Skip to main content

Natural products for enhancing the sensitivity or decreasing the adverse effects of anticancer drugs through regulating the redox balance

Abstract

Redox imbalance is reported to play a pivotal role in tumorigenesis, cancer development, and drug resistance. Severe oxidative damage is a general consequence of cancer cell responses to treatment and may cause cancer cell death or severe adverse effects. To maintain their longevity, cancer cells can rescue redox balance and enter a state of resistance to anticancer drugs. Therefore, targeting redox signalling pathways has emerged as an attractive and prospective strategy for enhancing the efficacy of anticancer drugs and decreasing their adverse effects. Over the past few decades, natural products (NPs) have become an invaluable source for developing new anticancer drugs due to their high efficacy and low toxicity. Increasing evidence has demonstrated that many NPs exhibit remarkable antitumour effects, whether used alone or as adjuvants, and are emerging as effective approaches to enhance sensitivity and decrease the adverse effects of conventional cancer therapies by regulating redox balance. Among them are several novel anticancer drugs based on NPs that have entered clinical trials. In this review, we summarize the synergistic anticancer effects and related redox mechanisms of the combination of NPs with conventional anticancer drugs. We believe that NPs targeting redox regulation will represent promising novel candidates and provide prospects for cancer treatment in the future.

Graphical Abstract

Highlights

  • An abnormal redox microenvironment plays a critical role in the tumorigenesis, development, and drug resistance of cancer.

  • NPs can selectively act as pro-oxidants, inducing cytotoxicity in cancer cells, while also serving as antioxidants and protectors in normal cells for detoxification.

  • NPs play a dual role in increasing the sensitivity and reversing the drug resistance of cancer cells, primarily through pro-oxidant action and, in some cases, through antioxidant action.

  • The heterogeneities in cellular and molecular biology provide the potential basis for the diverse regulation of NPs on the redox state.

  • The advances in novel redox-sensitive materials and clinical applications of NPs offer promising prospects.

Introduction

The term “redox” is derived from the concepts of reduction and oxidation, which are processes that are pivotal for maintaining cellular homeostasis, regulating a wide array of cellular processes, and impacting the overall functionality of cells. The redox balance is dynamically responsive, with relatively reducing steady states and relatively oxidizing steady states, which are maintained by the regulation of oxidation and antioxidant systems [1]. Reactive oxygen and nitrogen species (ROS/RNS) are key players in the regulation of the redox microenvironment. ROS/RNS are a class of highly reactive radicals, ions, or molecules that contain unpaired electrons, primarily consisting of superoxide anion radical (O2• −), hydroxyl radical (OH), hydrogen peroxide (H2O2), nitric oxide (NO), and peroxynitrite (ONOO) [2, 3]. Moreover, the antioxidant defence system is activated in living cells to combat the harmful effects of oxidative stress. The reduced glutathione (GSH)/glutathione disulfide (GSSG) redox couple is the predominant intracellular antioxidant that acts as a free radical scavenger and inhibitor of peroxidation [4]. Another critical system responsible for maintaining cellular redox homeostasis is the thioredoxin (Trx)/thioredoxin reductase (TrxR, encoded by TXNRD) dependent system, which protects organisms from oxidative stress and regulates the expression of proteins via transcription factors [5]. These main antioxidant systems require nicotinamide adenine dinucleotide phosphate (NADPH) as an electron donor. In addition, within cells, a series of regulatory mechanisms are involved in redox processes, such as the Keap1/Nrf2 transcription factor, NF-κB signalling, the MAPK pathway, PI3K-Akt cascades, and the AMPK-mTOR axis [6, 7]. These systems complement each other and jointly participate in the complex regulation of intracellular redox homeostasis, particularly in cases involving pathological conditions.

Cancer tumorigenesis and development are extremely complex and involve multistep processes that are highly dependent on the environment, from tissues to patients [8]. Redox levels play a critical role in modulating cell fate. Appropriate redox homeostasis is involved in signal transduction pathways in cells and maintains normal biological functions. However, imbalanced redox causes oxidative damage to important intracellular biomolecules, leading to abnormalities in the structure and function of cells [9]. Oxidative stress is closely associated with cancers and is defined as a state in which the level of ROS overrides the antioxidant defence mechanisms of the cell [10]. Specifically, most cancer cells exhibit more ROS compared with normal cells, indicating higher levels of oxidative stress, which confers advantages for increasing carcinogenesis and development [11]. Moreover, cancer cells can maintain a delicate balance of intracellular ROS levels. The activity of antioxidant systems in drug-resistant cancer cells is significantly increased compared with that in drug-sensitive cells [12]. This self-regulation not only allows cancer cells to adapt to persistent intrinsic oxidative stress and prevents cell death but is also considered an important mechanism of treatment resistance. In addition, abnormal redox reactions in cancer cells may increase their susceptibility to damage by ROS induced by exogenous reagents [13]. To date, various drugs with regulatory effects on ROS levels have been used for cancer treatment. However, the toxicity of these traditional treatments on nontumour cells, which is partially caused by oxidative stress, remains a critical issue. Consequently, targeting redox regulation in cancer cells is expected to represent an effective approach for enhancing sensitivity or reducing adverse effects.

Natural products (NPs) have various sources in nature and have served as a cornerstone in the search for new cancer therapies because of their diverse biological activities and chemical structures [14] (Fig. 1). Although a series of studies have helped elucidate the anticancer mechanisms of NPs, the findings in this field are still obtained from separate and individual studies and remain controversial. Therefore, systematic and comprehensive retrospective analyses are necessary. Herein, we provide an overview of the mechanisms of sensitization and detoxification in cancer cells by NPs through the regulation of the redox microenvironment. Overall, considering the imbalanced redox in cancer cells, NPs that are useful for redox regulation represent a promising class of drug candidates that may offer potential clinical and patient benefits, suggesting a potential shift in the paradigm of cancer therapy based on NPs.

Fig. 1
figure 1

The number of PubMed publications related to NPs, cancers, and redox over the past few years. An increasing number of reports revealed that NPs have served as a cornerstone for cancer treatment involving redox regulation

Background: association between redox regulation and cancer

The regulation of redox reactions is an ancient topic that has been extensively explored and involves numerous oxidation processes and reduction equivalents (Fig. 2). The intracellular oxidation process involves several biological mechanisms, including the mitochondrial electron transport chain, oxidoreductase enzyme mediation, metal-catalysed oxidation, and exogenous stimulation. Typically, intracellular reductive equivalents are generated from reductive substrates, and the relevant active couples constitute the major antioxidant systems that maintain redox balance and prevent intracellular oxidative damage. Understanding this background can help us discover potential targets and apply them in research on either physiological or pathogenic processes.

Fig. 2
figure 2

Regulation of the intracellular redox balance. Oxidation processes generate ROS/RNS to serve as key players in the regulation of redox balance and involve electron transport during OXPHOS, oxidoreductase enzymes, the Fenton reaction, and exogenous factor stimulation. The aerobic metabolism of cancer cells also progresses via the Warburg effect, which reduces ROS produced by mitochondrial OXPHOS. Typical reductive equivalents are generated from antioxidant systems, including GSH/GSSG and Trx/TrxR active couples, both of which require the NADP+/NADPH system as an electron donor. These substances interact with each other and collectively participate in the complex regulation of intracellular redox homeostasis

Regulation of oxidative and antioxidant pathways

It has been reported that mitochondria in cancer cells are characterized by ROS overproduction [15]. Mitochondrial redox signalling relies mainly on ROS generated by the electron transport chain via oxidative phosphorylation (OXPHOS), in which significant changes in the potential of electrons are related to the reduction of oxygen [16]. Conversely, the aerobic metabolism of cancer cells also progresses away from mitochondrial oxidative metabolism, known as the Warburg effect, which reduces ROS production by OXPHOS through the upregulation of the glycolysis rate in energy metabolism [17]. In such cancer cells, ROS are also generated through enhanced metabolism caused by abnormal signal transduction and gene activation [18]. Other pro-oxidant enzymes include NADPH oxidase (NOX/DUOX), lipoxygenase (LOX), cyclooxygenase (COX), and nitric oxide synthase (NOS), among others [19]. Among them, NOXs are transmembrane enzymes that are involved in the production of ROS and are overexpressed in multiple malignancies, leading to redox imbalance and tumour progression; thus, NOXs constitute potential targets for cancer treatment [20]. Additionally, the oxidative process can also be mediated by transition metals (such as ferrous and copper ions) through a nonenzymatic process known as the Fenton reaction. Exogenous carcinogens that contribute to the oxidation load include radiotherapy, chemicals, stress, and other toxicants [19].

GSH represents the most abundant low-molecular-weight thiol in cells, and the nucleophilic cysteine-thiol (R‒SH) group is the reactive group of the molecule, mediating its various biological activities. The ratio of GSH (reduced) to its disulfide, GSSG (oxidized), contributes to the redox homeostasis of the cell, and a decrease of the GSH/GSSG ratio serves as a biomarker of oxidative stress [21]. First, because of the cysteine residue, GSH is readily oxidized nonenzymatically to disulfide by electrophilic substances [22]. Second, GSH is extensively used as a cosubstrate by antioxidant enzymes such as glutathione peroxidase (GPX), glutaredoxin (Grx), and certain glutathione S-transferases (GSTs) [23]. In addition, Grx has been shown to promote reversible protein S-glutathionylation using a glutathionyl radical as the proximal donor [24], thereby preventing ROS-induced irreversible protein oxidative damage. GST is a key enzyme that catalyses the conjugation of GSH to a variety of electrophilic substances and the formation of GS-X and plays a critical role in cellular detoxification and signalling [25]. As a widely applied antioxidant, N-acetylcysteine (NAC) affects redox regulation by replenishing GSH, reducing disulfide bonds in proteins or other molecules, or directly scavenging oxidants and is often used as a supplement that may help with various conditions [26].

The Trx/TrxR active couple is overexpressed in various types of cancers [27]. Trx1 and Trx2 are two Trxs ubiquitously expressed in mammalian cells. As a thiol-disulfide reductase, Trx can catalyse the reduction of disulfides (S–S) within oxidized cellular proteins such as peroxiredoxin (Prx, HGNC root symbol PRDX) due to the presence of two cysteines in Trx’s active centre, Cys-Gly-Pro-Cys [28]. Trx also participates in cofactor activities and regulates the expression of many redox-related transcription factors, including nuclear factor kappa B (NF-κB), nuclear factor erythroid 2-related factor 2 (Nrf2), and redox factor-1 (Ref-1) [28]. TrxR is a homodimeric selenocysteine-containing flavoprotein that controls the redox state of Trx. The oxidized Trx can be further regenerated through reduction by TrxR at the expense of electrons provided by NADPH [29]. Three mammalian isoforms of TrxR have been characterized in humans: cytosolic TrxR1, mitochondrial TrxR2, and TrxR3 [30]. As an endogenous negative regulator of Trx, thioredoxin-interacting protein (TXNIP) is also a key redox protein and a potential therapeutic target [31].

Both GSH/GSSG and Trx/TrxR ultimately rely on the reducing power of NADPH, which functions as an important cofactor to provide electrons in antioxidant defence systems [32]. NADPH donates two electrons to reduce GSSG to GSH via GR. TrxR transfers electrons from NADPH to reduce oxidized thioredoxin (Trx-S2) to Trx-(SH)2, which provides reducing equivalents in the enzymatic removal of H2O2 and other organic hydroperoxides by Prx [33]. The generated NADP+ can primarily be reduced back to NADPH mainly by G6PD in the pentose phosphate pathway in the cytoplasm and by IDH2 and NNT in the mitochondria [34]. The regulation of enzymes involved in the biosynthesis of NADP+ and NADPH can influence NADP+/NADPH levels and the cellular redox potential [35]. Although NADPH is considered to be the electron source of antioxidant systems, it also acts as a substrate for the ROS-producing NOX family [36]. Overall, cellular redox homeostasis primarily results from a delicate balance between NADPH-dependent antioxidant systems (GSH/GSSG, Trx/TrxR, and Prx) and NADPH-dependent NOX or oxidation processes.

In addition to these critical redox systems, a series of transcription factors and signalling pathways participate in redox regulation within cells. Under normal conditions, Nrf2 is localized to the cytoplasm, where it interacts with Keap1. However, Keap1 becomes oxidized when exposed to oxidative or electrophilic stress, resulting in the synthesis of Nrf2 and its entry into the nucleus to regulate target gene expression [37]. Nrf2 activation is also regulated by the Trx/TrxR system, which acts as a negative regulator of Nrf2 transactivation [38]. Nrf2 has several target genes and regulates the expression of various antioxidants, such as GST, GPX, catalase (CAT), superoxide dismutase (SOD), and haem oxygenase-1 (HO-1) [37]. NF-κB is another redox-sensitive factor that can interact with the Nrf2 pathway. One of its most important influences in regulating redox balance is the increase in antioxidant proteins [39]. However, it is remarkably noteworthy that the regulation of redox reactions by NF-κB may be bidirectional, as it also plays a pro-oxidant role [39]. In addition, multiple signalling pathways can perceive redox changes and mediate downstream effects. As messengers, ROS/RNS-mediated cellular signalling involving the MAPK (ERK/JNK/p38), PI3K-Akt, and PKC signalling pathways, has been well reviewed [40, 41].

Redox imbalance, carcinogenesis, and drug resistance

Redox imbalance is considered an important stressor that controls tumour cell growth. Most reports on the associations between oxidative stress and carcinogenesis/progression revolve around ROS, which serve as the core factor that influences redox balance [42, 43]. ROS are considered to contribute to both tumorigenesis and development through the following mechanisms: (1) inducing DNA damage through the oxidation of nucleobases [44]; (2) regulating redox-sensitive transcription factors in cancer, including NF-κB, Nrf2, p53, and AP-1 [45]; (3) influencing the expression of oncogenes and tumor-suppressor genes by epigenetic modifications [46]; (4) acting as signalling molecules to drive cellular proliferation via the PI3K/AKT/mTOR and MAPK/ERK mitogenic signalling cascades [47]; and (5) promoting epithelial-to-mesenchymal transition [48]. Cancer cells generally exhibit increased oxidative stress, which promotes tumour initiation, growth, and proliferation; however, excessive ROS is harmful to cells. When ROS accumulation exceeds the tipping point, their carcinogenic effects on proliferation and development are shifted to antitumour effects via the induction of programmed cell death (PCD) [47, 49], including apoptosis, autophagy, necroptosis, pyroptosis, and ferroptosis. To neutralize ubiquitously elevated oxidative stress and maintain favourable redox homeostasis, cancer cells are able to adjust multiple antioxidant enzymes that support tumour progression, resulting in increased levels of GSH/GSSG [50], Trx/TrxR [27], NADPH [32], and other related proteins. The adaptation mechanism of cancer cells not only contributes to their survival under conditions of persistent oxidative stress but also results in resistance to certain anticancer agents [51]. This finding sets the tone for redox-mediated treatment, demonstrating that targeting redox regulation is a potent candidate approach for cancer therapy.

The efficacy of cancer therapies is partly due to the ROS production and consequent induction of oxidative stress in cancer cells. However, cancer cells can develop resistance to oxidative stress provoked by the treatment, which mechanistically involves an increase in antioxidation systems. For example, compared to those in drug-sensitive cells, the ratios of glutamate/glutamine and GSH/GSSG on the membrane of drug-resistant tumour cells are significantly greater (indicating “glutamine addiction”) [52]. Glutaminase (GLS) catalyses glutaminolysis to promote the formation of glutamate, which is important for GSH synthesis. Sorafenib-resistant hepatocellular carcinoma cells exhibit upregulated GLS1 and resistance to oxidative stress. Inhibiting glutamine metabolism sensitizes sorafenib-resistant cells to sorafenib [53]. The Trx/TrxR system is another contributor to cancer resistance, and increased Trx/TrxR expression is associated with cancer cell resistance to various chemotherapeutic agents. For example, Trx confers a growth advantage to pancreatic cancer cells and increases their resistance to cisplatin-induced apoptosis [54]. In malignant melanoma cells, intracellular Trx/TrxR expression together with endogenous TNFα is correlated with resistance to TNFα-induced cytotoxicity [55]. Trx1 is also involved in paclitaxel-induced drug resistance in ovarian cancer cells [56]. In addition, higher levels of NADPH are observed in drug-resistant cancer cells, which are considered to be more resistant to oxidative stress and to help to prevent ROS-mediated damage [53]. Therefore, targeting the antioxidation system that maintains cancer cell adaptation is a candidate approach for overcoming drug resistance. Some compounds have been shown to be effective in this context. For example, Trx inhibitors such as auranofin have entered clinical trials to improve therapeutic sensitivity and may be of significant clinical value [57].

Discovery: anticancer effects of NPs by regulating the redox balance

NPs are naturally derived from a wide range of plant sources and have various biological activities; thus, they constitute a cornerstone in the search for new cancer therapies. Here, we provide several typical examples to illustrate different types of NPs and their anticancer mechanisms (Fig. 3), and it is convincing that NPs act as attractive oxidative stress regulators. Given their excellent pharmacological potential, NPs and their derivatives and analogues are gradually being incorporated into therapeutic drug repositories for cancer treatment.

Fig. 3
figure 3

The anticancer activities of NPs in different cancers involve redox regulation. Various types of NPs, including terpenoids, phenolics, flavonoids, quinones, and alkaloids, exert inhibitory effects on multiple cancers by regulating the redox state, primarily by targeting intracellular antioxidant systems or directly promoting ROS generation, thereby inducing PCD

Terpenoids

Terpenoids constitute the largest class of NPs and have attracted considerable attention as drug libraries. The regulation of the redox state has been well established as a reason for the anticancer effect of terpenoids. For example, andrographolide reacts with the thiol group of GSH, thus increasing the level of intracellular ROS, and the combination of andrographolide and tanshinone IIA has synergistic effects by promoting crosstalk between ROS and p53 [58]. Tanshinone IIA promotes ferroptosis in breast [59] and gastric [60] cancers by inhibiting SLC7A11, which is accompanied by increased ROS. β-Elemene triggers ferroptosis and increases lipid ROS in lung cancer cells, which may be induced by TFEB-mediated GPX4 degradation [61]. Celastrol suppresses tumours by covalently targeting PRDX in both gastric [62] and colorectal [63] cancers. Celastrol triggers apoptosis and autophagy by activating the ROS/JNK pathway, as observed in cases of osteosarcoma [64] and glioma [65]. Artesunate induces autophagy-dependent apoptosis in bladder cancer by activating the ROS/AMPK/mTOR/ULK1 pathway [66]. Artesunate has also been identified as an activator of ROS-dependent ferroptosis in several cancers [67, 68].

