemistry, biology and the treatment of cancer. Biochim. Biophys. Acta 2012, 1826, 443–457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Ma, Y.; Chapman, J.; Levine, M.; Polireddy, K.; Drisko, J.; Chen, Q. High-dose parenteral ascorbate enhanced chemosensitivity of ovarian cancer and reduced toxicity of chemotherapy. Sci. Transl. Med. 2014, 6, 222ra218. [Google Scholar] [CrossRef] [PubMed]
  173. Yun, J.; Mullarky, E.; Lu, C.; Bosch, K.N.; Kavalier, A.; Rivera, K.; Roper, J.; Chio, I.I.; Giannopoulou, E.G.; Rago, C.; et al. Vitamin C selectively kills KRAS and BRAF mutant colorectal cancer cells by targeting GAPDH. Science 2015, 350, 1391–1396. [Google Scholar] [CrossRef] [Green Version]
  174. Wei, X.; Xu, Y.; Xu, F.F.; Chaiswing, L.; Schnell, D.; Noel, T.; Wang, C.; Chen, J.; St Clair, D.K.; St Clair, W.H. RelB Expression Determines the Differential Effects of Ascorbic Acid in Normal and Cancer Cells. Cancer Res. 2017, 77, 1345–1356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Shenoy, N.; Creagan, E.; Witzig, T.; Levine, M. Ascorbic Acid in Cancer Treatment: Let the Phoenix Fly. Cancer Cell 2018, 34, 700–706. [Google Scholar] [CrossRef] [Green Version]
  176. Vissers, M.C.M.; Das, A.B. Potential Mechanisms of Action for Vitamin C in Cancer: Reviewing the Evidence. Front. Physiol 2018, 9, 809. [Google Scholar] [CrossRef] [Green Version]
  177. Creagan, E.T.; Moertel, C.G.; O’Fallon, J.R.; Schutt, A.J.; O’Connell, M.J.; Rubin, J.; Frytak, S. Failure of high-dose vitamin C (ascorbic acid) therapy to benefit patients with advanced cancer. A controlled trial. N. Engl. J. Med. 1979, 301, 687–690. [Google Scholar] [CrossRef] [PubMed]
  178. Moertel, C.G.; Fleming, T.R.; Creagan, E.T.; Rubin, J.; O’Connell, M.J.; Ames, M.M. High-dose vitamin C versus placebo in the treatment of patients with advanced cancer who have had no prior chemotherapy. A randomized double-blind comparison. N. Engl. J. Med. 1985, 312, 137–141. [Google Scholar] [CrossRef]
  179. Padayatty, S.J.; Sun, H.; Wang, Y.; Riordan, H.D.; Hewitt, S.M.; Katz, A.; Wesley, R.A.; Levine, M. Vitamin C pharmacokinetics: Implications for oral and intravenous use. Ann. Internt. Med. 2004, 140, 533–537. [Google Scholar] [CrossRef]
  180. Verrax, J.; Calderon, P.B. Pharmacologic concentrations of ascorbate are achieved by parenteral administration and exhibit antitumoral effects. Free Radic. Biol. Med. 2009, 47, 32–40. [Google Scholar] [CrossRef]
  181. Nauman, G.; Gray, J.C.; Parkinson, R.; Levine, M.; Paller, C.J. Systematic Review of Intravenous Ascorbate in Cancer Clinical Trials. Antioxidants 2018, 7, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Schoenfeld, J.D.; Sibenaller, Z.A.; Mapuskar, K.A.; Wagner, B.A.; Cramer-Morales, K.L.; Furqan, M.; Sandhu, S.; Carlisle, T.L.; Smith, M.C.; Abu Hejleh, T.; et al. O2(-) and H2O2-Mediated Disruption of Fe Metabolism Causes the Differential Susceptibility of NSCLC and GBM Cancer Cells to Pharmacological Ascorbate. Cancer Cell 2017, 32, 268. [Google Scholar] [CrossRef] [PubMed]
  183. Spitz, D.R.; Sim, J.E.; Ridnour, L.A.; Galoforo, S.S.; Lee, Y.J. Glucose deprivation-induced oxidative stress in human tumor cells. A fundamental defect in metabolism? Ann. N Y Acad. Sci. 2000, 899, 349–362. [Google Scholar] [CrossRef]
  184. Aykin-Burns, N.; Ahmad, I.M.; Zhu, Y.; Oberley, L.W.; Spitz, D.R. Increased levels of superoxide and H2O2 mediate the differential susceptibility of cancer cells versus normal cells to glucose deprivation. Biochem. J. 2009, 418, 29–37. [Google Scholar] [CrossRef] [Green Version]
  185. Caltagirone, A.; Weiss, G.; Pantopoulos, K. Modulation of cellular iron metabolism by hydrogen peroxide. Effects of H2O2 on the expression and function of iron-responsive element-containing mRNAs in B6 fibroblasts. J. Biol. Chem. 2001, 276, 19738–19745. [Google Scholar] [CrossRef] [Green Version]