One of the anticancer mechanisms of terpenoids is the Michael addition of their α,β-unsaturated carbonyl units to cellular thiols, which can disrupt the redox state [69]. Rabdosia rubescens is a perennial herb native to regions of Asia that has been used for hundreds of years in traditional Chinese medicine [70]. Oridonin is one of the most important compounds isolated from Rabdosia rubescens. An investigation of the relationship between the structure and anticancer activity of oridonin revealed that the enone system in the D-ring of oridonin is the active anticancer centre [71]. This structure can undergo irreversible Michael addition reactions with nucleophilic groups of receptor amino acids/proteins, which mediate the regulation of redox-related targets by oridonin. As mentioned previously, GSH expression is generally increased in cancer cells. The anticancer effect of oridonin through GSH depletion has been reported in several human cancers, such as esophageal squamous cell carcinoma [72] and epidermoid carcinoma [73]. Supplementation with GSH or the GSH precursor NAC abolished the toxic effect of oridonin, suggesting that oridonin-induced cell death is ROS dependent. Moreover, the Michael effect between oridonin and GSH was greater than that between ponicidin and GSH, which may explain the better anticancer activity of oridonin. Novel dienone analogues with an additional α,β-unsaturated ketone system on the A-ring of oridonin have been prepared and demonstrated efficacy in inducing apoptosis in breast cancer cells [74].

Parthenolide, a natural sesquiterpene lactone isolated from the traditional Chinese medicinal plant Tanacetum parthenium, represents another electrophile that can modify biological nucleophilic molecules through Michael addition, and the highly reactive α-exo-methylene γ-butyrolactone (αMgB) group is the structure responsible for this bioactivity. For example, parthenolide increases GSH oxidation in hepatocellular carcinoma cells, and an increase in thiol oxidation results in an overall increase in lipid oxidation levels, increasing the sensitivity of cells to ferroptosis [75]. In colorectal cancer cells, parthenolide rapidly depletes GSH and protein thiols and increases intracellular ROS and endoplasmic reticulum (ER) stress, further inducing apoptosis [76]. Moreover, parthenolide likely follows the same mechanism to bind of the Sec residue of proteins. Skalska et al. identified surface Trx1 as a target of parthenolide and attributed its anti-lymphoma activity to its ability to modify the redox state of cell exofacial thiols [77]. Duan et al. reported that parthenolide selectively targets TrxR1 and TrxR2, and further induced ROS-mediated apoptosis in HeLa cells [78]. In addition, parthenolide selectively activated NOX and suppressed Trx, MnSOD, and CAT in prostate cancer cells [79]. Overall, an important component of parthenolide’s anticancer activity derives from the reaction of its lactone moiety with thiols.

Tripterygium wilfordii is a traditional Chinese medicine that was first recorded in the Compendium of Materia Medica (Ben Cao Gang Mu). Triptolide is a diterpene triepoxide extracted from Tripterygium wilfordii that has strong bioactivities. Previous studies have demonstrated that triptolide inhibits the translocation of Nrf2 from the cytoplasm to the nucleus, as well as the expression of the downstream genes NAD(P)H dehydrogenase [quinone] 1 (NQO1), CAT, and HO-1, thus abrogating Nrf2-mediated defense mechanisms against oxidative stress [80]. The suppression of Nrf2 by triptolide is closely related to its anticancer activity in several cancers, including glioma [81], lung cancer [82], and colorectal cancer [83]. Chen et al. identified PRDX2 as a direct binding target of triptolide, triptolide covalently binds to PRDX2 and inhibits its activity, inducing ROS accumulation and apoptosis in gastric cancer [84]. Triptolide induces ROS production, subsequently triggering PCD in different cancer cells. For example, triptolide induces apoptosis and autophagy in glioma cells by upregulating the ROS/JNK axis [85]. ROS generation and ERK activation are the major mechanisms by which triptolide induces apoptosis in breast cancer cells [86]. In contrast, triptolide was found to attenuate ionizing radiation-induced pulmonary fibrosis by inhibiting the axis of alveolar macrophages-NOXes-ROS-myofibroblasts axis, thereby reducing ROS [87]. This dual action appears to favor the benefits of triptolide under different pathological landscapes.

Phenolics

Plant tissues are rich in a wide variety of phenolic compounds, such as kaempferol, curcumin, resveratrol, catechins, and epigallocatechin-3-gallate (EGCG). Phenolic compounds are widely used in health care to prevent undesirable oxidation. EGCG is the most active and abundant polyphenol and has attracted considerable attention in cancer therapy because of its antioxidant activity. EGCG inhibits the process of multistage carcinogenesis and metastasis in multiple tumour models by regulating NOS, COX, Nrf2/Keap1, and other genes [88]. In contrast, phenolic compounds with anticancer potency also increase oxidative stress levels and cause cell death. ROS-mediated apoptosis plays a critical role in the anticancer effects of kaempferol in colorectal [89], glioblastoma [90], and pancreatic cancers [91]. Kaempferol is also used as a natural anticancer agent and sensitizer for lung cancer because it inhibits Nrf2 [92]. Resveratrol suppresses Notch1/PTEN/Akt signalling in ovarian cancer cells through ROS generation [93]. The anticancer activities of resveratrol are also related to manipulation of the levels of Nrf2, SOD, and CAT, for example [88]. Overall, the regulation of the redox state via pro-oxidants or antioxidants by phenolics holds significance for cancer therapy.

Curcumin is a main natural polyphenol that was first extracted from Curcuma species two centuries ago, and over 100 different clinical trials involving curcumin have been completed [94]. Curcumin increases Trx1 oxidation and subsequent apoptosis in prostate cancer cells [95]. A curcumin analogue (curcuminoid B63) has been shown to induce ROS-mediated paraptosis-like cell death by targeting TrxR1 in gastric cancer cells [96]. Indeed, the curcumin-induced inhibition of TrxR may depend on its Michael acceptor function. Specifically, curcumin has been shown to mediate the covalent modification of the active site cysteine (Cys) 496 and selenocysteine (Sec) 497 residues in TrxR, subsequently destroying TrxR reduction activity, a process that is irreversible and further leads to dramatic pro-oxidant effects, the induction of NOX activity, and the production of ROS [97]. As a reductive metabolite of curcumin, tetrahydrocurcumin lacks α,β-unsaturated ketone moieties and does not exhibit inhibitory activity against TrxR1 [98]. For TrxR2, high-affinity active sites that bind to mitocurcumin (a derivative of curcumin), including the active sites on the E, F, A, and B chains, have been identified. Mitocurcumin also regulates TrxR2 activity to generate NOX-like activity, resulting in increased ROS accumulation and apoptosis in lung cancer cells [99]. In addition, a significant decrease in the GSH/GSSG ratio was observed in lung cancer cells after treatment with curcumin [100] or its analogues [101], which led to ROS-induced apoptosis. NAC partially or completely reversed these effects, suggesting that ROS generation may be the underlying cause of curcumin-induced cell death. Curcumin regulates redox reactions, which are also mediated by metal ions. Recently, curcumin was reported to mobilize endogenous copper ions, form ROS, and further inhibit prostate cancer cells [102]. Mechanistically, curcumin reduces Cu(II) to Cu(I) and leads to the formation of H2O2, which further reacts with Cu(I) through the Fenton reaction to produce OH. Interestingly, the binding sites of curcumin to Cu also contribute to its antioxidant properties, indicating that both the antioxidant and pro-oxidant effects of curcuminoids could be attributed to the same structural moieties [103].

Erianin is a traditional Chinese medicine extracted from Dendrobium chrysotoxum Lindl and has a reputation as the “gold in medicine”. The anticancer property of erianin is mediated by multiple signalling pathways, such as the MAPK, PI3K/AKT, Wnt/β-catenin, and Nrf2 pathways [104]. Migration, invasion, and tumour angiogenesis have represented emerging anticancer targets of erianin in recent years. Redox imbalance-mediated PCD is involved in the anticancer effects of erianin. Erianin-induced apoptosis and autophagy in osteosarcoma cells are attributed to ROS induction, leading to the activation of the JNK/c-Jun signalling cascade [105]. Zhang et al. suggested that the anticancer activity of erianin in liver cancer cells was highly correlated with the regulation of ROS-mediated apoptosis; moreover, erianin-induced oxidative stress in HepG2 and SMMC-7721 cells may be related to immune function [106]. In KRAS colorectal cancer cells, erianin induces autophagy-dependent ferroptosis [107]. Erianin also induced ferroptosis in lung [108] and bladder [109] cancers, accompanied by ROS accumulation and lipid peroxidation, as well as a decrease in GSH and GPX4.

Flavonoids

Flavonoids are naturally found in a wide variety of sources, including berries, tea, wine, onions, and citrus fruits. The cytocidal effect that flavonoids possess when used in cancers is attributed primarily to their ability to evoke oxidative stress. For example, wogonin induces growth inhibition and cellular senescence in breast cancer cells by suppressing TXNRD2 by altering histone acetylation at its regulatory region [110]. Similarly, the anticancer activity of myricetin and quercetin in A549 cells may be due to the inhibition of TrxR [111]. Genistein suppresses the proliferation of leukaemia cells by decreasing the cellular redox potential (GSH/GSSG), accompanied by the downregulation of NADP-dependent isocitrate dehydrogenase [112]. Luteolin induces ferroptosis in clear cell renal cell carcinoma by inducing the Fenton reaction, GSH depletion, and lipid peroxidation [113]. Apigenin has been shown to have a cytotoxic effect through increasing the levels of ROS and lipid peroxidation (LPO) and represents a promising candidate for the treatment of cervical cancer [114]. Notably, the modulatory effect of flavonoids on the redox state is not absolute. Another perspective suggests that the role of flavonoids in cancer chemoprevention can be attributed to their capacity to quench ROS, RNS, and other radicals, as rutin acts as a tumour inhibitor and significantly reduces superoxide in HT29 cells [115].

Scutellaria baicalensis is a species of flowering plant that has been widely used as a medicine in East Asian regions. Baicalin is extracted from Scutellaria baicalensis and is a major representative of flavonoids with anticancer activities in multiple cancers. Baicalin acts as a pro-oxidant, triggering mitochondrial apoptosis in breast cancer cells through the mobilization of intracellular copper and the generation of ROS [116]. Baicalin induces ferroptosis in several cancers, including bladder [117] and gastric [118] cancers, with changes in ROS levels. Recently, baicalin was reported to inhibit the proliferation of diffuse large B-cell lymphoma cells through increasing intracellular ROS levels and GSDMD expression to induce pyroptosis [119]. Overall, baicalin promotes PCD in various cancer cells by increasing oxidative stress.

Quercetin, which bears the chemical name “3,5,7,3′,4′-pentahydroxyflavone”, is found in a wide range of berries. Quercetin enhances cellular lipid peroxidation and promotes ferroptosis in gastric cancer cells by targeting the SLC1A5/Nrf2 pathway and decreasing GSH and xCT [120]. Quercetin exposure leads to oxidative stress and ER stress in glioblastoma, accompanied by ROS generation and dysregulation of SOD [121]. Zhang et al. reported that PIG3 mediates the pro-oxidant activity of four flavonoids (kaempferol, quercetin, apigenin, and luteolin), which play crucial roles in the ROS-induced mitochondrial apoptosis in liver cancer cells [122]. Quercetin inhibits COX2 by binding to subunit A which has peroxidase activity, and the differential sensitivity of colorectal cancer cells to quercetin is attributed to COX2-dependent ROS generation [123].

Quinones

To date, quinones have become the second largest class of antitumour agents approved for clinical use in the US [124]. The most prominent chemical feature of quinones is their ability to undergo reversible oxidation–reduction and form semiquinone and oxygen radicals. In addition, the majority of the reactions of quinones with the nucleophile GSH can be characterized as reductive Michael additions, which produce GSH-hydroquinone conjugates [125]. Therefore, the cytotoxicity of quinone-containing chemotherapeutics may be mediated by the generation of semiquinone free radicals and oxyradicals [126]. Natural quinones significantly induce ROS-mediated apoptosis in pancreatic cancer cells, as observed with thymoquinone, plumbagin, and juglone [127]. Shikonin is considered a natural inducer of ROS, suppressing tumour growth and activating antitumour immunity through multiple molecular mechanisms [128]. Emodin regulates the ROS-mediated JNK and PI3K/AKT signalling pathways, consequently inducing necroptosis and inhibiting glycolysis in renal cancer cells [129]. The antioxidant GPX4 has been identified as a target of plumbagin, and its inhibition leads to PCD in glioma [130] and liver cancer cells [131]. Aloin is a natural antitumour anthraquinone glycoside. Wang et al. reported that aloin attenuates ROS generation in gastric cancer cells by downregulating NOX2 activation, consequently suppressing the phosphorylation of prosurvival signalling pathways [132].

The Rhubarb genus is commonly recognized for its edible and medicinal plants, the underground parts of which provide herbal materials. Rhein is a natural anthraquinone extracted from Rhubarb. Rhein has excellent antiproliferative effects on breast cancer cells because it induces the activation of the ROS-mediated NF-κB and p53 signalling pathways [133]. Rhein induces apoptosis in oral cancer cells by inducing ROS accumulation and inhibiting the AKT/mTOR pathway [134]. Rhein also increases the generation of ROS and activates JNK signalling in liver cancer cells [135]. Rhein has consistently demonstrated similar anticancer mechanisms that mainly involve the induction of ROS generation, which supports its promising anticancer activities.

The Lapacho tree (Handroanthus impetiginosus) is a native medicinal tree from southern Morocco and Mesoamerica. β-Lapachone is a natural naphthoquinone compound that was originally isolated from the Lapacho tree. The antitumour mechanism of β-lapachone is associated mainly with redox cycling and the production of ROS catalysed by quinone oxidoreductases. NQO1 metabolizes a two-electron reduction of β-lapachone using NADH and NADPH as electron sources, forming an unstable hydroquinone that rapidly auto-oxidizes and generates ROS, thereby inducing cell death [136]. Considering the characteristics of this natural quinone, β-lapachone has been developed to selectively target cancers with a specific increase in the quinone oxidoreductase NQO1. β-Lapachone-induced NQO1-dependent cell death has been reported in several cancers, including lung [137], breast [138], and liver cancers [139].

Other categories

NPs are diverse, and other categories, such as saponins, alkaloids, and organic acids, can also regulate intracellular redox processes. Ginseng is recognized as the "king of herbs", a plant whose roots have high medicinal value; ginsenosides are the main pharmaceutically bioactive components in ginseng. It has been reported that the anticancer effects of ginsenosides are associated with ROS regulation. In breast cancer cells, for example, ginsenoside Rh1 induces apoptosis and autophagy, which are related to the ROS-mediated Akt pathway [140]. Ginsenoside Rh2 induces increased ROS levels and activates the NF-κB pathway in colorectal cancer cells; it also induces paraptosis, a form of PCD that is characterized by cytoplasmic vacuolization [141]. Piper longum L. is a naturally existing edible and medicinal plant that grows as a perennial shrub or as an herbaceous vine. It is native to the Indo-Malaya region and is widely used in traditional medicine [142]. Piperlongumine is a biologically active alkaloid isolated from Piper longum L. that has anti-inflammatory, anticancer, antimicrobial, and immunomodulatory properties [143]. Piperlongumine inhibits esophageal carcinoma cells by inhibiting Nrf2 and promoting ROS-TXNIP-NLRP3-mediated pyroptosis [144]. Piperlongumine directly targets the Sec498 residue of TrxR1 to inhibit gastric cancer cells [145]. Additionally, it selectively induces ROS accumulation and cell death in cancer cells but not in normal cells. For example, piperlongumine reduces GSH and induces ROS accumulation in glioblastoma multiforme cells, which further activates the JNK and p38 pathways [146]. Piperlongumine increases the intracellular ROS levels sufficiently to cause lethal oxidative stress in breast cancer cells by inhibiting the antioxidant enzymes CAT, Trx1, and Prx2 [147]. Indeed, it has been widely reported that NPs exhibit significant anticancer effects in multiple cancers by regulating the redox microenvironment. Here, we summarize the related mechanisms that involve genes, proteins, transcription factors, signalling pathways, and PCD, with more detailed content presented in Table 1.

Table 1 Anticancer mechanisms of representative NPs involving redox regulation in cancers

Exploration: NPs enhance sensitivity and overcome resistance by regulating the redox balance

The insensitivity and resistance that evolve in cells during treatment have become increasingly prominent issues affecting therapeutic efficacy. Given the excellent anticancer effects of NPs targeting redox processes, there is increasing interest in combining NPs with conventional therapies to enhance sensitivity or reverse resistance. In recent years, broad reports have provided evidence that NPs are attractive redox-regulating adjuvants for cancer treatment (Fig. 4).

Fig. 4
figure 4

The synergistic mechanisms of the use of NPs as redox-regulating adjuvants in cancers. The combination of NPs with traditional therapies effectively alters the levels of ROS and redox-related targets in various cancer cells

Targeting the GSH/GSSG system: NPs with Michael acceptor functionality

As mentioned previously, the anticancer effect of oridonin, which targets GSH through the Michael addition reaction, has been reported. Therefore, targeting the GSH/GSSG system may represent an advantageous approach to enhancing oridonin sensitivity and overcoming drug resistance in cancer cells. Platinum compounds are among the most widely used chemotherapy drugs for various types of cancers; however, multidrug resistance is the main barrier preventing their application. It has been revealed that enhanced drug detoxification by GSH or GSH-related moieties is a major mechanism of platinum resistance [148, 149]. Yang et al. reported the synergistic anticancer effects of oridonin and cisplatin against esophageal squamous carcinoma cells, which are likely driven by the inhibition of GSH generation and increased ROS generation, ultimately inducing apoptosis [150]. Oridonin has also been shown to enhance the cytotoxic effect of 5-fluorouracil (5-FU) on renal cancer cells via GSH depletion and ROS formation, further leading to intracellular apoptosis and necroptosis, which are associated with the activation of MAPK and other pathways [151]; moreover, supplementation with GSH or NAC abolished this toxic effect of oridonin, supporting the synergistic effect of GSH/ROS. Tumour necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anticancer agent that induces cytotoxicity triggered by interactions with death receptors in cancer cells. Plumbagin is a bicyclic naphthoquinone produced in the roots of Plumabago zeylanica L., and its cytotoxic properties are related to its quinone core. Plumbagin is considered to be a potent electrophile that reacts with the thiol group of GSH [152]. Plumbagin enhances TRAIL-induced apoptosis of human leukaemia cells by depleting GSH and increasing the ROS-mediated DRs of TRAIL, exhibiting synergistic effects with TRAIL treatment [153]. Similarly, embelin is a naturally occurring benzoquinone compound that enhances TRAIL sensitivity by suppressing critical antioxidants and increasing ROS accumulation through the consumption of GSH and the downregulation of SOD1 in inflammatory breast cancer cells [154]. Emodin is a natural anthraquinone derivative found in Rhubarb that has significant pharmacological effects and clinical applications. Emodin enhances the chemosensitivity of bladder cancer cells to cisplatin by depleting GSH, decreasing GSH-cisplatin conjugates, and elevating ROS levels [155]. Quercetin is another quinone with Michael acceptor properties. It has been reported that GSH levels decrease when breast cancer cells are treated with high concentrations of quercetin, and quercetin combined with tamoxifen synergistically inhibits cell viability [156]. Overall, these observations suggest that compounds with Michael acceptor functionalities are potential sensitizers for cancer treatment [69] (Fig. 5A).