  186. Torti, S.V.; Torti, F.M. Iron and cancer: More ore to be mined. Nat. Rev. Cancer 2013, 13, 342–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Galaris, D.; Barbouti, A.; Pantopoulos, K. Iron homeostasis and oxidative stress: An intimate relationship. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 118535. [Google Scholar] [CrossRef]
  188. Shenoy, N.; Bhagat, T.; Nieves, E.; Stenson, M.; Lawson, J.; Choudhary, G.S.; Habermann, T.; Nowakowski, G.; Singh, R.; Wu, X.; et al. Upregulation of TET activity with ascorbic acid induces epigenetic modulation of lymphoma cells. Blood Cancer J. 2017, 7, e587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Shenoy, N.; Bhagat, T.D.; Cheville, J.; Lohse, C.; Bhattacharyya, S.; Tischer, A.; Machha, V.; Gordon-Mitchell, S.; Choudhary, G.; Wong, L.F.; et al. Ascorbic acid-induced TET activation mitigates adverse hydroxymethylcytosine loss in renal cell carcinoma. J. Clin. Invest. 2019, 129, 1612–1625. [Google Scholar] [CrossRef] [Green Version]
  190. Cimmino, L.; Dolgalev, I.; Wang, Y.; Yoshimi, A.; Martin, G.H.; Wang, J.; Ng, V.; Xia, B.; Witkowski, M.T.; Mitchell-Flack, M.; et al. Restoration of TET2 Function Blocks Aberrant Self-Renewal and Leukemia Progression. Cell 2017, 170, 1079–1095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Luchtel, R.A.; Bhagat, T.; Pradhan, K.; Jacobs, W.R., Jr.; Levine, M.; Verma, A.; Shenoy, N. High-dose ascorbic acid synergizes with anti-PD1 in a lymphoma mouse model. Proc. Natl. Acad. Sci. USA 2020, 117, 1666–1677. [Google Scholar] [CrossRef] [Green Version]
  192. Magri, A.; Germano, G.; Lorenzato, A.; Lamba, S.; Chila, R.; Montone, M.; Amodio, V.; Ceruti, T.; Sassi, F.; Arena, S.; et al. High-dose vitamin C enhances cancer immunotherapy. Sci. Transl. Med. 2020, 12, eaay8707. [Google Scholar] [CrossRef]
  193. Brune, K.; Patrignani, P. New insights into the use of currently available non-steroidal anti-inflammatory drugs. J. Pain Res. 2015, 8, 105–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Janne, P.A.; Mayer, R.J. Chemoprevention of colorectal cancer. N. Engl. J. Med. 2000, 342, 1960–1968. [Google Scholar] [CrossRef] [PubMed]
  195. Keller, J.J.; Giardiello, F.M. Chemoprevention strategies using NSAIDs and COX-2 inhibitors. Cancer Biol Ther 2003, 2, S140–S149. [Google Scholar] [CrossRef]
  196. Rao, C.V.; Reddy, B.S. NSAIDs and chemoprevention. Curr. Cancer Drug Targets 2004, 4, 29–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Wang, D.; Dubois, R.N. Prostaglandins and cancer. Gut 2006, 55, 115–122. [Google Scholar] [CrossRef]
  198. Finetti, F.; Travelli, C.; Ercoli, J.; Colombo, G.; Buoso, E.; Trabalzini, L. Prostaglandin E2 and Cancer: Insight into Tumor Progression and Immunity. Biology 2020, 9, 434. [Google Scholar] [CrossRef]
  199. Tinsley, H.N.; Gary, B.D.; Keeton, A.B.; Lu, W.; Li, Y.; Piazza, G.A. Inhibition of PDE5 by sulindac sulfide selectively induces apoptosis and attenuates oncogenic Wnt/beta-catenin-mediated transcription in human breast tumor cells. Cancer Prev. Res. 2011, 4, 1275–1284. [Google Scholar] [CrossRef] [Green Version]
  200. Li, N.; Xi, Y.; Tinsley, H.N.; Gurpinar, E.; Gary, B.D.; Zhu, B.; Li, Y.; Chen, X.; Keeton, A.B.; Abadi, A.H.; et al. Sulindac selectively inhibits colon tumor cell growth by activating the cGMP/PKG pathway to suppress Wnt/beta-catenin signaling. Mol. Cancer Ther. 2013, 12, 1848–1859. [Google Scholar] [CrossRef] [Green Version]
  201. Li, N.; Chen, X.; Zhu, B.; Ramirez-Alcantara, V.; Canzoneri, J.C.; Lee, K.; Sigler, S.; Gary, B.; Li, Y.; Zhang, W.; et al. Suppression of beta-catenin/TCF transcriptional activity and colon tumor cell growth by dual inhibition of PDE5 and 10. Oncotarget 2015, 6, 27403–27415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Li, N.; Lee, K.; Xi, Y.; Zhu, B.; Gary, B.D.; Ramirez-Alcantara, V.; Gurpinar, E.; Canzoneri, J.C.; Fajardo, A.; Sigler, S.; et al. Phosphodiesterase 10A: A novel target for selective inhibition of colon tumor cell growth and beta-catenin-dependent TCF transcriptional activity. Oncogene 2015, 34, 1499–1509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Tsutsumi, S.; Gotoh, T.; Tomisato, W.; Mima, S.; Hoshino, T.; Hwang, H.J.; Takenaka, H.; Tsuchiya, T.; Mori, M.; Mizushima, T. Endoplasmic reticulum stress response is involved in nonsteroidal anti-inflammatory drug-induced apoptosis. Cell Death Differ. 2004, 11, 1009–1016. [Google Scholar] [CrossRef] [Green Version]
  204. Adachi, M.; Sakamoto, H.; Kawamura, R.; Wang, W.; Imai, K.; Shinomura, Y. Nonsteroidal anti-inflammatory drugs and oxidative stress in cancer cells. Histol. Histopathol. 2007, 22, 437–442. [Google Scholar] [CrossRef] [PubMed]
  205. Ou, Y.C.; Yang, C.R.; Cheng, C.L.; Raung, S.L.; Hung, Y.Y.; Chen, C.J. Indomethacin induces apoptosis in 786-O renal cell carcinoma cells by activating mitogen-activated protein kinases and AKT. Eur. J. Pharmacol. 2007, 563, 49–60. [Google Scholar] [CrossRef]
  206. Du, H.; Li, W.; Wang, Y.; Chen, S.; Zhang, Y. Celecoxib induces cell apoptosis coupled with up-regulation of the expression of VEGF by a mechanism involving ER stress in human colorectal cancer cells. Oncol. Rep. 2011, 26, 495–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Zhang, X.; Lee, S.H.; Min, K.W.; McEntee, M.F.; Jeong, J.B.; Li, Q.; Baek, S.J. The involvement of endoplasmic reticulum stress in the suppression of colorectal tumorigenesis by tolfenamic acid. Cancer Prev. Res. 2013, 6, 1337–1347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Cha, W.; Park, S.W.; Kwon, T.K.; Hah, J.H.; Sung, M.W. Endoplasmic reticulum stress response as a possible mechanism of cyclooxygenase-2-independent anticancer effect of celecoxib. Anticancer Res. 2014, 34, 1731–1735. [Google Scholar] [PubMed]
  209. Ralph, S.J.; Pritchard, R.; Rodriguez-Enriquez, S.; Moreno-Sanchez, R.; Ralph, R.K. Hitting the Bull’s-Eye in Metastatic Cancers-NSAIDs Elevate ROS in Mitochondria, Inducing Malignant Cell Death. Pharmaceuticals 2015, 8, 62–106. [Google Scholar] [CrossRef] [Green Version]
  210. Marchetti, M.; Resnick, L.; Gamliel, E.; Kesaraju, S.; Weissbach, H.; Binninger, D. Sulindac enhances the killing of cancer cells exposed to oxidative stress. PLoS ONE 2009, 4, e5804. [Google Scholar] [CrossRef] [Green Version]
  211. Raza, H.; John, A.; Benedict, S. Acetylsalicylic acid-induced oxidative stress, cell cycle arrest, apoptosis and mitochondrial dysfunction in human hepatoma HepG2 cells. Eur. J. Pharmacol. 2011, 668, 15–24. [Google Scholar] [CrossRef] [PubMed]
  212. Tse, A.K.; Cao, H.H.; Cheng, C.Y.; Kwan, H.Y.; Yu, H.; Fong, W.F.; Yu, Z.L. Indomethacin sensitizes TRAIL-resistant melanoma cells to TRAIL-induced apoptosis through ROS-mediated upregulation of death receptor 5 and downregulation of survivin. J. Invest. Dermatol. 2014, 134, 1397–1407. [Google Scholar] [CrossRef] [Green Version]
  213. Pritchard, R.; Rodriguez-Enriquez, S.; Pacheco-Velazquez, S.C.; Bortnik, V.; Moreno-Sanchez, R.; Ralph, S. Celecoxib inhibits mitochondrial O2 consumption, promoting ROS dependent death of murine and human metastatic cancer cells via the apoptotic signalling pathway. Biochem. Pharmacol. 2018, 154, 318–334. [Google Scholar] [CrossRef] [Green Version]
  214. Ralph, S.J.; Nozuhur, S.; Moreno-Sánchez, R.; Rodríguez-Enríquez, S.; Pritchard, R. NSAID celecoxib: A potent mitochondrial pro-oxidant cytotoxic agent sensitizing metastatic cancers and cancer stem cells to chemotherapy. J. Cancer Metastasis Treat. 2018, 4, 49. [Google Scholar] [CrossRef] [Green Version]
  215. Sun, Y.; Rigas, B. The thioredoxin system mediates redox-induced cell death in human colon cancer cells: Implications for the mechanism of action of anticancer agents. Cancer Res. 2008, 68, 8269–8277. [Google Scholar] [CrossRef] [Green Version]
  216. Rousar, T.; Parik, P.; Kucera, O.; Bartos, M.; Cervinkova, Z. Glutathione reductase is inhibited by acetaminophen-glutathione conjugate in vitro. Physiol. Res. 2010, 59, 225–232. [Google Scholar] [CrossRef] [PubMed]