Fig. 5
figure 5

Representative combinations of NPs and chemotherapeutic agents that enhance sensitivity or reverse drug resistance. A NPs with Michael acceptor functionality represent potential sensitizers for cancer treatment by targeting GSH. B The presence of electrophilic centres may support the activities of NPs for enhancing sensitivity and reversing resistance. C Regulating the NADP+/NADPH system and increasing the levels of NOXs in cancer cells. D NPs targeting Nrf2 or other redox indicators to reverse drug resistance

Targeting Trx/TrxR: a classic pathway for enhancing sensitivity and reversing resistance

High Trx/TrxR levels are important components of the resistant phenotype in various cancers [157,158,159]. The Sec residue of TrxR seems to be a potent binding site for several NPs. As mentioned before, piperlongumine binds or inhibits the activity of antioxidant enzymes, including Trx1, TrxR1, and PRDX2, in cancer cells [145, 147]. Mechanistically, the α,β-unsaturated ketone moiety in piperlongumine, which serves as a Michael addition acceptor, can bind the C-terminal Sec residue of TrxR [160]. Recently, Zhang et al. reported that the combination of piperlongumine and oxaliplatin significantly inhibited TrxR1 activity in gastric cancer cells, leading to increased ROS production and oxidative stress, suggesting that piperlongumine synergistically potentiates the antitumour effect of oxaliplatin both in vitro and in vivo [161]. Shikonin is a natural naphthoquinone that modifies the Sec498 residue of TrxR1 to fully inhibit its antioxidant activity [162]. Shikonin targets TrxR and induces oxidative stress to promote apoptosis in gefitinib-resistant non-small cell lung cancer cells, with decreases in NADPH and the GSH/GSSG ratio [163]. Similarly, the combination of plumbagin and BAY-876 (a GLUT1 inhibitor) showed synergistic cytotoxicity in lung cancer cells by inhibiting TrxR and ROS generation. Specifically, plumbagin contains an α,β-unsaturated carbonyl functionality that serves as a Michael acceptor, modifies the Sec498 site of TrxR1, and promotes ROS production [164]. Quercetin pretreatment has been shown to act as a pro-oxidant in cisplatin-resistant ovarian adenocarcinoma cells, inducing ROS production by effectively reducing the levels of the Trx/TrxR antioxidant system (Trx1, Trx2, and TrxR1) and downregulating the mTOR/STAT3 pathways, which synergistically potentiate cisplatin cytotoxicity [165]. Zhu et al. synthesized a derivative of oleanolic acid, olean-28,13b-olide 2, which reversed the cisplatin resistance of lung cancer cells by inhibiting TrxR, regulating other pathways, and enhancing cisplatin-induced ROS accumulation [166]. Considering that the inhibition of TrxR by electrophiles usually occurs through Michael addition, the electrophilic centres in NPs likely support their anticancer activity and synergistic effects by targeting the Trx/TrxR system [167] (Fig. 5B).

Targeting NADP+/NADPH: increased expression of the pro-oxidant enzyme NOX

Currently, several NPs have been shown to enhance the sensitivity or reverse the resistance of cancer cells by regulating NADPH or NOXs (Fig. 5C). Epidermal growth factor receptor (EGFR) tyrosine kinase inhibitor (TKI) resistance is caused by an additional T790M mutation in EGFR. Sanguinarine is a naturally derived isoquinoline alkaloid from Sanguinaria canadensis with various bioactivities, including anti-inflammatory, anticancer, and neuroprotective effects [168]. In gefitinib-resistant non-small cell lung cancer cells, sanguinarine-mediated M790 oxidation occurs through the upregulation of NOX3 and subsequent H2O2 production. NOX3 oxidizes NADPH, causing severe NADPH depletion and an increase in the NADP+/NADPH ratio, ultimately triggering EGFR overoxidation and apoptosis [169]. The ginsenoside Ro-triggered accumulation of superoxides inhibits autophagy and sensitizes esophageal cancer cells to 5-FU-mediated cell death. As important sources of superoxides, both NOX2 and NCF1 were shown to be upregulated by ginsenoside Ro treatment in a dose-dependent manner, contributing to the synergistic anticancer effect of ginsenoside Ro [170]. Baicalin enhances the efficacy of 5-FU in gastric cancer cells by inducing ferroptosis. In this case, alterations occurred within the cellular redox microenvironment, marked by elevated cellular oxidation, increased levels of the oxidase NOX1, and decreased expression of the antioxidant enzyme GPX4 [118]. In addition to forming the antioxidant couple, another role of NADPH is as a substrate for NOX to promote the production of ROS. Therefore, the NADPH concentration could also be increased by NP treatment in certain cases. For example, combination treatment with ursolic acid and oxaliplatin increased NADPH levels, a resource for ROS production, in colorectal cancer cells. Elevated ROS levels markedly reduce the expression of drug resistance genes, including permeability glycoprotein, multidrug resistance-associated protein, and breast cancer resistance protein [171]. In contrast, NPs act as sensitizers by downregulating NOX levels in certain cases. It has been reported that emodin combined with cisplatin downregulates the expression of NOX, along with a decrease in ROS, which is considered to inhibit the expression of drug resistance genes and enhance chemosensitivity in endometrial cancer cells [172].

Targeting the Nrf2/HO-1 axis and other redox indicators: a potent pathway for reversing drug resistance

High Nrf2 expression is associated with the resistance of several cancers to chemotherapy [173]. The inhibition of Nrf2 by wogonin contributes to its ability to overcome drug resistance in multiple cancers, including liver cancer [174], breast cancer [175], and head and neck cancers [176]. For example, wogonin inhibits drug detoxification by downregulating Nrf2 and its effector GSTP1, thus reversing cisplatin resistance in HN4-cisR and HN9-cisR cells. Suppression of Nrf2/HO-1 by wogonin reverses doxorubicin resistance in MCF-7/DOX cells. In cisplatin-resistant non-small cell lung cancer cells, ginsenoside Rd was found to restore sensitivity to chemotherapy drugs by suppressing Nrf2, which was accompanied by a reduction in the expression of the downstream genes NQO1 and HO-1, as well as a decrease in the levels of the drug resistance genes MDR1 and MRP1 [177]. Berberine resensitizes breast cancer cells to lapatinib by suppressing the transcriptional activation of Nrf2 and increasing ROS levels [178]. Quercetin suppresses the expression of critical antioxidant enzymes, such as Nrf2/HO-1, GPX, SOD, and CAT, in both 5-FU-resistant colorectal cancer cells [179] and cisplatin-resistant ovarian cancer cells [180]. These mechanisms are considered to contribute to overcoming the resistance of colorectal cancer cells to quercetin. Parthenolide prevents the resistance of breast cancer cells to doxorubicin and mitoxantrone through the downregulation of Nrf2, CAT, and MnSOD [181]. NF-κB is not only associated with redox regulation but is also involved in the development of drug resistance in ovarian cancer cells with increased activity. Triptolide increases cellular ROS production by inhibiting complex I of the mitochondrial respiratory chain, and ROS further mediate the inactivation of the NF-κB survival pathway and cell apoptosis, thus reversing cisplatin resistance in ovarian cancer [182]. Overall, targeting these antioxidant indicators is a potent pathway for reversing the drug resistance of multiple cancers (Fig. 5D).

The regulation of ROS: a bidirectional regulatory effect

ROS are important regulatory factors that determine cell survival and death. It is reasonable to speculate that the signalling pathways discussed earlier may lead to changes in the levels of intracellular ROS, which act as signalling molecules and mediate PCD in cancer cells, consequently supporting the anticancer properties of NPs. Several studies have reported the impact of NPs on tumour ROS; more importantly, future research is necessary to explore the underlying pathways and specific targets involved in these studies. Although most of the mechanisms reviewed above suggest that NPs generally positively regulate ROS in cancers, the bidirectional modulation of the redox microenvironment by NPs cannot be ignored (Fig. 6A, B). For example, quercetin has been shown to enhance the antitumour activity of doxorubicin combined with cyclophosphamide in breast cancer cells by inhibiting ROS accumulation and the ERK1/2 pathway [183]. A reduction in ROS has also been reported as the mechanism by which quercetin overcomes 5-FU resistance in colorectal cancer cells [179]. The regulation of ROS by quercetin may depend on the drug dosage, given that tamoxifen-induced ROS generation was found to be suppressed by low-concentration quercetin, whereas high concentrations of quercetin synergized with tamoxifen to increase the production of ROS [156].

Fig. 6
figure 6

NPs play a dual role in the regulation of redox reactions in different cases. A Cancer cells typically exhibit abnormal redox states characterized by elevated levels of oxidative stress and increased antioxidant systems, which contribute to "self-adaptation" and resistance to treatment toxicity. B NPs typically act selectively as pro-oxidants in cancer cells to increase sensitivity and reverse drug resistance; however, they can exert synergistic effects by inhibiting oxidative stress. C NPs decrease the adverse effects of cancer therapy in normal cells by serving as antioxidants and protectors

Resveratrol is a phenolic compound that is widely present in various dietary compounds, such as mulberries, peanuts, blueberries, and grapes. It has been reported that resveratrol synergizes with 5-FU to combat colorectal cancer, increasing the ROS levels and decreasing the CAT and GPX levels [184]. The combination of resveratrol and platinum (IV) complex induces increased levels of O2• −, H2O2, and NO2, further amplifying the pro-oxidative effects of the platinum (IV) complex on both breast adenocarcinoma and choriocarcinoma cells [185]. Resveratrol enhances the antitumour effects of temozolomide in glioblastoma cells via the ROS-dependent AMPK-TSC-mTOR signalling pathway [186]. However, the synergistic effect also likely depends on the antioxidant properties of resveratrol. Resveratrol sensitizes glioma cells to temozolomide-induced cell death synergistically through the downregulation of protective autophagy, leading to a decrease in ROS [187]. EGCG is another major form of dietary polyphenol, and its combined use with EGFR-TKIs has been shown to significantly reverse the Warburg effect, leading to increased OXPHOS and ROS and overcoming drug resistance in non-small cell lung cancer [188]. Wei et al. also showed that EGCG can sensitize renal cell carcinoma cells to TRAIL-induced apoptosis by downregulating c-FLIP via a ROS-dependent pathway [189]. In contrast, EGCG in conjunction with irinotecan has been shown to suppress the production of ROS in colorectal cancer, where the altered redox homeostasis, along with the regulation of GRP78, collectively enhances the chemosensitivity of colorectal cancer cells to irinotecan [190]. The antioxidant properties of EGCG also contribute to reversing multidrug resistance in oral cancer [191]. Overall, NPs enhance cancer sensitivity or reverse resistance by modulating redox balance, a fact that has been well characterized (Table 2).

Table 2 Representative NPs enhance sensitivity or reverse drug resistance in cancers by regulating redox balance

Detoxification effect of NPs: focus on antioxidant properties

The burden of cancer treatment-induced toxicity and adverse effects remains significant, including, but not limited to, various organ toxicities, weight loss, and systemic reactions. Many studies have demonstrated that NPs protect normal cells by attenuating chemotherapy/radiation-induced oxidative stress (Fig. 6C). Another advantage of NPs as adjuncts in cancer therapy is that they do not interfere with the efficacy of existing anticancer agents [192]. Although oxidative stress is one of the anticancer pathways of platinum compounds, it also plays a key role in the systemic toxicities induced by platinum [193]. Honokiol is a natural polyphenolic compound extracted from Magnolia grandiflora and has multiple anticancer effects. Systemic application of honokiol prevents cisplatin-induced hearing loss by ROS detoxification without compromising its antitumour effect [192]. Nanoparticulated honokiol attenuates cisplatin-induced nephrotoxicity by maintaining mitochondrial antioxidant capacity [194]. Tanshinone is a natural compound derived from the traditional Chinese medicine Salvia miltiorrhiza and has various bioactivities. Tanshinone IIA reduces oxaliplatin-induced neurotoxicity by reducing ROS levels [195]. Curcumin adjuvant administration to cisplatin therapy exerts antioxidant and protective effects by upregulating the endogenous antioxidant defense system (Nrf-2/HO-1) in cochlear cells [196]. The combination of curcumin and α-tocopherol protects against cisplatin-induced hepatotoxicity by reducing NOX and oxidative markers (ROS and MDA) and increasing CAT [197]. In addition, curcumin pretreatment significantly increases Nrf2 in the nucleus and decreases the expression of Keap1, as well as reverses the doxorubicin-induced reduction of HO-1 and NQO1, which provides a rational mechanism against doxorubicin-induced neurotoxicity [198]. Mitochondrial dysfunction-induced oxidative stress represents a major cause of doxorubicin-induced cardiotoxicity. Oridonin shows a cardiac protective effect and improves doxorubicin-induced oxidative stress. The pretreatment with oridonin has been shown to significantly elevate the activities of antioxidant defense components such as SOD, Mn-SOD, HO-1, and GPX, while reducing NOX subunit p47phox activation levels [199]. Another benefit of combining oridonin with chemotherapeutic agents is the reduction in the doses of chemodrugs, which serves to decrease the incidence of adverse effects [200]. Similarly, berberine alleviates doxorubicin-induced cardiotoxicity by suppressing oxidative stress via activating the Nrf2 pathway [201]. A recent study has demonstrated that kaempferol offers protection against doxorubicin-induced nephrotoxicity by suppressing ROS/ASK1-mediated MAPK signaling and attenuating oxidative stress [202]. Previous reviews have explored phytochemicals with radioprotective effects, primarily including phenols, flavonoids, and polysaccharides [203]. For example, ferulic acid pretreatment has been shown to protect peripheral blood mononuclear cells from radiation damage by significantly preventing radiation-induced ROS generation, restoring GSH, and increasing NF-κB and Nrf2 levels within the nuclei [204]. The testis, as one of the most radiosensitive organs, can be significantly impaired by even low doses of radiation. Kim et al. reported that genistein has a protective effect against radiation-induced testicular injury, primarily through the mitigation of ROS [205]. Overall, the anti-oxidative pathway may serve as a major pathway of NPs in normal cells within the context of chemotherapy/radiation protection.

Contemporary research: potential challenges and promising future

Although NPs are widely available and affordable, they still face challenges from the laboratory to the clinic due to limitations in their properties, effectiveness, safety, and translation. To date, numerous improvements in NP-based properties, derivative preparations, and novel delivery systems have been widely reported as advanced approaches to address these challenges and increase their efficiency. With progress in product conversion and clinical trials (Fig. 7), NPs are well developed and advancing, providing promising prospects for cancer treatment strategies.

Fig. 7
figure 7

Advances in clinical trials of anticancer drugs based on NPs over the past few decades. Clinical trials of anticancer drugs based on NPs are emerging with maturity, whether they are used alone, in combination with sensitizers, or to reduce adverse effects, indicating promising prospects for the benefit of patients

Overcoming limitations: novel redox-sensitive materials

Most NPs have common characteristics such as poor solubility, low stability, and low bioavailability [206]. An innovative approach to overcome these limitations is the multifunctional design of structural modifications and drug delivery systems based on the structure‒activity relationship (SAR). For example, the structural modification of neohesperidin using an immobilized lipase improves lipophilicity, benefiting its action on the cancer cell wall and enhancing bioavailability in vivo [207]. The water solubility of β-elemene could be improved by introducing hydrophilic moieties like hydroxyl and amino groups [208]. A novel nanodrug delivery system was constructed to coload paclitaxel and triptolide, with favourable stability and biocompatibility, enhancing ferroptosis in lung cancer cells via ROS accumulation [209]. Zhen et al. prepared biomineralized oxidized EGCG-molybdenum ion-coordinated nanoparticles with good solubility and long-term safety that reacted with GSH via Michael addition to form aggregates, which further inhibited cancer progression [210]. Yan et al. prepared a nano-metal-organic framework (MOF)-encapsulated dihydroartemisinin, with better cytocompatibility and anticancer activity, and induced apoptosis in ovarian cancer cells by inhibiting ROS [211]. Cui et al. synthesized hawthorn fruit extract-mediated selenium nanoparticles, which were stable and inhibited liver cancer cells by inducing oxidative stress and mitochondrial dysfunction [212]. Collectively, these novel redox-sensitive materials may induce a cascade amplification of oxidative stress and exert significant anticancer activity (Fig. 8).

Fig. 8
figure 8

Novel redox-sensitive materials enhance the utilization efficiency of NPs. Considering the abnormal redox microenvironment in cancer cells, novel redox-sensitive materials (structural modifications, derivatives, and nanodelivery systems) based on NPs and drugs serve as innovative approaches to overcome the limitations of NPs and may further induce cascade amplification of oxidative stress to exert significant anticancer activity

Preclinical strategies: improving selectivity and evaluating safety

Although NPs can attenuate the adverse effects of conventional therapies, we cannot hold the preconceived notion that they are absolutely safe. Several approaches could avoid undesirable effects and achieve the "high efficacy and low toxicity" of NPs by increasing selectivity and reducing off-target effects. The cisplatin and curcumin coloaded liposome system extends the drug duration and promotes drug accumulation in tumours. The nano-liposomes generate more ROS compared with the single drug agent [213]. Liposomes loaded with paclitaxel-doxorubicin respond to stimulation of the tumour microenvironment with high levels of ROS/GSH to release drugs [214]. Advanced formulations of resveratrol increase the accumulation of loaded drugs in cancer areas and prevent off-target drug release [215]. A ROS-responsive artesunate complex carrier was prepared, with hyaluronic acid in this carrier enabling deep tumour penetration and selective drug release [216]. In addition, some biological probes based on NPs have been prepared and can be used for target identification and improving selectivity. Efforts are needed to evaluate the safety of NPs in vitro and in vivo before any pharmaceutical administration, with a focus on distribution, metabolism, and organ toxicity. The clinical application of celastrol is limited due to severe side effects. Compound 19–048 derived from celastrol significantly reduces side effects according to target identification and structural modifications [63]. The identification, response, and management of potential toxicity and side effects are critical challenges for the clinical application of NPs, which still have a long way to go.