  217. Raza, H.; John, A. Implications of altered glutathione metabolism in aspirin-induced oxidative stress and mitochondrial dysfunction in HepG2 cells. PLoS ONE 2012, 7, e36325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Hussain, M.; Javeed, A.; Ashraf, M.; Al-Zaubai, N.; Stewart, A.; Mukhtar, M.M. Non-steroidal anti-inflammatory drugs, tumour immunity and immunotherapy. Pharmacol. Res. 2012, 66, 7–18. [Google Scholar] [CrossRef]
  219. Serafini, P.; Meckel, K.; Kelso, M.; Noonan, K.; Califano, J.; Koch, W.; Dolcetti, L.; Bronte, V.; Borrello, I. Phosphodiesterase-5 inhibition augments endogenous antitumor immunity by reducing myeloid-derived suppressor cell function. J. Exp. Med. 2006, 203, 2691–2702. [Google Scholar] [CrossRef]
  220. Noonan, K.A.; Ghosh, N.; Rudraraju, L.; Bui, M.; Borrello, I. Targeting immune suppression with PDE5 inhibition in end-stage multiple myeloma. Cancer Immunol. Res. 2014, 2, 725–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Spranger, S.; Bao, R.; Gajewski, T.F. Melanoma-intrinsic beta-catenin signalling prevents anti-tumour immunity. Nature 2015, 523, 231–235. [Google Scholar] [CrossRef]
  222. Spranger, S.; Dai, D.; Horton, B.; Gajewski, T.F. Tumor-Residing Batf3 Dendritic Cells Are Required for Effector T Cell Trafficking and Adoptive T Cell Therapy. Cancer Cell 2017, 31, 711–723. [Google Scholar] [CrossRef] [Green Version]
  223. Fletcher, R.; Tong, J.; Risnik, D.; Leibowitz, B.J.; Wang, Y.-J.; Concha-Benavente, F.; DeLiberty, J.M.; Stolz, D.B.; Pai, R.K.; Ferris, R.L.; et al. Non-steroidal anti-inflammatory drugs induce immunogenic cell death in suppressing colorectal tumorigenesis. Oncogene 2021. [Google Scholar] [CrossRef] [PubMed]
  224. Majchrzak, K.; Nelson, M.H.; Bowers, J.S.; Bailey, S.R.; Wyatt, M.M.; Wrangle, J.M.; Rubinstein, M.P.; Varela, J.C.; Li, Z.; Himes, R.A.; et al. beta-catenin and PI3Kdelta inhibition expands precursor Th17 cells with heightened stemness and antitumor activity. JCI Insight 2017, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Zelenay, S.; van der Veen, A.G.; Bottcher, J.P.; Snelgrove, K.J.; Rogers, N.; Acton, S.E.; Chakravarty, P.; Girotti, M.R.; Marais, R.; Quezada, S.A.; et al. Cyclooxygenase-Dependent Tumor Growth through Evasion of Immunity. Cell 2015, 162, 1257–1270. [Google Scholar] [CrossRef] [Green Version]
  226. Liu, J.; Xia, X.; Huang, P. xCT: A Critical Molecule That Links Cancer Metabolism to Redox Signaling. Mol. Ther. 2020, 28, 2358–2366. [Google Scholar] [CrossRef]
  227. Huang, Y.; Dai, Z.; Barbacioru, C.; Sadee, W. Cystine-glutamate transporter SLC7A11 in cancer chemosensitivity and chemoresistance. Cancer Res. 2005, 65, 7446–7454. [Google Scholar] [CrossRef] [Green Version]
  228. Lo, M.; Wang, Y.Z.; Gout, P.W. The x(c)- cystine/glutamate antiporter: A potential target for therapy of cancer and other diseases. J. Cell Physiol. 2008, 215, 593–602. [Google Scholar] [CrossRef]
  229. Daher, B.; Vucetic, M.; Pouyssegur, J. Cysteine Depletion, a Key Action to Challenge Cancer Cells to Ferroptotic Cell Death. Front. Oncol. 2020, 10, 723. [Google Scholar] [CrossRef]
  230. Wang, W.; Zou, W. Amino Acids and Their Transporters in T Cell Immunity and Cancer Therapy. Mol. Cell 2020, 80, 384–395. [Google Scholar] [CrossRef] [PubMed]
  231. Koppula, P.; Zhuang, L.; Gan, B. Cystine transporter SLC7A11/xCT in cancer: Ferroptosis, nutrient dependency, and cancer therapy. Protein Cell 2020. [Google Scholar] [CrossRef]
  232. Gout, P.W.; Buckley, A.R.; Simms, C.R.; Bruchovsky, N. Sulfasalazine, a potent suppressor of lymphoma growth by inhibition of the x(c)- cystine transporter: A new action for an old drug. Leukemia 2001, 15, 1633–1640. [Google Scholar] [CrossRef] [Green Version]
  233. Nagane, M.; Kanai, E.; Shibata, Y.; Shimizu, T.; Yoshioka, C.; Maruo, T.; Yamashita, T. Sulfasalazine, an inhibitor of the cystine-glutamate antiporter, reduces DNA damage repair and enhances radiosensitivity in murine B16F10 melanoma. PLoS ONE 2018, 13, e0195151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Iida, Y.; Okamoto-Katsuyama, M.; Maruoka, S.; Mizumura, K.; Shimizu, T.; Shikano, S.; Hikichi, M.; Takahashi, M.; Tsuya, K.; Okamoto, S.; et al. Effective ferroptotic small-cell lung cancer cell death from SLC7A11 inhibition by sulforaphane. Oncol. Lett. 2021, 21, 71. [Google Scholar] [CrossRef] [PubMed]