Product conversion: technology and science

The transformation of achievements and standardized production is an important step for bringing a natural compound from the laboratory to the clinic. Emerging technologies and advances in science provide opportunities for the discovery, production, and engineering of NPs. Traditional NP-based research starts with screening, identifying, and extracting, which is a costly process. Recently, analytical techniques, genome mining and engineering, and microbial culturing have been identified as advanced technologies for discovering NPs [14]. Most NPs require improvements before they can be used in the clinic; rather than being used directly, biotechnology and engineering will be highly important in this context. Plant cell suspension culture is emerging as a viable technology to overcome quality and quantity challenges during the production of NPs [217]. Li et al. reported the use of protein engineering for improving and diversifying NP synthesis [218]. Additional challenges with production that need to be resolved include purity and titer, quality control, standardization, and sustainability. Science and technology progress in other important research areas, such as computational tools [219], database platforms [220], and multiomics [221], also significantly influence NP-based therapies and offer great potential for the development of NPs.

Promising progression: clinical trials and patents

Given the promising aspects of anticancer therapy, several NPs and their analogs or derivatives are emerging for use in clinical trials or patents (Table 3). Elemene, a major ingredient of the traditional Chinese medicine Curcumae Rhizoma, has been confirmed by our team to have inhibitory activity against various cancers [222, 223]. We have disclosed several patents for elemene oral microemulsion (US20120322892A1), injectable solution (CN101306181A), and sustained-release tablets (US20130059922A1). The synergistic treatment of elemene plus TKIs for EGFR-mutated advanced non-small-cell lung cancer has entered Phase IV clinical trials (NCT04401059). Clinical trials of genistein in combination therapy have been conducted in breast cancer (NCT00244933) and lung cancer (NCT01628471). Recently, quercetin combined with dasatinib has been approved to enter Phase II clinical trials (NCT06355037) for reversing chemotherapy resistance in breast cancer. In addition, the anticancer applications of some classic NPs such as paclitaxel, vincristine, and camptothecin are already becoming increasingly mature. Collectively, the application of NPs into clinical practice is a tantalizing prospect that may facilitate patient benefits.

Table 3 Advances in clinical trials of anticancer drugs based on NPs have occurred over the past few decades

Conclusions and perspectives

Oxidative stress is a “double-edged sword” in cancer biology and treatment. Although the antioxidant capacity of NPs endows them with the ability to favourably modulate some pathological processes and reduce the side effects of chemotherapeutic drugs, caution is warranted when discussing the mechanisms and the use of NPs as sensitizers. In contrast to their antioxidant activity, the synergistic effect of NPs, which is mediated by increasing oxidative stress, has been documented in many cancers, and such a phenomenon is not uncommon in anticancer drugs. A reasonable explanation for this paradoxical phenomenon is that an antioxidant in one system is not necessarily an antioxidant in all systems. NPs may exert distinct effects on the regulation of redox reactions in different types of cells, i.e., nontumour cells undergoing stressful conditions versus cancer cells. In the former, NPs exert a protective effect as antioxidants and detoxifiers, whereas in cancer cells, NPs are mostly converted to mediate pro-oxidant activity, promoting cell death and overcoming therapeutic resistance. In other words, NPs appear to be flexible and display relatively selective benefits in different types of cells. Several mechanisms seem to explain this phenomenon on the basis of the preceding discussion.

The heterogeneity in the redox state between cancer and normal cells provides a biological basis for regulating redox reactions as a strategy to selectively kill cancer cells with NPs. It has been well-established that cancer cells exhibit higher levels of oxidative stress compared with normal cells. Excessive oxidative stress not only is an inducing factor for carcinogenesis but also renders cancer cells more vulnerable to oxidative damage because they survive with a heightened basal level of ROS-mediated signalling, which is required for high metabolism, enhanced energy, and rapid proliferation [13, 224, 225]. Under these conditions, further exposure to threshold-excessive ROS overwhelmingly pushes cancer cells toward peroxidative damage and cell death. In contrast, normal cells may tolerate exogenous intervention better because of their low baseline levels of oxidant signalling and normal metabolic modulation [224, 225]. For example, flavonoids can scavenge low levels of ROS in normal cells. However, high and sensitive levels of ROS in cancer cells are difficult to scavenge with flavonoids. Furthermore, the mitochondrial membrane disruption and decrease in the membrane potential caused by flavonoids are accompanied by a pro-oxidative response [226]. In addition, considering that part of the anticancer mechanism of NPs occurs through exofacial protein modification, the extracellular redox state may also contribute to this selective effect. There is evidence that in contrast to the oxidative extracellular redox state of normal cells, cancer cells exist in a reduced extracellular environment [227]. For example, it is possible that the exofacial thiol targets of parthenolide are reduced on cancer cells and, thus, able to interact with parthenolide but are oxidized on normal cells and, thus, unavailable for interaction with parthenolide [77].

Another possibility links the different effects of NPs to redox-related targets from the perspective of molecular biology. For example, the antioxidant and pro-oxidant effects of quercetin depend on the availability of reduced intracellular GSH. In the presence of high GSH concentrations in normal cells, the reversibility between oxidized quercetin and GSH ensures protection against quercetin toxicity [228]. Parthenolide selectively activates NOX and increases oxidative stress in prostate cancer cells; however, parthenolide may become trapped in normal cells by sulfa-Michael addition under the catalysis of GSTs, rendering them more tolerant [79, 229]. Xu et al. identified Keap1 as a determinant controlling the differential effects of parthenolide on cancer and normal cells [230]. Parthenolide susceptibility is also determined by CAT activity [231]. In summary, this speculation is plausible, and we hypothesize that the antioxidant mechanism of NPs in normal cells may be accessed by stimulating the production of more antioxidants, following a "consumption–replenishment" model. However, this concept seems insufficient to explain all cases because some cancer cells, as mentioned before, particularly drug-resistant cells, also exhibit significantly increased antioxidant defenses. For these cells, targeting the antioxidant system is a strong strategy for sensitization and reversing resistance.

Importantly, the anticancer activity of the NPs is not immutable, with several pieces of evidence indicating that the anticancer effects of the NPs involve both their pro-oxidant and antioxidant activities. Specifically, the effects of NPs may be influenced by different concentrations, durations, test conditions, and cancer types, which may explain their biphasic effects. As mentioned previously, quercetin at low concentrations has antioxidant activity, whereas high concentrations of it synergize to increase the production of ROS [111]. Another hypothesis for the biphasic effect is that a burst of ROS is not always beneficial for inducing cytotoxicity-PCD but may also mediate protective autophagy in cancer cells. For example, ROS-mediated autophagy induction by honokiol in prostate cancer cells is cytoprotective [232]. In this context, the use of antioxidants is promising. As reported by Lin et al., the combination of resveratrol and temozolomide affects malignant glioma by inhibiting ROS/ERK-mediated autophagy [187].

In summary, it has been widely confirmed that NPs serve as adjuvants for enhancing drug sensitivity and overcoming resistance, as well as for reducing adverse effects by regulating the redox microenvironment. NPs exhibit differential, selective, and bidirectional effects on tumour redox states. The underlying mechanisms of this phenomenon appear complex, multifaceted, and interconnected, involving different cellular contexts, molecular regulations, pathological landscapes, and conditions of action. Whether there are critical thresholds between these effects and how to identify and delineate them to ensure the optimal effect of NPs as intended in different diseases remain unresolved issues. Future research is expected to explore more specific targets and mechanisms of NPs, critically identifying the roles of the targets in redox balance and the differences between normal and malignant environments. Currently, research progress on novel redox-sensitive NP materials is inspiring, with some NPs already being developed for clinical trials. These findings support the promising prospects of NPs, which are expected to provide clinical and patient benefits in the future.

Availability of data and materials

Not applicable.

Abbreviations

NPs:

Natural products

ROS:

Reactive oxygen species

RNS:

Reactive nitrogen species

H2O2 :

Hydrogen peroxide

NO:

Nitric oxide

GSH:

Glutathione

GSSG:

Glutathione disulfide

Trx:

Thioredoxin

TrxR:

Thioredoxin reductase

NADPH:

Nicotinamide adenine dinucleotide phosphate

OXPHOX:

Oxidative phosphorylation

NOX:

NADPH oxidase

LOX:

Lipoxygenase

COX:

Cyclooxygenase

NOS:

Nitric oxide synthase

GPX:

Glutathione peroxidase

Grx:

Glutaredoxin

GST:

Glutathione S-transferase

NAC:

N-acetylcysteine

Prx:

Peroxiredoxin

NF-κB:

Nuclear factor kappa B

Nrf2:

Nuclear factor erythroid 2-related factor 2

Ref-1:

Redox factor-1

TXNIP:

Thioredoxin-interacting protein

CAT:

Catalase

SOD:

Superoxide dismutase

HO-1:

Haem oxygenase-1

PCD:

Programmed cell death

GLS:

Glutaminase

ER:

Endoplasmic reticulum

NQO1:

NAD(P)H dehydrogenase [quinone] 1

EGCG:

Epigallocatechin-3-gallate

Cys:

Cysteine

Sec:

Selenocysteine

LPO:

Lipid peroxidation

5-FU:

5-Fluorouracil

TRAIL:

Tumour necrosis factor-related apoptosis-inducing ligand

EGFR:

Epidermal growth factor receptor

TKI:

Tyrosine kinase inhibitor

MOF:

Metal–organic framework

References

  1. Jones DP. Redox organization of living systems. Free Radic Biol Med. 2024;217:179–89.

    Article  PubMed  CAS  Google Scholar 

  2. Sies H, Jones DP. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat Rev Mol Cell Biol. 2020;21(7):363–83.

    Article  PubMed  CAS  Google Scholar 

  3. Chen XM, Chen HS, Xu MJ, Shen JG. Targeting reactive nitrogen species: a promising therapeutic strategy for cerebral ischemia-reperfusion injury. Acta Pharmacol Sin. 2013;34(1):67–77.

    Article  PubMed  CAS  Google Scholar 

  4. Forman HJ, Zhang H, Rinna A. Glutathione: overview of its protective roles, measurement, and biosynthesis. Mol Aspects Med. 2009;30(1–2):1–12.

    Article  PubMed  CAS  Google Scholar 

  5. Jastrzab A, Skrzydlewska E. Thioredoxin-dependent system. Application of inhibitors. J Enzyme Inhib Med Chem. 2021;36(1):362–71.

    Article  PubMed  CAS  Google Scholar 

  6. Jomova K, Raptova R, Alomar SY, Alwasel SH, Nepovimova E, Kuca K, et al. Reactive oxygen species, toxicity, oxidative stress, and antioxidants: chronic diseases and aging. Arch Toxicol. 2023;97(10):2499–574.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Zhao Y, Hu X, Liu Y, Dong S, Wen Z, He W, et al. ROS signaling under metabolic stress: cross-talk between AMPK and AKT pathway. Mol Cancer. 2017;16(1):79.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Poillet-Perez L, Despouy G, Delage-Mourroux R, Boyer-Guittaut M. Interplay between ROS and autophagy in cancer cells, from tumor initiation to cancer therapy. Redox Biol. 2015;4:184–92.

    Article  PubMed  CAS  Google Scholar 

  9. Juan CA, de la Perez Lastra JM, Plou FJ, Perez-Lebena E. The chemistry of reactive oxygen species (ROS) revisited: outlining their role in biological macromolecules (DNA, lipids and proteins) and induced pathologies. Int J Mol Sci. 2021;22(9):4642.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Klaunig JE. Oxidative stress and cancer. Curr Pharm Des. 2018;24(40):4771–8.

    Article  PubMed  CAS  Google Scholar 

  11. Cheung EC, Vousden KH. The role of ROS in tumour development and progression. Nat Rev Cancer. 2022;22(5):280–97.

    Article  PubMed  CAS  Google Scholar 

  12. Cui Q, Wang JQ, Assaraf YG, Ren L, Gupta P, Wei L, et al. Modulating ROS to overcome multidrug resistance in cancer. Drug Resist Updat. 2018;41:1–25.

    Article  PubMed  Google Scholar 

  13. Jiang H, Zuo J, Li B, Chen R, Luo K, Xiang X, et al. Drug-induced oxidative stress in cancer treatments: angel or devil? Redox Biol. 2023;63: 102754.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Atanasov AG, Zotchev SB, Dirsch VM. International natural product sciences T, Supuran CT. Natural products in drug discovery: advances and opportunities. Nat Rev Drug Discov. 2021;20(3):200–16.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Yang Y, Karakhanova S, Hartwig W, D’Haese JG, Philippov PP, Werner J, et al. Mitochondria and mitochondrial ROS in cancer: novel targets for anticancer therapy. J Cell Physiol. 2016;231(12):2570–81.

    Article  PubMed  CAS  Google Scholar 

  16. Nakamura H, Takada K. Reactive oxygen species in cancer: current findings and future directions. Cancer Sci. 2021;112(10):3945–52.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Lu J, Tan M, Cai Q. The Warburg effect in tumor progression: mitochondrial oxidative metabolism as an anti-metastasis mechanism. Cancer Lett. 2015;356(2):156–64.

    Article  PubMed  CAS  Google Scholar 

  18. Nagano O, Okazaki S, Saya H. Redox regulation in stem-like cancer cells by CD44 variant isoforms. Oncogene. 2013;32(44):5191–8.

    Article  PubMed  CAS  Google Scholar 

  19. Wang J, Sun D, Huang L, Wang S, Jin Y. Targeting reactive oxygen species capacity of tumor cells with repurposed drug as an anticancer therapy. Oxid Med Cell Longev. 2021;2021:8532940.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Reis J, Gorgulla C, Massari M, Marchese S, Valente S, Noce B, et al. Targeting ROS production through inhibition of NADPH oxidases. Nat Chem Biol. 2023;19(12):1540–50.

    Article  PubMed  CAS  Google Scholar 

  21. Chai YC, Mieyal JJ. Glutathione and glutaredoxin-key players in cellular redox homeostasis and signaling. Antioxidants. 2023;12(8):1553.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Wu G, Fang YZ, Yang S, Lupton JR, Turner ND. Glutathione metabolism and its implications for health. J Nutr. 2004;134(3):489–92.

    Article  PubMed  CAS  Google Scholar 

  23. Cassier-Chauvat C, Marceau F, Farci S, Ouchane S, Chauvat F. The glutathione system: a journey from cyanobacteria to higher eukaryotes. Antioxidants. 2023;12(6):1199.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Starke DW, Chock PB, Mieyal JJ. Glutathione-thiyl radical scavenging and transferase properties of human glutaredoxin (thioltransferase). Potential role in redox signal transduction. J Biol Chem. 2003;278(17):14607–13.

    Article  PubMed  CAS  Google Scholar 

  25. Laborde E. Glutathione transferases as mediators of signaling pathways involved in cell proliferation and cell death. Cell Death Differ. 2010;17(9):1373–80.

    Article  PubMed  CAS  Google Scholar 

  26. Pedre B, Barayeu U, Ezerina D, Dick TP. The mechanism of action of N-acetylcysteine (NAC): the emerging role of H(2)S and sulfane sulfur species. Pharmacol Ther. 2021;228: 107916.

    Article  PubMed  CAS  Google Scholar 

  27. Hasan AA, Kalinina E, Tatarskiy V, Shtil A. The thioredoxin system of mammalian cells and its modulators. Biomedicines. 2022;10(7):1757.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Karlenius TC, Tonissen KF. Thioredoxin and cancer: a role for thioredoxin in all states of tumor oxygenation. Cancers. 2010;2(2):209–32.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Yang B, Lin Y, Huang Y, Shen YQ, Chen Q. Thioredoxin (Trx): a redox target and modulator of cellular senescence and aging-related diseases. Redox Biol. 2024;70: 103032.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Fritz-Wolf K, Kehr S, Stumpf M, Rahlfs S, Becker K. Crystal structure of the human thioredoxin reductase-thioredoxin complex. Nat Commun. 2011;2:383.

    Article  PubMed  Google Scholar 

  31. Choi EH, Park SJ. TXNIP: a key protein in the cellular stress response pathway and a potential therapeutic target. Exp Mol Med. 2023;55(7):1348–56.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Ju HQ, Lin JF, Tian T, Xie D, Xu RH. NADPH homeostasis in cancer: functions, mechanisms and therapeutic implications. Signal Transduct Target Ther. 2020;5(1):231.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Xiao W, Wang RS, Handy DE, Loscalzo J. NAD(H) and NADP(H) redox couples and cellular energy metabolism. Antioxid Redox Signal. 2018;28(3):251–72.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Xie N, Zhang L, Gao W, Huang C, Huber PE, Zhou X, et al. NAD(+) metabolism: pathophysiologic mechanisms and therapeutic potential. Signal Transduct Target Ther. 2020;5(1):227.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Koju N, Qin ZH, Sheng R. Reduced nicotinamide adenine dinucleotide phosphate in redox balance and diseases: a friend or foe? Acta Pharmacol Sin. 2022;43(8):1889–904.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87(1):245–313.

    Article  PubMed  CAS  Google Scholar 

  37. Kobayashi M, Yamamoto M. Molecular mechanisms activating the Nrf2-Keap1 pathway of antioxidant gene regulation. Antioxid Redox Signal. 2005;7(3–4):385–94.

    Article  PubMed  CAS  Google Scholar 

  38. Cebula M, Schmidt EE, Arner ES. TrxR1 as a potent regulator of the Nrf2-Keap1 response system. Antioxid Redox Signal. 2015;23(10):823–53.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Morgan MJ, Liu ZG. Crosstalk of reactive oxygen species and NF-kappaB signaling. Cell Res. 2011;21(1):103–15.

    Article  PubMed  CAS  Google Scholar 

  40. Averill-Bates D. Reactive oxygen species and cell signalling. Review. Biochim Biophys Acta Mol Cell Res. 2024;1871(2):119573.

    Article  PubMed  CAS  Google Scholar 

  41. Steinberg SF. Mechanisms for redox-regulation of protein kinase C. Front Pharmacol. 2015;6:128.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Zhang X, Ma L, Wang J. Cross-regulation between redox and epigenetic systems in tumorigenesis: molecular mechanisms and clinical applications. Antioxid Redox Signal. 2023;39(7–9):445–71.

    Article  PubMed  CAS  Google Scholar 

  43. Harris IS, DeNicola GM. The complex interplay between antioxidants and ROS in cancer. Trends Cell Biol. 2020;30(6):440–51.

    Article  PubMed  CAS  Google Scholar 

  44. Jiang D, Rusling JF. Oxidation chemistry of DNA and p53 tumor suppressor gene. ChemistryOpen. 2019;8(3):252–65.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Helfinger V, Schroder K. Redox control in cancer development and progression. Mol Aspects Med. 2018;63:88–98.

    Article  PubMed  CAS  Google Scholar 

  46. Ebrahimi SO, Reiisi S, Shareef S. miRNAs, oxidative stress, and cancer: a comprehensive and updated review. J Cell Physiol. 2020;235(11):8812–25.

    Article  PubMed  CAS  Google Scholar 

  47. Wang Y, Qi H, Liu Y, Duan C, Liu X, Xia T, et al. The double-edged roles of ROS in cancer prevention and therapy. Theranostics. 2021;11(10):4839–57.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Chatterjee R, Chatterjee J. ROS and oncogenesis with special reference to EMT and stemness. Eur J Cell Biol. 2020;99(2–3): 151073.

    Article  PubMed  CAS  Google Scholar 

  49. Perillo B, Di Donato M, Pezone A, Di Zazzo E, Giovannelli P, Galasso G, et al. ROS in cancer therapy: the bright side of the moon. Exp Mol Med. 2020;52(2):192–203.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Kennedy L, Sandhu JK, Harper ME, Cuperlovic-Culf M. Role of glutathione in cancer: from mechanisms to therapies. Biomolecules. 2020;10(10):1429.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Trachootham D, Alexandre J, Huang P. Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat Rev Drug Discov. 2009;8(7):579–91.

    Article  PubMed  CAS  Google Scholar 

  52. Hu X, Jin H, Zhu L. Effect of glutamine metabolism on chemoresistance and its mechanism in tumors. Zhejiang Da Xue Xue Bao Yi Xue Ban. 2021;50(1):32–40.

    PubMed  PubMed Central  CAS  Google Scholar 

  53. Kim MJ, Choi YK, Park SY, Jang SY, Lee JY, Ham HJ, et al. PPARdelta reprograms glutamine metabolism in sorafenib-resistant HCC. Mol Cancer Res. 2017;15(9):1230–42.

    Article  PubMed  CAS  Google Scholar 

  54. Arnold NB, Ketterer K, Kleeff J, Friess H, Buchler MW, Korc M. Thioredoxin is downstream of Smad7 in a pathway that promotes growth and suppresses cisplatin-induced apoptosis in pancreatic cancer. Cancer Res. 2004;64(10):3599–606.

    Article  PubMed  CAS  Google Scholar 

  55. Barral AM, Kallstrom R, Sander B, Rosen A. Thioredoxin, thioredoxin reductase and tumour necrosis factor-alpha expression in melanoma cells: correlation to resistance against cytotoxic attack. Melanoma Res. 2000;10(4):331–43.

    Article  PubMed  CAS  Google Scholar 

  56. Wang J, Yang H, Li W, Xu H, Yang X, Gan L. Thioredoxin 1 upregulates FOXO1 transcriptional activity in drug resistance in ovarian cancer cells. Biochim Biophys Acta. 2015;1852(3):395–405.

    Article  PubMed  CAS  Google Scholar 

  57. Jovanovic M, Podolski-Renic A, Krasavin M, Pesic M. The role of the thioredoxin detoxification system in cancer progression and resistance. Front Mol Biosci. 2022;9: 883297.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Li YX, Jiang B, Wang RY, Wang JYN, Li YC, Bao YM. Synergistic effects of tanshinone IIA and andrographolide on the apoptosis of cancer cells via crosstalk between p53 and reactive oxygen species pathways. Pharmacol Rep. 2020;72(2):400.

    Article  PubMed  CAS  Google Scholar 

  59. Luo N, Zhang K, Li X, Hu Y, Guo L. Tanshinone IIA destabilizes SLC7A11 by regulating PIAS4-mediated SUMOylation of SLC7A11 through KDM1A, and promotes ferroptosis in breast cancer. J Adv Res. 2024. https://doi.org/10.1016/j.jare.2024.04.009.

    Article  PubMed  Google Scholar 

  60. Guan ZH, Chen J, Li XL, Dong N. Tanshinone IIA induces ferroptosis in gastric cancer cells through p53-mediated SLC7A11 down-regulation. 2020. Biosci Rep. https://doi.org/10.1042/BSR20201807.

  61. Zhao LP, Wang HJ, Hu D, Hu JH, Guan ZR, Yu LH, et al. beta-Elemene induced ferroptosis via TFEB-mediated GPX4 degradation in EGFR wide-type non-small cell lung cancer. J Adv Res. 2023. https://doi.org/10.1016/j.jare.2023.08.018.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Xi C, Ying Z, Wu L, Sian C, Feng L, Xi Z, et al. Celastrol induces ROS-mediated apoptosis via directly targeting peroxiredoxin-2 in gastric cancer cells. Theranostics. 2020;10(22):10290.

    Article  Google Scholar 

  63. Heng X, Hongfang Z, Chunyong D, Defang J, Zijie Z, Yang L, et al. Celastrol suppresses colorectal cancer via covalent targeting peroxiredoxin 1. Signal Transduct Target Ther. 2023. https://doi.org/10.1038/s41392-022-01231-4.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Li H, Zhang J, Sun L, Li B, Gao H, Xie T, et al. Celastrol induces apoptosis and autophagy via the ROS/JNK signaling pathway in human osteosarcoma cells: an in vitro and in vivo study. Cell Death Dis. 2015;6(1): e1604.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Liu X, Zhao P, Wang X, Wang L, Zhu Y, Song Y, et al. Celastrol mediates autophagy and apoptosis via the ROS/JNK and Akt/mTOR signaling pathways in glioma cells. Journal of experimental & clinical cancer research : CR. 2019;38(1):184.

    Article  PubMed Central  Google Scholar 

  66. Zhou XJ, Chen Y, Wang FF, Wu HS, Zhang Y, Liu JX, et al. Artesunate induces autophagy dependent apoptosis through upregulating ROS and activating AMPK-mTOR-ULK1 axis in human bladder cancer cells. Chem Biol Interact. 2020;331:109273.

    Article  PubMed  CAS  Google Scholar 

  67. Zhang QT, Yi HM, Yao H, Lu L, He GC, Wu M, et al. Artemisinin derivatives inhibit non-small cell lung cancer cells through induction of ROS-dependent apoptosis/ferroptosis. J Cancer. 2021;12(13):4075.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Roh J, Kim E, Jang H, Shin D. Nrf2 inhibition reverses the resistance of cisplatin-resistant head and neck cancer cells to artesunate-induced ferroptosis. Redox Biol. 2017;11:254–62.

    Article  PubMed  CAS  Google Scholar 

  69. Liang ST, Chen C, Chen RX, Li R, Chen WL, Jiang GH, et al. Michael acceptor molecules in natural products and their mechanism of action. Front Pharmacol. 2022;13:1033003.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Chen X, Dai X, Liu Y, He X, Gong G. Isodon rubescens (Hemls.) Hara.: a comprehensive review on traditional uses, phytochemistry, and pharmacological activities. Front Pharmacol. 2022;13:766581.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  71. Ding Y, Ding C, Ye N, Liu Z, Wold EA, Chen H, et al. Discovery and development of natural product oridonin-inspired anticancer agents. Eur J Med Chem. 2016;122:102–17.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Li Y, Li N, Shi J, Ahmed T, Liu H, Guo J, et al. Involvement of glutathione depletion in selective cytotoxicity of oridonin to p53-mutant esophageal squamous carcinoma cells. Front Oncol. 2019;9:1525.

    Article  PubMed  Google Scholar 

  73. Yu Y, Fan SM, Song JK, Tashiro S, Onodera S, Ikejima T. Hydroxyl radical (.OH) played a pivotal role in oridonin-induced apoptosis and autophagy in human epidermoid carcinoma A431 cells. Biol Pharm Bull. 2012;35(12):2148–59.

    Article  PubMed  CAS  Google Scholar 

  74. Ding C, Zhang Y, Chen H, Yang Z, Wild C, Ye N, et al. Oridonin ring A-based diverse constructions of enone functionality: identification of novel dienone analogues effective for highly aggressive breast cancer by inducing apoptosis. J Med Chem. 2013;56(21):8814–25.

    Article  PubMed  CAS  Google Scholar 

  75. LoBianco F, Krager K, Johnson E, Godwin C, Allen A, Crooks P, et al. Parthenolide induces rapid thiol oxidation that leads to ferroptosis in hepatocellular carcinoma cells. Front Toxicol. 2022;4: 936149.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Zhang S, Ong C, Shen H. Critical roles of intracellular thiols and calcium in parthenolide-induced apoptosis in human colorectal cancer cells. Cancer Lett. 2004;208(2):143–53.

    Article  PubMed  CAS  Google Scholar 

  77. Skalska J, Brookes P, Nadtochiy S, Hilchey S, Jordan C, Guzman M, et al. Modulation of cell surface protein free thiols: a potential novel mechanism of action of the sesquiterpene lactone parthenolide. PLoS ONE. 2009;4(12): e8115.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Duan D, Zhang J, Yao J, Liu Y, Fang J. Targeting thioredoxin reductase by parthenolide contributes to inducing apoptosis of HeLa cells. J Biol Chem. 2016;291(19):10021–31.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  79. Sun Y, St Clair D, Xu Y, Crooks P, St CW. A NADPH oxidase-dependent redox signaling pathway mediates the selective radiosensitization effect of parthenolide in prostate cancer cells. Can Res. 2010;70(7):2880–90.

    Article  CAS  Google Scholar 

  80. Wang Y, Guo SH, Shang XJ, Yu LS, Zhu JW, Zhao A, et al. Triptolide induces Sertoli cell apoptosis in mice via ROS/JNK-dependent activation of the mitochondrial pathway and inhibition of Nrf2-mediated antioxidant response. Acta Pharmacol Sin. 2018;39(2):311–27.

    Article  PubMed  CAS  Google Scholar 

  81. Yu D, Liu Y, Zhou Y, Ruiz-Rodado V, Larion M, Xu G, et al. Triptolide suppresses IDH1-mutated malignancy via Nrf2-driven glutathione metabolism. Proc Natl Acad Sci U S A. 2020;117(18):9964–72.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  82. Zhu J, Wang H, Chen F, Lv H, Xu Z, Fu J, et al. Triptolide enhances chemotherapeutic efficacy of antitumor drugs in non-small-cell lung cancer cells by inhibiting Nrf2-ARE activity. Toxicol Appl Pharmacol. 2018;358:1–9.

    Article  PubMed  CAS  Google Scholar 

  83. Wang HF, Zhao ZL. Triptolide inhibits proliferation and invasion of colorectal cancer cells by blocking Nrf2 expression. Chem Biol Drug Des. 2024;103(1): e14410.

    Article  PubMed  Google Scholar 

  84. Chen P, Zhong X, Song Y, Zhong W, Wang S, Wang J, et al. Triptolide induces apoptosis and cytoprotective autophagy by ROS accumulation via directly targeting peroxiredoxin 2 in gastric cancer cells. Cancer Lett. 2024;587: 216622.

    Article  PubMed  CAS  Google Scholar 

  85. Liu X, Zhao P, Wang X, Wang L, Zhu Y, Gao W. Triptolide induces glioma cell autophagy and apoptosis via upregulating the ROS/JNK and downregulating the Akt/mTOR signaling pathways. Front Oncol. 2019;9:387.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Tan BJ, Chiu GN. Role of oxidative stress, endoplasmic reticulum stress and ERK activation in triptolide-induced apoptosis. Int J Oncol. 2013;42(5):1605–12.

    Article  PubMed  CAS  Google Scholar 

  87. Chen C, Yang S, Zhang M, Zhang Z, Hong J, Han D, et al. Triptolide mitigates radiation-induced pulmonary fibrosis via inhibition of axis of alveolar macrophages-NOXes-ROS-myofibroblasts. Cancer Biol Ther. 2016;17(4):381–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  88. Li L, Jin P, Guan Y, Luo M, Wang Y, He B, et al. Exploiting polyphenol-mediated redox reorientation in cancer therapy. Pharmaceuticals. 2022;15(12):1540.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Choi J, Kim J, Lee H, Pak J, Shim B, Kim S. Reactive oxygen species and p53 mediated activation of p38 and caspases is critically involved in kaempferol induced apoptosis in colorectal cancer cells. J Agric Food Chem. 2018;66(38):9960–7.

    Article  PubMed  CAS  Google Scholar 

  90. Sharma V, Joseph C, Ghosh S, Agarwal A, Mishra M, Sen E. Kaempferol induces apoptosis in glioblastoma cells through oxidative stress. Mol Cancer Ther. 2007;6(9):2544–53.

    Article  PubMed  CAS  Google Scholar 

  91. Wang F, Wang L, Qu C, Chen L, Geng Y, Cheng C, et al. Kaempferol induces ROS-dependent apoptosis in pancreatic cancer cells via TGM2-mediated Akt/mTOR signaling. BMC Cancer. 2021;21(1):396.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Fouzder C, Mukhuty A, Kundu R. Kaempferol inhibits Nrf2 signalling pathway via downregulation of Nrf2 mRNA and induces apoptosis in NSCLC cells. Arch Biochem Biophys. 2021;697: 108700.

    Article  PubMed  CAS  Google Scholar 

  93. Kim T, Park J, Woo J. Resveratrol induces cell death through ROS-dependent downregulation of Notch1/PTEN/Akt signaling in ovarian cancer cells. Mol Med Rep. 2019;19(4):3353–60.

    PubMed  CAS  Google Scholar 

  94. Kunnumakkara AB, Bordoloi D, Padmavathi G, Monisha J, Roy NK, Prasad S, et al. Curcumin, the golden nutraceutical: multitargeting for multiple chronic diseases. Br J Pharmacol. 2017;174(11):1325–48.

    Article  PubMed  CAS  Google Scholar 

  95. Rodriguez-Garcia A, Hevia D, Mayo JC, Gonzalez-Menendez P, Coppo L, Lu J, et al. Thioredoxin 1 modulates apoptosis induced by bioactive compounds in prostate cancer cells. Redox Biol. 2017;12:634–47.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  96. Chen X, Chen X, Zhang X, Wang L, Cao P, Rajamanickam V, et al. Curcuminoid B63 induces ROS-mediated paraptosis-like cell death by targeting TrxR1 in gastric cells. Redox Biol. 2019;21: 101061.

    Article  PubMed  CAS  Google Scholar 

  97. Fang J, Lu J, Holmgren A. Thioredoxin reductase is irreversibly modified by curcumin: a novel molecular mechanism for its anticancer activity. J Biol Chem. 2005;280(26):25284–90.

    Article  PubMed  CAS  Google Scholar 

  98. Javvadi P, Hertan L, Kosoff R, Datta T, Kolev J, Mick R, et al. Thioredoxin reductase-1 mediates curcumin-induced radiosensitization of squamous carcinoma cells. Cancer Res. 2010;70(5):1941–50.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. Jayakumar S, Patwardhan RS, Pal D, Singh B, Sharma D, Kutala VK, et al. Mitochondrial targeted curcumin exhibits anticancer effects through disruption of mitochondrial redox and modulation of TrxR2 activity. Free Radic Biol Med. 2017;113:530–8.

    Article  PubMed  CAS  Google Scholar 

  100. Kaushik G, Kaushik T, Yadav SK, Sharma SK, Ranawat P, Khanduja KL, et al. Curcumin sensitizes lung adenocarcinoma cells to apoptosis via intracellular redox status mediated pathway. Indian J Exp Biol. 2012;50(12):853–61.

    PubMed  CAS  Google Scholar 

  101. Liu GY, Zhai Q, Chen JZ, Zhang ZQ, Yang J. 2,2’-Fluorine mono-carbonyl curcumin induce reactive oxygen species-Mediated apoptosis in Human lung cancer NCI-H460 cells. Eur J Pharmacol. 2016;786:161–8.

    Article  PubMed  CAS  Google Scholar 

  102. Alhasawi MAI, Aatif M, Muteeb G, Alam MW, Oirdi ME, Farhan M. Curcumin and its derivatives induce apoptosis in human cancer cells by mobilizing and redox cycling genomic copper ions. Molecules. 2022;27(21):7410.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  103. Ahsan H, Parveen N, Khan NU, Hadi SM. Pro-oxidant, anti-oxidant and cleavage activities on DNA of curcumin and its derivatives demethoxycurcumin and bisdemethoxycurcumin. Chem Biol Interact. 1999;121(2):161–75.

    Article  PubMed  CAS  Google Scholar 

  104. Ma L, Li M, Zhang Y, Liu K. Recent advances of antitumor leading compound erianin: mechanisms of action and structural modification. Eur J Med Chem. 2023;261: 115844.

    Article  PubMed  CAS  Google Scholar 

  105. Wang H, Zhang T, Sun W, Wang Z, Zuo D, Zhou Z, et al. Erianin induces G2/M-phase arrest, apoptosis, and autophagy via the ROS/JNK signaling pathway in human osteosarcoma cells in vitro and in vivo. Cell Death Dis. 2016;7(6): e2247.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  106. Zhang X, Wang Y, Li X, Yang A, Li Z, Wang D. The anti-carcinogenesis properties of erianin in the modulation of oxidative stress-mediated apoptosis and immune response in liver cancer. Aging. 2019;11(22):10284–300.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  107. Miao Q, Deng W, Lyu W, Sun Z, Fan S, Qi M, et al. Erianin inhibits the growth and metastasis through autophagy-dependent ferroptosis in KRAS colorectal cancer. Free Radical Biol Med. 2023;204:301–12.

    Article  CAS  Google Scholar 

  108. Chen P, Wu Q, Feng J, Yan L, Sun Y, Liu S, et al. Erianin, a novel dibenzyl compound in dendrobium extract, inhibits lung cancer cell growth and migration via calcium/calmodulin-dependent ferroptosis. Signal Transduct Target Ther. 2020;5(1):51.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  109. Xiang Y, Chen X, Wang W, Zhai L, Sun X, Feng J, et al. Natural product erianin inhibits bladder cancer cell growth by inducing ferroptosis via NRF2 inactivation. Front Pharmacol. 2021;12: 775506.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  110. Yang D, Guo Q, Liang Y, Zhao Y, Tian X, Ye Y, et al. Wogonin induces cellular senescence in breast cancer via suppressing TXNRD2 expression. Arch Toxicol. 2020;94(10):3433–47.

    Article  PubMed  CAS  Google Scholar 

  111. Lu J, Papp L, Fang J, Rodriguez-Nieto S, Zhivotovsky B, Holmgren A. Inhibition of mammalian thioredoxin reductase by some flavonoids: implications for myricetin and quercetin anticancer activity. Can Res. 2006;66(8):4410–8.