  235. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang, W.S.; et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef] [Green Version]
  236. Liu, D.S.; Duong, C.P.; Haupt, S.; Montgomery, K.G.; House, C.M.; Azar, W.J.; Pearson, H.B.; Fisher, O.M.; Read, M.; Guerra, G.R.; et al. Inhibiting the system xC(-)/glutathione axis selectively targets cancers with mutant-p53 accumulation. Nat. Commun. 2017, 8, 14844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Sato, M.; Kusumi, R.; Hamashima, S.; Kobayashi, S.; Sasaki, S.; Komiyama, Y.; Izumikawa, T.; Conrad, M.; Bannai, S.; Sato, H. The ferroptosis inducer erastin irreversibly inhibits system xc- and synergizes with cisplatin to increase cisplatin’s cytotoxicity in cancer cells. Sci. Rep. 2018, 8, 968. [Google Scholar] [CrossRef] [Green Version]
  238. Cobler, L.; Zhang, H.; Suri, P.; Park, C.; Timmerman, L.A. xCT inhibition sensitizes tumors to gamma-radiation via glutathione reduction. Oncotarget 2018, 9, 32280–32297. [Google Scholar] [CrossRef] [PubMed]
  239. Zhao, Y.; Li, Y.; Zhang, R.; Wang, F.; Wang, T.; Jiao, Y. The Role of Erastin in Ferroptosis and Its Prospects in Cancer Therapy. Oncol. Targets Ther. 2020, 13, 5429–5441. [Google Scholar] [CrossRef]
  240. Badgley, M.A.; Kremer, D.M.; Maurer, H.C.; DelGiorno, K.E.; Lee, H.J.; Purohit, V.; Sagalovskiy, I.R.; Ma, A.; Kapilian, J.; Firl, C.E.M.; et al. Cysteine depletion induces pancreatic tumor ferroptosis in mice. Science 2020, 368, 85–89. [Google Scholar] [CrossRef]
  241. Arensman, M.D.; Yang, X.S.; Leahy, D.M.; Toral-Barza, L.; Mileski, M.; Rosfjord, E.C.; Wang, F.; Deng, S.; Myers, J.S.; Abraham, R.T.; et al. Cystine-glutamate antiporter xCT deficiency suppresses tumor growth while preserving antitumor immunity. Proc. Natl. Acad. Sci. USA 2019, 116, 9533–9542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Lang, X.; Green, M.D.; Wang, W.; Yu, J.; Choi, J.E.; Jiang, L.; Liao, P.; Zhou, J.; Zhang, Q.; Dow, A.; et al. Radiotherapy and Immunotherapy Promote Tumoral Lipid Oxidation and Ferroptosis via Synergistic Repression of SLC7A11. Cancer Discov. 2019, 9, 1673–1685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Xu, X.; Saw, P.E.; Tao, W.; Li, Y.; Ji, X.; Bhasin, S.; Liu, Y.; Ayyash, D.; Rasmussen, J.; Huo, M.; et al. ROS-Responsive Polyprodrug Nanoparticles for Triggered Drug Delivery and Effective Cancer Therapy. Adv. Mater. 2017, 29, 1700141. [Google Scholar] [CrossRef] [PubMed]
  244. Tao, W.; He, Z. ROS-responsive drug delivery systems for biomedical applications. Asian J. Pharm. Sci. 2018, 13, 101–112. [Google Scholar] [CrossRef]
  245. Peiro Cadahia, J.; Previtali, V.; Troelsen, N.S.; Clausen, M.H. Prodrug strategies for targeted therapy triggered by reactive oxygen species. Medchemcomm 2019, 10, 1531–1549. [Google Scholar] [CrossRef]
  246. Wang, Y.; Zhang, Y.; Ru, Z.; Song, W.; Chen, L.; Ma, H.; Sun, L. A ROS-responsive polymeric prodrug nanosystem with self-amplified drug release for PSMA (-) prostate cancer specific therapy. J. Nanobiotechnol. 2019, 17, 91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Yang, B.; Gao, J.; Pei, Q.; Xu, H.; Yu, H. Engineering Prodrug Nanomedicine for Cancer Immunotherapy. Adv. Sci. (Weinh) 2020, 7, 2002365. [Google Scholar] [CrossRef]
  248. Gong, Y.; Chen, M.; Tan, Y.; Shen, J.; Jin, Q.; Deng, W.; Sun, J.; Wang, C.; Liu, Z.; Chen, Q. Injectable Reactive Oxygen Species-Responsive SN38 Prodrug Scaffold with Checkpoint Inhibitors for Combined Chemoimmunotherapy. ACS Appl. Mater. Interfaces 2020, 12, 50248–50259. [Google Scholar] [CrossRef]
  249. Kuang, Y.; Balakrishnan, K.; Gandhi, V.; Peng, X. Hydrogen peroxide inducible DNA cross-linking agents: Targeted anticancer prodrugs. J. Am. Chem. Soc. 2011, 133, 19278–19281. [Google Scholar] [CrossRef] [Green Version]