    Article  CAS  Google Scholar 

  112. Kim I, Kim J, Lee J, Cho E. Genistein decreases cellular redox potential, partially suppresses cell growth in HL-60 leukemia cells and sensitizes cells to γ-radiation-induced cell death. Mol Med Rep. 2014;10(6):2786–92.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Han S, Lin F, Qi Y, Liu C, Zhou L, Xia Y, et al. HO-1 contributes to luteolin-triggered ferroptosis in clear cell renal cell carcinoma via increasing the labile iron pool and promoting lipid peroxidation. Oxid Med Cell Longev. 2022;2022:3846217.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Souza R, Bonfim-Mendonça P, Gimenes F, Ratti B, Kaplum V, Bruschi M, et al. Oxidative stress triggered by apigenin induces apoptosis in a comprehensive panel of human cervical cancer-derived cell lines. Oxid Med Cell Longev. 2017;2017:1512745.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Ben Sghaier M, Pagano A, Mousslim M, Ammari Y, Kovacic H, Luis J. Rutin inhibits proliferation, attenuates superoxide production and decreases adhesion and migration of human cancerous cells. Biomed Pharmacother. 2016;84:1972–8.

    Article  PubMed  CAS  Google Scholar 

  116. Liu ZH, Yang CX, Zhang L, Yang CY, Xu XQ. Baicalein, as a prooxidant, triggers mitochondrial apoptosis in MCF-7 human breast cancer cells through mobilization of intracellular copper and reactive oxygen species generation. Onco Targets Ther. 2019;12:10749–61.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  117. Kong N, Chen X, Feng J, Duan T, Liu S, Sun X, et al. Baicalin induces ferroptosis in bladder cancer cells by downregulating FTH1. Acta Pharm Sin B. 2021;11(12):4045–54.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Yuan J, Khan SU, Yan J, Lu J, Yang C, Tong Q. Baicalin enhances the efficacy of 5-fluorouracil in gastric cancer by promoting ROS-mediated ferroptosis. Biomed Pharmacother. 2023;164: 114986.

    Article  PubMed  CAS  Google Scholar 

  119. Lu M, He CL, Wu ZT, Lyu Y, Duan XH, Wang BX, et al. Effect of baicalin on pyroptosis of diffuse large B-cell lymphoma cell lines DB and its mechanism. Zhongguo Shi Yan Xue Ye Xue Za Zhi. 2023;31(6):1706–13.

    PubMed  Google Scholar 

  120. Ding L, Dang S, Sun M, Zhou D, Sun Y, Li E, et al. Quercetin induces ferroptosis in gastric cancer cells by targeting SLC1A5 and regulating the p-Camk2/p-DRP1 and NRF2/GPX4 Axes. Free Radic Biol Med. 2024;213:150–63.

    Article  PubMed  CAS  Google Scholar 

  121. Kusaczuk M, Krętowski R, Naumowicz M, Stypułkowska A, Cechowska-Pasko M. A preliminary study of the effect of quercetin on cytotoxicity, apoptosis, and stress responses in glioblastoma cell lines. Int J Mol Sci. 2022;23(3):1345.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  122. Zhang Q, Cheng G, Qiu H, Zhu L, Ren Z, Zhao W, et al. The p53-inducible gene 3 involved in flavonoid-induced cytotoxicity through the reactive oxygen species-mediated mitochondrial apoptotic pathway in human hepatoma cells. Food Funct. 2015;6(5):1518–25.

    Article  PubMed  CAS  Google Scholar 

  123. Raja S, Rajendiran V, Kasinathan N, Amirthalakshmi P, Venkatabalasubramanian S, Murali M, et al. Differential cytotoxic activity of Quercetin on colonic cancer cells depends on ROS generation through COX-2 expression. Food Chem Toxicol. 2017;106:92–106.

    Article  PubMed  CAS  Google Scholar 

  124. Sanchez-Cruz P, Alegria AE. Quinone-enhanced reduction of nitric oxide by xanthine/xanthine oxidase. Chem Res Toxicol. 2009;22(5):818–23.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  125. Bolton JL, Dunlap T. Formation and biological targets of quinones: cytotoxic versus cytoprotective effects. Chem Res Toxicol. 2017;30(1):13–37.

    Article  PubMed  CAS  Google Scholar 

  126. Mimnaugh EG, Dusre L, Atwell J, Myers CE. Differential oxygen radical susceptibility of adriamycin-sensitive and -resistant MCF-7 human breast tumor cells. Cancer Res. 1989;49(1):8–15.

    PubMed  CAS  Google Scholar 

  127. Narayanan P, Farghadani R, Nyamathulla S, Rajarajeswaran J, Thirugnanasampandan R, Bhuwaneswari G. Natural quinones induce ROS-mediated apoptosis and inhibit cell migration in PANC-1 human pancreatic cancer cell line. J Biochem Mol Toxicol. 2022;36(5): e23008.

    Article  PubMed  CAS  Google Scholar 

  128. Guo C, He J, Song X, Tan L, Wang M, Jiang P, et al. Pharmacological properties and derivatives of shikonin-A review in recent years. Pharmacol Res. 2019;149: 104463.

    Article  PubMed  CAS  Google Scholar 

  129. Wang KJ, Meng XY, Chen JF, Wang KY, Zhou C, Yu R, et al. Emodin induced necroptosis and inhibited glycolysis in the renal cancer cells by enhancing ROS. Oxid Med Cell Longev. 2021;2021:8840590.

    Article  PubMed  PubMed Central  Google Scholar 

  130. Zhan S, Lu L, Pan SS, Wei XQ, Miao RR, Liu XH, et al. Targeting NQO1/GPX4-mediated ferroptosis by plumbagin suppresses in vitro and in vivo glioma growth. Br J Cancer. 2022;127(2):364–76.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  131. Yao L, Yan D, Jiang B, Xue Q, Chen X, Huang Q, et al. Plumbagin is a novel GPX4 protein degrader that induces apoptosis in hepatocellular carcinoma cells. Free Radical Biol Med. 2023;203:1–10.

    Article  CAS  Google Scholar 

  132. Wang Z, Tang T, Wang S, Cai T, Tao H, Zhang Q, et al. Aloin inhibits the proliferation and migration of gastric cancer cells by regulating NOX2-ROS-mediated pro-survival signal pathways. Drug Des Dev Ther. 2020;14:145–55.

    Article  CAS  Google Scholar 

  133. Chang CY, Chan HL, Lin HY, Way TD, Kao MC, Song MZ, et al. Rhein induces apoptosis in human breast cancer cells. Evid Based Complement Alternat Med. 2012;2012: 952504.

    Article  PubMed  Google Scholar 

  134. Zhang H, Ma L, Kim E, Yi J, Huang H, Kim H, et al. Rhein induces oral cancer cell apoptosis and ROS via suppresse AKT/mTOR signaling pathway In Vitro and In Vivo. Int J Mol Sci. 2023;24(10):8507.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  135. Wang A, Jiang H, Liu Y, Chen J, Zhou X, Zhao C, et al. Rhein induces liver cancer cells apoptosis via activating ROS-dependent JNK/Jun/caspase-3 signaling pathway. J Cancer. 2020;11(2):500–7.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Ross D, Siegel D. The diverse functionality of NQO1 and its roles in redox control. Redox Biol. 2021;41: 101950.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  137. Bey EA, Bentle MS, Reinicke KE, Dong Y, Yang CR, Girard L, et al. An NQO1- and PARP-1-mediated cell death pathway induced in non-small-cell lung cancer cells by beta-lapachone. Proc Natl Acad Sci USA. 2007;104(28):11832–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  138. Bey EA, Reinicke KE, Srougi MC, Varnes M, Anderson VE, Pink JJ, et al. Catalase abrogates beta-lapachone-induced PARP1 hyperactivation-directed programmed necrosis in NQO1-positive breast cancers. Mol Cancer Ther. 2013;12(10):2110–20.

    Article  PubMed  CAS  Google Scholar 

  139. Zhao W, Jiang L, Fang T, Fang F, Liu Y, Zhao Y, et al. beta-lapachone selectively kills hepatocellular carcinoma cells by targeting NQO1 to Induce extensive DNA damage and PARP1 hyperactivation. Front Oncol. 2021;11: 747282.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Huynh DTN, Jin Y, Myung CS, Heo KS. Ginsenoside Rh1 induces MCF-7 cell apoptosis and autophagic cell death through ROS-mediated akt signaling. Cancers. 2021;13(8):1892.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  141. Li B, Zhao J, Wang CZ, Searle J, He TC, Yuan CS, et al. Ginsenoside Rh2 induces apoptosis and paraptosis-like cell death in colorectal cancer cells through activation of p53. Cancer Lett. 2011;301(2):185–92.

    Article  PubMed  CAS  Google Scholar 

  142. Biswas P, Ghorai M, Mishra T, Gopalakrishnan AV, Roy D, Mane AB, et al. Piper longum L.: a comprehensive review on traditional uses, phytochemistry, pharmacology, and health-promoting activities. Phytother Res. 2022;36(12):4425–76.

    Article  PubMed  CAS  Google Scholar 

  143. Parama D, Rana V, Girisa S, Verma E, Daimary UD, Thakur KK, et al. The promising potential of piperlongumine as an emerging therapeutics for cancer. Explor Target Antitumor Ther. 2021;2(4):323–54.

    PubMed  PubMed Central  CAS  Google Scholar 

  144. Cui Y, Chen XB, Liu Y, Wang Q, Tang J, Chen MJ. Piperlongumine inhibits esophageal squamous cell carcinoma in vitro and in vivo by triggering NRF2/ROS/TXNIP/NLRP3-dependent pyroptosis. Chem Biol Interact. 2024;390: 110875.

    Article  PubMed  CAS  Google Scholar 

  145. Zou P, Xia Y, Ji J, Chen W, Zhang J, Chen X, et al. Piperlongumine as a direct TrxR1 inhibitor with suppressive activity against gastric cancer. Cancer Lett. 2016;375(1):114–26.

    Article  PubMed  CAS  Google Scholar 

  146. Liu JM, Pan F, Li L, Liu QR, Chen Y, Xiong XX, et al. Piperlongumine selectively kills glioblastoma multiforme cells via reactive oxygen species accumulation dependent JNK and p38 activation. Biochem Biophys Res Commun. 2013;437(1):87–93.

    Article  PubMed  CAS  Google Scholar 

  147. Baranoski A, Semprebon SC, Biazi BI, Zanetti TA, Corveloni AC, Areal Marques L, et al. Piperlongumine inhibits antioxidant enzymes, increases ROS levels, induces DNA damage and G2/M cell cycle arrest in breast cell lines. J Toxicol Environ Health A. 2024;87(7):294–309.

    Article  PubMed  CAS  Google Scholar 

  148. Bansal A, Simon MC. Glutathione metabolism in cancer progression and treatment resistance. J Cell Biol. 2018;217(7):2291–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  149. Niu B, Zhou Y, Liao K, Wen T, Lao S, Quan G, et al. “Pincer movement”: reversing cisplatin resistance based on simultaneous glutathione depletion and glutathione S-transferases inhibition by redox-responsive degradable organosilica hybrid nanoparticles. Acta Pharm Sin B. 2022;12(4):2074–88.

    Article  PubMed  CAS  Google Scholar 

  150. Yang H, Wang J, Khan S, Zhang Y, Zhu K, Zhou E, et al. Selective synergistic anticancer effects of cisplatin and oridonin against human p53-mutant esophageal squamous carcinoma cells. Anticancer Drugs. 2022;33(1):e444–52.

    Article  PubMed  CAS  Google Scholar 

  151. Zheng W, Zhou CY, Zhu XQ, Wang XJ, Li ZY, Chen XC, et al. Oridonin enhances the cytotoxicity of 5-FU in renal carcinoma cells by inducting necroptotic death. Biomed Pharmacother. 2018;106:175–82.

    Article  PubMed  CAS  Google Scholar 

  152. Tripathi SK, Panda M, Biswal BK. Emerging role of plumbagin: cytotoxic potential and pharmaceutical relevance towards cancer therapy. Food Chem Toxicol. 2019;125:566–82.

    Article  PubMed  CAS  Google Scholar 

  153. Kong X, Luo J, Xu T, Zhou Y, Pan Z, Xie Y, et al. Plumbagin enhances TRAIL-induced apoptosis of human leukemic Kasumi-1 cells through upregulation of TRAIL death receptor expression, activation of caspase-8 and inhibition of cFLIP. Oncol Rep. 2017;37(6):3423–32.

    Article  PubMed  CAS  Google Scholar 

  154. Allensworth JL, Aird KM, Aldrich AJ, Batinic-Haberle I, Devi GR. XIAP inhibition and generation of reactive oxygen species enhances TRAIL sensitivity in inflammatory breast cancer cells. Mol Cancer Ther. 2012;11(7):1518–27.

    Article  PubMed  CAS  Google Scholar 

  155. Li X, Wang H, Wang J, Chen Y, Yin X, Shi G, et al. Emodin enhances cisplatin-induced cytotoxicity in human bladder cancer cells through ROS elevation and MRP1 downregulation. BMC Cancer. 2016;16:578.

    Article  PubMed  PubMed Central  Google Scholar 

  156. Xu Z, Zhao D, Zheng X, Huang B, Xia X, Pan X. Quercetin exerts bidirectional regulation effects on the efficacy of tamoxifen in estrogen receptor-positive breast cancer therapy: an in vitro study. Environ Toxicol. 2020;35(11):1179–93.

    Article  PubMed  CAS  Google Scholar 

  157. Kim SJ, Miyoshi Y, Taguchi T, Tamaki Y, Nakamura H, Yodoi J, et al. High thioredoxin expression is associated with resistance to docetaxel in primary breast cancer. Clin Cancer Res. 2005;11(23):8425–30.

    Article  PubMed  CAS  Google Scholar 

  158. Yokomizo A, Ono M, Nanri H, Makino Y, Ohga T, Wada M, et al. Cellular levels of thioredoxin associated with drug sensitivity to cisplatin, mitomycin C, doxorubicin, and etoposide. Cancer Res. 1995;55(19):4293–6.

    PubMed  CAS  Google Scholar 

  159. Yamada M, Tomida A, Yoshikawa H, Taketani Y, Tsuruo T. Increased expression of thioredoxin/adult T-cell leukemia-derived factor in cisplatin-resistant human cancer cell lines. Clin Cancer Res. 1996;2(2):427–32.

    PubMed  CAS  Google Scholar 

  160. Zhu P, Qian J, Xu Z, Meng C, Zhu W, Ran F, et al. Overview of piperlongumine analogues and their therapeutic potential. Eur J Med Chem. 2021;220: 113471.

    Article  PubMed  CAS  Google Scholar 

  161. Zhang P, Shi L, Zhang T, Hong L, He W, Cao P, et al. Piperlongumine potentiates the antitumor efficacy of oxaliplatin through ROS induction in gastric cancer cells. Cell Oncol. 2019;42(6):847–60.

    Article  CAS  Google Scholar 

  162. Zhang Y, Sun S, Xu W, Yang R, Yang Y, Guo J, et al. Thioredoxin reductase 1 inhibitor shikonin promotes cell necroptosis via SecTRAPs generation and oxygen-coupled redox cycling. Free Radic Biol Med. 2022;180:52–62.

    Article  PubMed  CAS  Google Scholar 

  163. Li X, Fan XX, Jiang ZB, Loo WT, Yao XJ, Leung EL, et al. Shikonin inhibits gefitinib-resistant non-small cell lung cancer by inhibiting TrxR and activating the EGFR proteasomal degradation pathway. Pharmacol Res. 2017;115:45–55.

    Article  PubMed  CAS  Google Scholar 

  164. Sun S, Zhang Y, Xu W, Yang R, Yang Y, Guo J, et al. Plumbagin reduction by thioredoxin reductase 1 possesses synergy effects with GLUT1 inhibitor on KEAP1-mutant NSCLC cells. Biomed Pharmacother. 2022;146: 112546.

    Article  PubMed  CAS  Google Scholar 

  165. Hasan AA, Kalinina E, Nuzhina J, Volodina Y, Shtil A, Tatarskiy V. Potentiation of cisplatin cytotoxicity in resistant ovarian cancer skov3/cisplatin cells by quercetin pre-treatment. Int J Mol Sci. 2023;24(13):10960.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  166. Zhu B, Ren C, Du K, Zhu H, Ai Y, Kang F, et al. Olean-28,13b-olide 2 plays a role in cisplatin-mediated apoptosis and reverses cisplatin resistance in human lung cancer through multiple signaling pathways. Biochem Pharmacol. 2019;170: 113642.

    Article  PubMed  CAS  Google Scholar 

  167. Zhang J, Li X, Han X, Liu R, Fang J. Targeting the thioredoxin system for cancer therapy. Trends Pharmacol Sci. 2017;38(9):794–808.

    Article  PubMed  CAS  Google Scholar 

  168. Huang LJ, Lan JX, Wang JH, Huang H, Lu K, Zhou ZN, et al. Bioactivity and mechanism of action of sanguinarine and its derivatives in the past 10 years. Biomed Pharmacother. 2024;173: 116406.

    Article  PubMed  CAS  Google Scholar 

  169. Leung EL, Fan XX, Wong MP, Jiang ZH, Liu ZQ, Yao XJ, et al. Targeting tyrosine kinase inhibitor-resistant non-small cell lung cancer by inducing epidermal growth factor receptor degradation via methionine 790 oxidation. Antioxid Redox Signal. 2016;24(5):263–79.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  170. Zheng K, Li Y, Wang S, Wang X, Liao C, Hu X, et al. Inhibition of autophagosome-lysosome fusion by ginsenoside Ro via the ESR2-NCF1-ROS pathway sensitizes esophageal cancer cells to 5-fluorouracil-induced cell death via the CHEK1-mediated DNA damage checkpoint. Autophagy. 2016;12(9):1593–613.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  171. Zhang Y, Huang L, Shi H, Chen H, Tao J, Shen R, et al. Ursolic acid enhances the therapeutic effects of oxaliplatin in colorectal cancer by inhibition of drug resistance. Cancer Sci. 2018;109(1):94–102.

    Article  PubMed  CAS  Google Scholar 

  172. Ding N, Zhang H, Su S, Ding Y, Yu X, Tang Y, et al. Emodin enhances the chemosensitivity of endometrial cancer by inhibiting ROS-mediated cisplatin-resistance. Anticancer Agents Med Chem. 2018;18(7):1054–63.

    Article  PubMed  CAS  Google Scholar 

  173. Sporn MB, Liby KT. NRF2 and cancer: the good, the bad and the importance of context. Nat Rev Cancer. 2012;12(8):564–71.

    Article  PubMed  CAS  Google Scholar 

  174. Qian C, Wang Y, Zhong Y, Tang J, Zhang J, Li Z, et al. Wogonin-enhanced reactive oxygen species-induced apoptosis and potentiated cytotoxic effects of chemotherapeutic agents by suppression Nrf2-mediated signaling in HepG2 cells. Free Radic Res. 2014;48(5):607–21.