  250. Peng, X.; Gandhi, V. ROS-activated anticancer prodrugs: A new strategy for tumor-specific damage. Ther. Deliv. 2012, 3, 823–833. [Google Scholar] [CrossRef] [Green Version]
  251. Chen, W.; Fan, H.; Balakrishnan, K.; Wang, Y.; Sun, H.; Fan, Y.; Gandhi, V.; Arnold, L.A.; Peng, X. Discovery and Optimization of Novel Hydrogen Peroxide Activated Aromatic Nitrogen Mustard Derivatives as Highly Potent Anticancer Agents. J. Med. Chem. 2018, 61, 9132–9145. [Google Scholar] [CrossRef]
  252. Chen, W.; Balakrishnan, K.; Kuang, Y.; Han, Y.; Fu, M.; Gandhi, V.; Peng, X. Reactive oxygen species (ROS) inducible DNA cross-linking agents and their effect on cancer cells and normal lymphocytes. J. Med. Chem. 2014, 57, 4498–4510. [Google Scholar] [CrossRef] [PubMed]
  253. Hagen, H.; Marzenell, P.; Jentzsch, E.; Wenz, F.; Veldwijk, M.R.; Mokhir, A. Aminoferrocene-based prodrugs activated by reactive oxygen species. J. Med. Chem. 2012, 55, 924–934. [Google Scholar] [CrossRef] [PubMed]
  254. Schikora, M.; Reznikov, A.; Chaykovskaya, L.; Sachinska, O.; Polyakova, L.; Mokhir, A. Activity of aminoferrocene-based prodrugs against prostate cancer. Bioorg. Med. Chem. Lett. 2015, 25, 3447–3450. [Google Scholar] [CrossRef]
  255. Daum, S.; Reshetnikov, M.S.V.; Sisa, M.; Dumych, T.; Lootsik, M.D.; Bilyy, R.; Bila, E.; Janko, C.; Alexiou, C.; Herrmann, M.; et al. Lysosome-Targeting Amplifiers of Reactive Oxygen Species as Anticancer Prodrugs. Angew. Chem. Int. Ed. Engl. 2017, 56, 15545–15549. [Google Scholar] [CrossRef] [PubMed]
  256. Reshetnikov, V.; Daum, S.; Janko, C.; Karawacka, W.; Tietze, R.; Alexiou, C.; Paryzhak, S.; Dumych, T.; Bilyy, R.; Tripal, P.; et al. ROS-Responsive N-Alkylaminoferrocenes for Cancer-Cell-Specific Targeting of Mitochondria. Angew. Chem. Int. Ed. Engl. 2018, 57, 11943–11946. [Google Scholar] [CrossRef] [PubMed]
  257. Yoo, H.J.; Liu, Y.; Wang, L.; Schubert, M.L.; Hoffmann, J.M.; Wang, S.; Neuber, B.; Huckelhoven-Krauss, A.; Gern, U.; Schmitt, A.; et al. Tumor-Specific Reactive Oxygen Species Accelerators Improve Chimeric Antigen Receptor T Cell Therapy in B Cell Malignancies. Int. J. Mol. Sci. 2019, 20, 2469. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Interactions between cells in the TME lead to changes in redox status. Tumor-reactive effector T cells (CD4+ and CD8+) can induce increased levels of ROS in cancer cells via the actions of IFNγ and TNFα. TNFα signaling in tumor cells activates NADPH oxidases, which lead to increased production of ROS. IFNγ signaling in tumor cells diminishes xCT expression through transcriptional inhibition, reducing tumor uptake of extracellular cystine and subsequent GSH synthesis. IFNγ signaling in cancer-associated fibroblasts (CAF) blocks the release of GSH and cysteine, further depleting the extracellular pool of cystine and cysteine available to cancer cells. The severe redox imbalance in cancer cells, caused by TNFα-driven ROS production and IFNγ-induced GSH deficiency, leads to extensive oxidative damages and eventual tumor cell death. However, effector T cells are also susceptible to oxidative stress in the TME. ROS induced upon TCR engagement are counterbalanced by increased antioxidant systems such as GSH and catalase. T cell dysfunction may occur when effector T cells are exposed to ROS produced by MDSCs and Tregs, which are more resistant to oxidative stress due to their increased antioxidant systems. High levels of extracellular ROS can disrupt antigen-presentation between T cells and DCs, and can affect tumor antigen recognition by T cells. Some apoptosis-prone Tregs can increase the presence of adenosine in the TME, which suppresses the function of effector T cells in an A2AR-dependent manner. Therapeutic interventions should be directed to enhance T cell-induced tumor oxidative stress while enabling T cells to resist the elevated oxidative stress in the TME.
Figure 1. Interactions between cells in the TME lead to changes in redox status. Tumor-reactive effector T cells (CD4+ and CD8+) can induce increased levels of ROS in cancer cells via the actions of IFNγ and TNFα. TNFα signaling in tumor cells activates NADPH oxidases, which lead to increased production of ROS. IFNγ signaling in tumor cells diminishes xCT expression through transcriptional inhibition, reducing tumor uptake of extracellular cystine and subsequent GSH synthesis. IFNγ signaling in cancer-associated fibroblasts (CAF) blocks the release of GSH and cysteine, further depleting the extracellular pool of cystine and cysteine available to cancer cells. The severe redox imbalance in cancer cells, caused by TNFα-driven ROS production and IFNγ-induced GSH deficiency, leads to extensive oxidative damages and eventual tumor cell death. However, effector T cells are also susceptible to oxidative stress in the TME. ROS induced upon TCR engagement are counterbalanced by increased antioxidant systems such as GSH and catalase. T cell dysfunction may occur when effector T cells are exposed to ROS produced by MDSCs and Tregs, which are more resistant to oxidative stress due to their increased antioxidant systems. High levels of extracellular ROS can disrupt antigen-presentation between T cells and DCs, and can affect tumor antigen recognition by T cells. Some apoptosis-prone Tregs can increase the presence of adenosine in the TME, which suppresses the function of effector T cells in an A2AR-dependent manner. Therapeutic interventions should be directed to enhance T cell-induced tumor oxidative stress while enabling T cells to resist the elevated oxidative stress in the TME.
Cancers 13 00986 g001
Figure 2. Hypothetical mechanisms by which certain pro-oxidants enhance the efficacy of cancer immunotherapies. Tumor-reactive effector T cells, emerging in the TME after either ICB therapy or adoptive transfer, can mediate tumor killing via well-characterized mechanisms involving cytolytic granules such as perforin and granzymes, and apoptosis-inducing ligands such as FASL and TRAIL. Increasing evidence reveals that inflammatory cytokines produced by effector T cells, including TNFα and IFNγ, can exert antitumor effect by modulating tumor redox. TNFα signaling in cancer cells activates NOX-dependent ROS production, while IFNγ signaling exacerbates GSH deficiency by suppressing the cysteine/glutamate transporter xCT. The combined effects of TNFα and IFNγ lead to substantial ROS accumulation in tumor cells, rendering them vulnerable to further redox disruption which can be incited by a pro-oxidant. A suitable pro-oxidant should preferentially induce oxidative stress in cancer cells without harming antitumor T cells. The potential mechanisms of action of four immunotherapy-compatible pro-oxidants are illustrated. (A). Pharmacological dose of ascorbate, in its oxidized form DHA, can be preferentially taken into cancer cells via the glucose transporter (GLUT1). Intracellular DHA is reduced to ascorbate at the expense of GSH. The GSH shortage aggravates ROS accumulation, which damages DNA/protein/lipid and derails cell metabolism, leading to tumor cell death. Meanwhile, ascorbate may induce TET activities in antitumor T cells and tumor cells. TET activation in T cells leads to enhanced function of T cells through epigenetic modifications. TET activation in tumor cells results in demethylation and activation of SMAD1, which increases tumor chemosensitivity. (B). NSAIDs can act as pro-oxidants to reduce GSH and thereby increase the levels of ROS in tumor cells. NSAID-induced ER stress may lead to release of calreticulin (CRT), a DAMP molecule characteristic of ICD, which can attract and activate DCs, which in turn elicit antitumor CD8+ T cell responses. In addition, NSAIDs can suppress tumor cell growth by its inhibitory effect on β-catenin, COX2, and PGE2. Some NSAIDs may reduce MDSC activity by inhibiting PDE5 function in MDSCs. (C). Cyst(e)inase or xCT inhibitors can reduce the presence or block the uptake of extracellular cystine and cysteine, which tumor cells rely on to synthesize GSH, respectively. Cyst(e)inase can act in concert with ICB-induced antitumor T cells to drive tumor cell ferroptosis. (D). ROS-responsive prodrugs can be effectively delivered to tumor loci by nanoparticles. The increased levels of ROS in the TME can activate these prodrugs, which give rise to alkylating metabolites to cause further DNA damage and intensify oxidative stress in tumor cells. These prodrugs may synergize with antitumor T cells because ROS accumulated in tumor cells after immunotherapies such as ICB or CAR-T therapy can effectively activate prodrugs, which in turn further amplify ROS in tumor cells to drive apoptosis.