    Article  PubMed  CAS  Google Scholar 

  175. Zhong Y, Zhang F, Sun Z, Zhou W, Li ZY, You QD, et al. Drug resistance associates with activation of Nrf2 in MCF-7/DOX cells, and wogonin reverses it by down-regulating Nrf2-mediated cellular defense response. Mol Carcinog. 2013;52(10):824–34.

    Article  PubMed  CAS  Google Scholar 

  176. Kim EH, Jang H, Shin D, Baek SH, Roh JL. Targeting Nrf2 with wogonin overcomes cisplatin resistance in head and neck cancer. Apoptosis. 2016;21(11):1265–78.

    Article  PubMed  CAS  Google Scholar 

  177. Chian S, Zhao Y, Xu M, Yu X, Ke X, Gao R, et al. Ginsenoside Rd reverses cisplatin resistance in non-small-cell lung cancer A549 cells by downregulating the nuclear factor erythroid 2-related factor 2 pathway. Anticancer Drugs. 2019;30(8):838–45.

    Article  PubMed  CAS  Google Scholar 

  178. Zhang R, Qiao H, Chen S, Chen X, Dou K, Wei L, et al. Berberine reverses lapatinib resistance of HER2-positive breast cancer cells by increasing the level of ROS. Cancer Biol Ther. 2016;17(9):925–34.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  179. Tang Z, Wang L, Chen Y, Zheng X, Wang R, Liu B, et al. Quercetin reverses 5-fluorouracil resistance in colon cancer cells by modulating the NRF2/HO-1 pathway. Eur J Histochem. 2023;67(3):3719.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  180. Hasan AAS, Kalinina EV, Tatarskiy VV, Volodina YL, Petrova Capital AC, Novichkova MD, et al. Suppression of the antioxidant system and PI3K/Akt/mTOR signaling pathway in cisplatin-resistant cancer cells by quercetin. Bull Exp Biol Med. 2022;173(6):760–4.

    Article  PubMed  CAS  Google Scholar 

  181. Carlisi D, De Blasio A, Drago-Ferrante R, Di Fiore R, Buttitta G, Morreale M, et al. Parthenolide prevents resistance of MDA-MB231 cells to doxorubicin and mitoxantrone: the role of Nrf2. Cell Death Discov. 2017;3:17078.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  182. Zhong YY, Chen HP, Tan BZ, Yu HH, Huang XS. Triptolide avoids cisplatin resistance and induces apoptosis via the reactive oxygen species/nuclear factor-kappaB pathway in SKOV3(PT) platinum-resistant human ovarian cancer cells. Oncol Lett. 2013;6(4):1084–92.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  183. Zhang P, Zhang J, Zhao L, Li S, Li K. Quercetin attenuates the cardiotoxicity of doxorubicin-cyclophosphamide regimen and potentiates its chemotherapeutic effect against triple-negative breast cancer. Phytother Res. 2022;36(1):551–61.

    Article  PubMed  CAS  Google Scholar 

  184. Moutabian H, Majdaeen M, Ghahramani-Asl R, Yadollahi M, Gharepapagh E, Ataei G, et al. A systematic review of the therapeutic effects of resveratrol in combination with 5-fluorouracil during colorectal cancer treatment: with a special focus on the oxidant, apoptotic, and anti-inflammatory activities. Cancer Cell Int. 2022;22(1):142.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  185. Paunovic MG, Matic MM, Obradovic AD, Jevtic VV, Stojkovic DL, Ognjanovic BI. Antiproliferative, antimigratory, and prooxidative potential of novel platinum(IV) complexes and resveratrol on breast cancer (MDA-MB-231) and choriocarcinoma (JEG-3) cell lines. Drug Dev Res. 2022;83(3):688–98.

    PubMed  CAS  Google Scholar 

  186. Yuan Y, Xue X, Guo RB, Sun XL, Hu G. Resveratrol enhances the antitumor effects of temozolomide in glioblastoma via ROS-dependent AMPK-TSC-mTOR signaling pathway. CNS Neurosci Ther. 2012;18(7):536–46.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  187. Lin CJ, Lee CC, Shih YL, Lin TY, Wang SH, Lin YF, et al. Resveratrol enhances the therapeutic effect of temozolomide against malignant glioma in vitro and in vivo by inhibiting autophagy. Free Radic Biol Med. 2012;52(2):377–91.

    Article  PubMed  CAS  Google Scholar 

  188. Zhou Y, Huang S, Guo Y, Ran M, Shan W, Chen WH, et al. Epigallocatechin gallate circumvents drug-induced resistance in non-small-cell lung cancer by modulating glucose metabolism and AMPK/AKT/MAPK axis. Phytother Res. 2023;37(12):5837–53.

    Article  PubMed  CAS  Google Scholar 

  189. Wei R, Zhu G, Jia N, Yang W. Epigallocatechin-3-gallate sensitizes human 786-O renal cell carcinoma cells to TRAIL-induced apoptosis. Cell Biochem Biophys. 2015;72(1):157–64.

    Article  PubMed  CAS  Google Scholar 

  190. Wu W, Gou H, Xiang B, Geng R, Dong J, Yang X, et al. EGCG enhances the chemosensitivity of colorectal cancer to irinotecan through GRP78-mediatedendoplasmic reticulum stress. J Oncol. 2022;2022:7099589.

    Article  PubMed  PubMed Central  Google Scholar 

  191. Mei Y, Wei D, Liu J. Reversal of multidrug resistance in KB cells with tea polyphenol antioxidant capacity. Cancer Biol Ther. 2005;4(4):468–73.

    Article  PubMed  CAS  Google Scholar 

  192. Tan X, Zhou Y, Agarwal A, Lim M, Xu Y, Zhu Y, et al. Systemic application of honokiol prevents cisplatin ototoxicity without compromising its antitumor effect. Am J Cancer Res. 2020;10(12):4416–34.

    PubMed  PubMed Central  CAS  Google Scholar 

  193. Katanic Stankovic JS, Selakovic D, Rosic G. Oxidative damage as a fundament of systemic toxicities induced by Cisplatin-the crucial limitation or potential therapeutic target? Int J Mol Sci. 2023;24(19):14574.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  194. Liu HT, Wang TE, Hsu YT, Chou CC, Huang KH, Hsu CC, et al. Nanoparticulated Honokiol mitigates cisplatin-induced chronic kidney injury by maintaining mitochondria antioxidant capacity and reducing caspase 3-associated cellular apoptosis. Antioxidants. 2019;8(10):466.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  195. Cheng W, Xiang W, Wang S, Xu K. Tanshinone IIA ameliorates oxaliplatin-induced neurotoxicity via mitochondrial protection and autophagy promotion. Am J Transl Res. 2019;11(5):3140–9.

    PubMed  PubMed Central  CAS  Google Scholar 

  196. Paciello F, Fetoni AR, Mezzogori D, Rolesi R, Di Pino A, Paludetti G, et al. The dual role of curcumin and ferulic acid in counteracting chemoresistance and cisplatin-induced ototoxicity. Sci Rep. 2020;10(1):1063.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  197. Palipoch S, Punsawad C, Koomhin P, Suwannalert P. Hepatoprotective effect of curcumin and alpha-tocopherol against cisplatin-induced oxidative stress. BMC Complement Altern Med. 2014;14:111.

    Article  PubMed  PubMed Central  Google Scholar 

  198. Liao D, Shangguan D, Wu Y, Chen Y, Liu N, Tang J, et al. Curcumin protects against doxorubicin induced oxidative stress by regulating the Keap1-Nrf2-ARE and autophagy signaling pathways. Psychopharmacology. 2023;240(5):1179–90.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  199. Yu D, Li J, Wang Y, Guo D, Zhu C, Sun B, et al. Oridonin ameliorates doxorubicin induced-cardiotoxicity via the E2F1/Sirt6/PGC1alpha pathway in mice. Food Chem Toxicol. 2023;181: 114050.

    Article  PubMed  CAS  Google Scholar 

  200. Liu X, Xu J, Zhou J, Shen Q. Oridonin and its derivatives for cancer treatment and overcoming therapeutic resistance. Genes Dis. 2021;8(4):448–62.

    Article  PubMed  CAS  Google Scholar 

  201. Wang YY, Liao J, Luo YL, Li MS, Su XY, Yu B, et al. Berberine alleviates doxorubicin-induced myocardial injury and fibrosis by eliminating oxidative stress and mitochondrial damage via promoting Nrf-2 pathway activation. Int J Mol Sci. 2023;24(4):448.

    Google Scholar 

  202. Wu Q, Chen J, Zheng X, Song J, Yin L, Guo H, et al. Kaempferol attenuates doxorubicin-induced renal tubular injury by inhibiting ROS/ASK1-mediated activation of the MAPK signaling pathway. Biomed Pharmacother. 2023;157:114087.

    Article  PubMed  CAS  Google Scholar 

  203. Yi J, Zhu J, Zhao C, Kang Q, Zhang X, Suo K, et al. Potential of natural products as radioprotectors and radiosensitizers: opportunities and challenges. Food Funct. 2021;12(12):5204–18.

    Article  PubMed  CAS  Google Scholar 

  204. Das U, Manna K, Adhikary A, Mishra S, Saha KD, Sharma RD, et al. Ferulic acid enhances the radiation sensitivity of lung and liver carcinoma cells by collapsing redox homeostasis: mechanistic involvement of Akt/p38 MAPK signalling pathway. Free Radic Res. 2019;53(9–10):944–67.

    Article  PubMed  CAS  Google Scholar 

  205. Kim JS, Heo K, Yi JM, Gong EJ, Yang K, Moon C, et al. Genistein mitigates radiation-induced testicular injury. Phytother Res. 2012;26(8):1119–25.

    Article  PubMed  CAS  Google Scholar 

  206. Peng S, Wang Y, Sun Z, Zhao L, Huang Y, Fu X, et al. Nanoparticles loaded with pharmacologically active plant-derived natural products: Biomedical applications and toxicity. Colloids Surf B Biointerfaces. 2023;225: 113214.

    Article  PubMed  CAS  Google Scholar 

  207. Xia N, Wan W, Zhu S, Liu Q. Synthesis of hydrophobic propionyl neohesperidin ester using an immobilied enzyme and description of its anti-proliferative and pro-apoptotic effects on MCF-7 human breast cancer cells. Front Bioeng Biotechnol. 2020;8:1025.

    Article  PubMed  PubMed Central  Google Scholar 

  208. Bai ZQ, Yao CS, Zhu JL, Xie YY, Ye XY, Bai RR, et al. Anti-tumor drug discovery based on natural product β-elemene: anti-tumor mechanisms and structural modification. Molecules. 2021;26(6):1499.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  209. Wen Y, Li K, Ni M, Jiang H, Wu H, Yu Q, et al. Dendritic polylysine with paclitaxel and triptolide codelivery for enhanced cancer ferroptosis through the accumulation of ROS. ACS Appl Mater Interfaces. 2024. https://doi.org/10.1021/acsami.4c00558.

    Article  PubMed  Google Scholar 

  210. Zhen W, Liu Y, An S, Jiang X. Glutathione-induced In Situ michael addition between nanoparticles for pyroptosis and immunotherapy. Angew Chem Int Ed Engl. 2023;62(20): e202301866.

    Article  PubMed  CAS  Google Scholar 

  211. Yan Y, Yang X, Han N, Liu Y, Liang Q, Li L, et al. Metal-organic framework-encapsulated dihydroartemisinin nanoparticles induces apoptotic cell death in ovarian cancer by blocking ROMO1-mediated ROS production. J Nanobiotechnol. 2023;21(1):204.

    Article  CAS  Google Scholar 

  212. Cui D, Liang T, Sun L, Meng L, Yang C, Wang L, et al. Green synthesis of selenium nanoparticles with extract of hawthorn fruit induced HepG2 cells apoptosis. Pharm Biol. 2018;56(1):528–34.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  213. Cheng Y, Zhao P, Wu S, Yang T, Chen Y, Zhang X, et al. Cisplatin and curcumin co-loaded nano-liposomes for the treatment of hepatocellular carcinoma. Int J Pharm. 2018;545:261–73.

    Article  PubMed  CAS  Google Scholar 

  214. Yu J, Wang Y, Zhou S, Li J, Wang J, Chi D, et al. Remote loading paclitaxel-doxorubicin prodrug into liposomes for cancer combination therapy. Acta Pharm Sinica B. 2020;10(9):1730–40.

    Article  CAS  Google Scholar 

  215. Ahmad J, Ahamad J, Algahtani M, Garg A, Shahzad N, Ahmad M, et al. Nanotechnology-mediated delivery of resveratrol as promising strategy to improve therapeutic efficacy in triple negative breast cancer (TNBC): progress and promises. Expert Opin Drug Deliv. 2024;21(2):229–44.

    Article  PubMed  CAS  Google Scholar 

  216. Huang Q, Li Y, Huang Z, Jun M, Wang W, Chen X, et al. Artesunate carriers induced ferroptosis to overcome biological barriers for anti-cancer. Eur J Pharm Biopharm. 2023;190:284–93.

    Article  PubMed  CAS  Google Scholar 

  217. Kolewe M, Gaurav V, Roberts S. Pharmaceutically active natural product synthesis and supply via plant cell culture technology. Mol Pharm. 2008;5(2):243–56.

    Article  PubMed  CAS  Google Scholar 

  218. Li C, Zhang R, Wang J, Wilson L, Yan Y. Protein engineering for improving and diversifying natural product biosynthesis. Trends Biotechnol. 2020;38(7):729–44.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  219. Atanasov A, Waltenberger B, Pferschy-Wenzig E, Linder T, Wawrosch C, Uhrin P, et al. Discovery and resupply of pharmacologically active plant-derived natural products: a review. Biotechnol Adv. 2015;33(8):1582–614.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  220. Sun XN, Zhang YT, Zhou Y, Lian XC, Yan LL, Pan T, et al. NPCDR: natural product-based drug combination and its disease-specific molecular regulation. Nucleic Acids Res. 2022;50:D1324–33.

    Article  PubMed  CAS  Google Scholar 

  221. Ma XX, Meng YJ, Wang P, Tang ZH, Wang HZ, Xie T. Bioinformatics-assisted, integrated omics studies on medicinal plants. Brief Bioinform. 2020;21(6):1857.

    Article  PubMed  CAS  Google Scholar 

  222. Zhang RN, Pan T, Xiang Y, Zhang MM, Feng J, Liu SP, et al. β-elemene reverses the resistance of p53-deficient colorectal cancer cells to 5-fluorouracil by inducing pro-death autophagy and cyclin D3-dependent cycle arrest. Front Bioeng Biotechnol. 2020;8:378.

    Article  PubMed  PubMed Central  Google Scholar 

  223. Chen P, Li XJ, Zhang RN, Liu SP, Xiang Y, Zhang MM, et al. Combinative treatment of β-elemene and cetuximab is sensitive to KRAS mutant colorectal cancer cells by inducing ferroptosis and inhibiting epithelial-mesenchymal transformation. Theranostics. 2020;10(11):5107.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  224. Schumacker P. Reactive oxygen species in cancer cells: live by the sword, die by the sword. Cancer Cell. 2006;10(3):175–6.

    Article  PubMed  CAS  Google Scholar 

  225. Trachootham D, Zhou Y, Zhang H, Demizu Y, Chen Z, Pelicano H, et al. Selective killing of oncogenically transformed cells through a ROS-mediated mechanism by beta-phenylethyl isothiocyanate. Cancer Cell. 2006;10(3):241–52.

    Article  PubMed  CAS  Google Scholar 

  226. Lee Y, Lee J, Lim C. Anticancer activity of flavonoids accompanied by redox state modulation and the potential for a chemotherapeutic strategy. Food Sci Biotechnol. 2021;30(3):321–40.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  227. Chaiswing L, Zhong W, Cullen J, Oberley L, Oberley T. Extracellular redox state regulates features associated with prostate cancer cell invasion. Can Res. 2008;68(14):5820–6.

    Article  CAS  Google Scholar 

  228. Lara G, Marcello P, Milena N, Jonas PM, Sara DB, Erika R, et al. Quercetin and cancer chemoprevention. Evid Based Complement Alternat Med. 2011. https://doi.org/10.1093/ecam/neq053.

    Article  Google Scholar 

  229. Freund R, Gobrecht P, Fischer D, Arndt H. Advances in chemistry and bioactivity of parthenolide. Nat Prod Rep. 2020;37(4):541–65.

    Article  PubMed  CAS  Google Scholar 

  230. Xu Y, Fang F, Miriyala S, Crooks PA, Oberley TD, Chaiswing L, et al. KEAP1 is a redox sensitive target that arbitrates the opposing radiosensitive effects of parthenolide in normal and cancer cells. Cancer Res. 2013;73(14):4406.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  231. Wang W, Adachi M, Kawamura R, Sakamoto H, Hayashi T, Ishida T, et al. Parthenolide-induced apoptosis in multiple myeloma cells involves reactive oxygen species generation and cell sensitivity depends on catalase activity. Apoptosis. 2006;11(12):2225–35.

    Article  PubMed  CAS  Google Scholar 

  232. Eun-Ryeong H, Kozue S, Shivendra VS. Honokiol activates reactive oxygen species-mediated cytoprotective autophagy in human prostate cancer cells. Prostate. 2014;74(12):1209.

    Article  Google Scholar 

  233. Song GQ, Wu P, Dong XM, Cheng LH, Lu HQ, Lin YY, et al. Elemene induces cell apoptosis via inhibiting glutathione synthesis in lung adenocarcinoma. J Ethnopharmacol. 2023;311:116409.

    Article  PubMed  CAS  Google Scholar 

  234. Huang Z, Gan S, Zhuang X, Chen Y, Lu L, Wang Y, et al. Artesunate inhibits the cell growth in colorectal cancer by promoting ROS-dependent cell senescence and autophagy. Cells. 2022;11(16):6472.

    Article  Google Scholar 

  235. Zhang F, Hao Y, Yang N, Liu M, Luo Y, Zhang Y, et al. Oridonin-induced ferroptosis and apoptosis: a dual approach to suppress the growth of osteosarcoma cells. BMC Cancer. 2024;24(1):198.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  236. D’Anneo A, Carlisi D, Lauricella M, Puleio R, Martinez R, Di Bella S, et al. Parthenolide generates reactive oxygen species and autophagy in MDA-MB231 cells. A soluble parthenolide analogue inhibits tumour growth and metastasis in a xenograft model of breast cancer. Cell Death Dis. 2013;4(10):891.

    Article  Google Scholar 

  237. Zhang J, Li Y, Duan D, Yao J, Gao K, Fang J. Inhibition of thioredoxin reductase by alantolactone prompts oxidative stress-mediated apoptosis of HeLa cells. Biochem Pharmacol. 2016;102:34–44.