Figure 2. Hypothetical mechanisms by which certain pro-oxidants enhance the efficacy of cancer immunotherapies. Tumor-reactive effector T cells, emerging in the TME after either ICB therapy or adoptive transfer, can mediate tumor killing via well-characterized mechanisms involving cytolytic granules such as perforin and granzymes, and apoptosis-inducing ligands such as FASL and TRAIL. Increasing evidence reveals that inflammatory cytokines produced by effector T cells, including TNFα and IFNγ, can exert antitumor effect by modulating tumor redox. TNFα signaling in cancer cells activates NOX-dependent ROS production, while IFNγ signaling exacerbates GSH deficiency by suppressing the cysteine/glutamate transporter xCT. The combined effects of TNFα and IFNγ lead to substantial ROS accumulation in tumor cells, rendering them vulnerable to further redox disruption which can be incited by a pro-oxidant. A suitable pro-oxidant should preferentially induce oxidative stress in cancer cells without harming antitumor T cells. The potential mechanisms of action of four immunotherapy-compatible pro-oxidants are illustrated. (A). Pharmacological dose of ascorbate, in its oxidized form DHA, can be preferentially taken into cancer cells via the glucose transporter (GLUT1). Intracellular DHA is reduced to ascorbate at the expense of GSH. The GSH shortage aggravates ROS accumulation, which damages DNA/protein/lipid and derails cell metabolism, leading to tumor cell death. Meanwhile, ascorbate may induce TET activities in antitumor T cells and tumor cells. TET activation in T cells leads to enhanced function of T cells through epigenetic modifications. TET activation in tumor cells results in demethylation and activation of SMAD1, which increases tumor chemosensitivity. (B). NSAIDs can act as pro-oxidants to reduce GSH and thereby increase the levels of ROS in tumor cells. NSAID-induced ER stress may lead to release of calreticulin (CRT), a DAMP molecule characteristic of ICD, which can attract and activate DCs, which in turn elicit antitumor CD8+ T cell responses. In addition, NSAIDs can suppress tumor cell growth by its inhibitory effect on β-catenin, COX2, and PGE2. Some NSAIDs may reduce MDSC activity by inhibiting PDE5 function in MDSCs. (C). Cyst(e)inase or xCT inhibitors can reduce the presence or block the uptake of extracellular cystine and cysteine, which tumor cells rely on to synthesize GSH, respectively. Cyst(e)inase can act in concert with ICB-induced antitumor T cells to drive tumor cell ferroptosis. (D). ROS-responsive prodrugs can be effectively delivered to tumor loci by nanoparticles. The increased levels of ROS in the TME can activate these prodrugs, which give rise to alkylating metabolites to cause further DNA damage and intensify oxidative stress in tumor cells. These prodrugs may synergize with antitumor T cells because ROS accumulated in tumor cells after immunotherapies such as ICB or CAR-T therapy can effectively activate prodrugs, which in turn further amplify ROS in tumor cells to drive apoptosis.
Cancers 13 00986 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aboelella, N.S.; Brandle, C.; Kim, T.; Ding, Z.-C.; Zhou, G. Oxidative Stress in the Tumor Microenvironment and Its Relevance to Cancer Immunotherapy. Cancers 2021, 13, 986. https://doi.org/10.3390/cancers13050986

AMA Style

Aboelella NS, Brandle C, Kim T, Ding Z-C, Zhou G. Oxidative Stress in the Tumor Microenvironment and Its Relevance to Cancer Immunotherapy. Cancers. 2021; 13(5):986. https://doi.org/10.3390/cancers13050986

Chicago/Turabian Style

Aboelella, Nada S., Caitlin Brandle, Timothy Kim, Zhi-Chun Ding, and Gang Zhou. 2021. "Oxidative Stress in the Tumor Microenvironment and Its Relevance to Cancer Immunotherapy" Cancers 13, no. 5: 986. https://doi.org/10.3390/cancers13050986

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