    Article  PubMed  CAS  Google Scholar 

  238. Yang S, Evens AM, Prachand S, Singh AT, Bhalla S, David K, et al. Mitochondrial-mediated apoptosis in lymphoma cells by the diterpenoid lactone andrographolide, the active component of Andrographis paniculata. Clin Cancer Res. 2010;16(19):4755–68.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  239. Xie JH, Lai ZQ, Zheng XH, Liao HJ, Xian YF, Li Q, et al. Apoptotic activities of brusatol in human non-small cell lung cancer cells: involvement of ROS-mediated mitochondrial-dependent pathway and inhibition of Nrf2-mediated antioxidant response. Toxicology. 2021. https://doi.org/10.1016/j.tox.2021.152680.

    Article  PubMed  Google Scholar 

  240. Lee N, Meng R, Rah S, Jin H, Ray N, Kim S, et al. Reactive oxygen species-mediated autophagy by ursolic acid inhibits growth and metastasis of esophageal cancer cells. Int J Mol Sci. 2020;21(24):9409.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  241. Delaram M, Reza Z, Amirabbas R, Mohammad S-N, Rozita G, Amir S, et al. The growth inhibitory effect of resveratrol and gallic acid on prostate cancer cell lines through the alteration of oxidative stress balance: the interplay between Nrf2, HO-1, and BACH1 genes. Anticancer Agents Med Chem. 2024. https://doi.org/10.2174/0118715206317999240708062744.

    Article  Google Scholar 

  242. Liao K, Lee Y, Chao W, Huang Y, Chung H, Chen S, et al. Honokiol suppresses cell proliferation and tumor migration through ros in human anaplastic thyroid cancer cells. Endocrine Metab Immune Disord Drug Targets. 2024. https://doi.org/10.2174/0118715303295608240408082523.

    Article  Google Scholar 

  243. Yeh PS, Wang W, Chang YA, Lin CJ, Wang JJ, Chen RM. Honokiol induces autophagy of neuroblastoma cells through activating the PI3K/Akt/mTOR and endoplasmic reticular stress/ERK1/2 signaling pathways and suppressing cell migration. Cancer Lett. 2016;370(1):66–77.

    Article  PubMed  CAS  Google Scholar 

  244. Qian X, Zhu L, Xu M, Liu H, Yu X, Shao Q, et al. Shikonin suppresses small cell lung cancer growth via inducing ATF3-mediated ferroptosis to promote ROS accumulation. Chem Biol Interact. 2023;382: 110588.

    Article  PubMed  CAS  Google Scholar 

  245. Tsai M, Chen S, Ong A, Chung Y, Chen P, Hsieh Y, et al. Shikonin induced program cell death through generation of reactive oxygen species in renal cancer cells. Antioxidants. 2021;10(11):1831.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  246. Wang P, Gao C, Wang W, Yao LP, Zhang J, Zhang SD, et al. Juglone induces apoptosis and autophagy via modulation of mitogen-activated protein kinase pathways in human hepatocellular carcinoma cells. Food Chem Toxicol. 2018;116(Pt B):40–50.

    PubMed  CAS  Google Scholar 

  247. Zhou X, Wang F, Wu H, Chen X, Zhang Y, Lin J, et al. Thymoquinone suppresses the proliferation, migration and invasiveness through regulating ROS, autophagic flux and miR-877-5p in human bladder carcinoma cells. Int J Biol Sci. 2021;17(13):3456–75.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  248. Zheng K, Liao C, Li Y, Fan X, Fan L, Xu H, et al. Gypenoside L, isolated from gynostemma pentaphyllum, induces cytoplasmic vacuolation death in hepatocellular carcinoma cells through reactive-oxygen-species-mediated unfolded protein response. J Agric Food Chem. 2016;64(8):1702–11.

    Article  PubMed  CAS  Google Scholar 

  249. Hussain AR, Al-Jomah NA, Siraj AK, Manogaran P, Al-Hussein K, Abubaker J, et al. Sanguinarine-dependent induction of apoptosis in primary effusion lymphoma cells. Cancer Res. 2007;67(8):3888–97.

    Article  PubMed  CAS  Google Scholar 

  250. Liu M, Sun S, Meng Y, Wang L, Liu H, Shi W, et al. Benzophenanthridine alkaloid chelerythrine elicits necroptosis of gastric cancer cells via selective conjugation at the redox hyperreactive C-terminal sec(498) residue of cytosolic selenoprotein thioredoxin reductase. Molecules. 2023;28(19):6842.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  251. Wen-Hsiu H, Yih-Shou H, Hsing-Chun K, Chun-Yuh T, Hai-I H, Chau-Jong W, et al. Berberine induces apoptosis in SW620 human colonic carcinoma cells through generation of reactive oxygen species and activation of JNK/p38 MAPK and FasL. Arch Toxicol. 2007;81(10):719.

    Article  Google Scholar 

  252. Cao Y, Zhang H, Tang J, Wang R. Ferulic acid mitigates growth and invasion of esophageal squamous cell carcinoma through inducing ferroptotic cell death. Dis Markers. 2022;2022:4607966.

    Article  PubMed  PubMed Central  Google Scholar 

  253. Kim JW, Choi J, Park MN, Kim B. Apoptotic effect of gallic acid via regulation of p-p38 and ER stress in PANC-1 and MIA PaCa-2 cells pancreatic cancer cells. Int J Mol Sci. 2023;24(20):15236.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  254. Chen G, Wang K, Yang BY, Tang B, Chen JX, Hua ZC. Synergistic antitumor activity of oridonin and arsenic trioxide on hepatocellular carcinoma cells. Int J Oncol. 2012;40(1):139–47.

    PubMed  Google Scholar 

  255. Cao S, Xia M, Mao Y, Zhang Q, Donkor PO, Qiu F, et al. Combined oridonin with cetuximab treatment shows synergistic anticancer effects on laryngeal squamous cell carcinoma: involvement of inhibition of EGFR and activation of reactive oxygen species-mediated JNK pathway. Int J Oncol. 2016;49(5):2075–87.

    Article  PubMed  CAS  Google Scholar 

  256. Xue D, Pan ST, Zhou X, Ye F, Zhou Q, Shi F, et al. Plumbagin enhances the anticancer efficacy of cisplatin by increasing intracellular ROS in human tongue squamous cell carcinoma. Oxid Med Cell Longev. 2020;2020:5649174.

    Article  PubMed  PubMed Central  Google Scholar 

  257. Wang W, Sun Y, Huang X, He M, Chen Y, Shi G, et al. Emodin enhances sensitivity of gallbladder cancer cells to platinum drugs via glutathion depletion and MRP1 downregulation. Biochem Pharmacol. 2010;79(8):1134–40.

    Article  PubMed  CAS  Google Scholar 

  258. Ma J, Yang J, Wang C, Zhang N, Dong Y, Wang C, et al. Emodin augments cisplatin cytotoxicity in platinum-resistant ovarian cancer cells via ROS-dependent MRP1 downregulation. Biomed Res Int. 2014;2014: 107671.

    Article  PubMed  PubMed Central  Google Scholar 

  259. Markowitsch S, Schupp P, Lauckner J, Vakhrusheva O, Slade K, Mager R, et al. Artesunate inhibits growth of sunitinib-resistant renal cell carcinoma cells through cell cycle arrest and induction of ferroptosis. Cancers. 2020;12(11):3150.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  260. Li J, Sharkey CC, King MR. Piperlongumine and immune cytokine TRAIL synergize to promote tumor death. Sci Rep. 2015;5:9987.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  261. Zheng L, Fang S, Chen A, Chen W, Qiao E, Chen M, et al. Piperlongumine synergistically enhances the antitumour activity of sorafenib by mediating ROS-AMPK activation and targeting CPSF7 in liver cancer. Pharmacol Res. 2022;177: 106140.

    Article  PubMed  CAS  Google Scholar 

  262. Ni M, Zhou J, Zhu Z, Xu Q, Yin Z, Wang Y, et al. Shikonin and cisplatin synergistically overcome cisplatin resistance of ovarian cancer by inducing ferroptosis via upregulation of HMOX1 to promote Fe(2+) accumulation. Phytomedicine. 2023;112: 154701.

    Article  PubMed  CAS  Google Scholar 

  263. Wang L, Stadlbauer B, Lyu C, Buchner A, Pohla H. Shikonin enhances the antitumor effect of cabazitaxel in prostate cancer stem cells and reverses cabazitaxel resistance by inhibiting ABCG2 and ALDH3A1. Am J Cancer Res. 2020;10(11):3784–800.

    PubMed  PubMed Central  CAS  Google Scholar 

  264. Lange M, Abhari BA, Hinrichs TM, Fulda S, Liese J. Identification of a novel oxidative stress induced cell death by Sorafenib and oleanolic acid in human hepatocellular carcinoma cells. Biochem Pharmacol. 2016;118:9–17.

    Article  PubMed  CAS  Google Scholar 

  265. Sarkhosh-Inanlou R, Molaparast M, Mohammadzadeh A, Shafiei-Irannejad V. Sanguinarine enhances cisplatin sensitivity via glutathione depletion in cisplatin-resistant ovarian cancer (A2780) cells. Chem Biol Drug Des. 2020;95(2):215–23.

    Article  PubMed  CAS  Google Scholar 

  266. Lin MY, Cheng WT, Cheng HC, Chou WC, Chen HI, Ou HC, et al. Baicalin enhances chemosensitivity to doxorubicin in breast cancer cells via upregulation of oxidative stress-mediated mitochondria-dependent apoptosis. Antioxidants. 2021;10(10):1506.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  267. Luo F, Zhao J, Liu S, Xue Y, Tang D, Yang J, et al. Ursolic acid augments the chemosensitivity of drug-resistant breast cancer cells to doxorubicin by AMPK-mediated mitochondrial dysfunction. Biochem Pharmacol. 2022;205: 115278.

    Article  PubMed  CAS  Google Scholar 

  268. Li H, Yu Y, Liu Y, Luo Z, Law BYK, Zheng Y, et al. Ursolic acid enhances the antitumor effects of sorafenib associated with Mcl-1-related apoptosis and SLC7A11-dependent ferroptosis in human cancer. Pharmacol Res. 2022;182: 106306.

    Article  PubMed  CAS  Google Scholar 

  269. Lee DH, Rhee JG, Lee YJ. Reactive oxygen species up-regulate p53 and Puma; a possible mechanism for apoptosis during combined treatment with TRAIL and wogonin. Br J Pharmacol. 2009;157(7):1189–202.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  270. Hou D, Xu G, Zhang C, Li B, Qin J, Hao X, et al. Berberine induces oxidative DNA damage and impairs homologous recombination repair in ovarian cancer cells to confer increased sensitivity to PARP inhibition. Cell Death Dis. 2017;8(10): e3070.

    Article  PubMed  PubMed Central  Google Scholar 

  271. Chen F, Liu Y, Wang S, Guo X, Shi P, Wang W, et al. Triptolide, a Chinese herbal extract, enhances drug sensitivity of resistant myeloid leukemia cell lines through downregulation of HIF-1alpha and Nrf2. Pharmacogenomics. 2013;14(11):1305–17.

    Article  PubMed  CAS  Google Scholar 

  272. Yang Y, Zhang LJ, Bai XG, Xu HJ, Jin ZL, Ding M. Synergistic antitumour effects of triptolide plus gemcitabine in bladder cancer. Biomed Pharmacother. 2018;106:1307–16.

    Article  PubMed  CAS  Google Scholar 

  273. Xiang Y, Ye W, Huang C, Yu D, Chen H, Deng T, et al. Brusatol enhances the chemotherapy efficacy of gemcitabine in pancreatic cancer via the Nrf2 signalling pathway. Oxid Med Cell Longev. 2018;2018:2360427.

    Article  PubMed  PubMed Central  Google Scholar 

  274. Yang Y, Tian Z, Guo R, Ren F. Nrf2 Inhibitor, Brusatol in combination with trastuzumab exerts synergistic antitumor activity in HER2-positive cancers by inhibiting Nrf2/HO-1 and HER2-AKT/ERK1/2 pathways. Oxid Med Cell Longev. 2020;2020:9867595.

    Article  PubMed  PubMed Central  Google Scholar 

  275. Wu MF, Huang YH, Chiu LY, Cherng SH, Sheu GT, Yang TY. Curcumin induces apoptosis of chemoresistant lung cancer cells via ROS-regulated p38 MAPK phosphorylation. Int J Mol Sci. 2022;23(15):8248.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  276. Gaur T, Ali A, Sharma D, Gupta SK, Gota V, Bagal B, et al. Mitocurcumin utilizes oxidative stress to upregulate JNK/p38 signaling and overcomes Cytarabine resistance in acute myeloid leukemia. Cell Signal. 2024;114: 111004.

    Article  PubMed  CAS  Google Scholar 

  277. Hormann V, Kumi-Diaka J, Durity M, Rathinavelu A. Anticancer activities of genistein-topotecan combination in prostate cancer cells. J Cell Mol Med. 2012;16(11):2631–6.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  278. Jiang H, Ma Y, Chen X, Pan S, Sun B, Krissansen GW, et al. Genistein synergizes with arsenic trioxide to suppress human hepatocellular carcinoma. Cancer Sci. 2010;101(4):975–83.

    Article  PubMed  CAS  Google Scholar 

  279. Wang R, Ma L, Weng D, Yao J, Liu X, Jin F. Gallic acid induces apoptosis and enhances the anticancer effects of cisplatin in human small cell lung cancer H446 cell line via the ROS-dependent mitochondrial apoptotic pathway. Oncol Rep. 2016;35(5):3075–83.

    Article  PubMed  CAS  Google Scholar 

  280. Sanchez-Carranza JN, Diaz JF, Redondo-Horcajo M, Barasoain I, Alvarez L, Lastres P, et al. Gallic acid sensitizes paclitaxel-resistant human ovarian carcinoma cells through an increase in reactive oxygen species and subsequent downregulation of ERK activation. Oncol Rep. 2018;39(6):3007–14.

    PubMed  CAS  Google Scholar 

  281. Gan D, He W, Yin H, Gou X. beta-elemene enhances cisplatin-induced apoptosis in bladder cancer cells through the ROS-AMPK signaling pathway. Oncol Lett. 2020;19(1):291–300.

    PubMed  CAS  Google Scholar 

  282. Gao AM, Ke ZP, Wang JN, Yang JY, Chen SY, Chen H. Apigenin sensitizes doxorubicin-resistant hepatocellular carcinoma BEL-7402/ADM cells to doxorubicin via inhibiting PI3K/Akt/Nrf2 pathway. Carcinogenesis. 2013;34(8):1806–14.

    Article  PubMed  CAS  Google Scholar 

  283. Wang Q, Zheng XL, Yang L, Shi F, Gao LB, Zhong YJ, et al. Reactive oxygen species-mediated apoptosis contributes to chemosensitization effect of saikosaponins on cisplatin-induced cytotoxicity in cancer cells. J Exp Clin Cancer Res. 2010;29(1):159.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  284. Liu Y, Shi C, He Z, Zhu F, Wang M, He R, et al. Inhibition of PI3K/AKT signaling via ROS regulation is involved in Rhein-induced apoptosis and enhancement of oxaliplatin sensitivity in pancreatic cancer cells. Int J Biol Sci. 2021;17(2):589–602.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  285. Perez-Rojas JM, Gonzalez-Macias R, Gonzalez-Cortes J, Jurado R, Pedraza-Chaverri J, Garcia-Lopez P. Synergic effect of alpha-mangostin on the cytotoxicity of cisplatin in a cervical cancer model. Oxid Med Cell Longev. 2016;2016:7981397.

    Article  PubMed  PubMed Central  Google Scholar 

  286. Wang K, Liu X, Liu Q, Ho IH, Wei X, Yin T, et al. Hederagenin potentiated cisplatin- and paclitaxel-mediated cytotoxicity by impairing autophagy in lung cancer cells. Cell Death Dis. 2020;11(8):611.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  287. Kim EH, Baek S, Shin D, Lee J, Roh JL. Hederagenin induces apoptosis in cisplatin-resistant head and neck cancer cells by inhibiting the Nrf2-ARE antioxidant pathway. Oxid Med Cell Longev. 2017;2017:5498908.

    Article  PubMed  PubMed Central  Google Scholar 

  288. Chang CC, Kuan CP, Lin JY, Lai JS, Ho TF. Tanshinone IIA facilitates TRAIL sensitization by up-regulating DR5 through the ROS-JNK-CHOP signaling axis in human ovarian carcinoma cell lines. Chem Res Toxicol. 2015;28(8):1574–83.

    Article  PubMed  CAS  Google Scholar 

  289. Hong H, Cao W, Wang Q, Liu C, Huang C. Synergistic antitumor effect of Andrographolide and cisplatin through ROS-mediated ER stress and STAT3 inhibition in colon cancer. Med Oncol. 2022;39(5):101.

    Article  PubMed  CAS  Google Scholar 

  290. Wei RJ, Zhang XS, He DL. Andrographolide sensitizes prostate cancer cells to TRAIL-induced apoptosis. Asian J Androl. 2018;20(2):200–4.

    Article  PubMed  CAS  Google Scholar 

  291. Rawat L, Balan M, Sasamoto Y, Sabarwal A, Pal S. A novel combination therapy with Cabozantinib and Honokiol effectively inhibits c-Met-Nrf2-induced renal tumor growth through increased oxidative stress. Redox Biol. 2023;68: 102945.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  292. Jayasooriya RG, Choi YH, Hyun JW, Kim GY. Camptothecin sensitizes human hepatoma Hep3B cells to TRAIL-mediated apoptosis via ROS-dependent death receptor 5 upregulation with the involvement of MAPKs. Environ Toxicol Pharmacol. 2014;38(3):959–67.

    Article  PubMed  CAS  Google Scholar 

  293. Mohammadhosseinpour S, Weaver A, Sudhakaran M, Ho LC, Le T, Doseff AI, et al. Arachidin-1, a prenylated stilbenoid from peanut, enhances the anticancer effects of paclitaxel in triple-negative breast cancer cells. Cancers. 2023;15(2):399.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  294. Zhang T, Xu C, Zheng P, Zhang X, Qiu C, Wu F, et al. Glaucocalyxin B attenuates ovarian cancer cell growth and cisplatin resistance In Vitro via activating oxidative stress. Oxid Med Cell Longev. 2022;2022:6324292.

    PubMed  PubMed Central  Google Scholar 

  295. Cao P, Xia Y, He W, Zhang T, Hong L, Zheng P, et al. Enhancement of oxaliplatin-induced colon cancer cell apoptosis by alantolactone, a natural product inducer of ROS. Int J Biol Sci. 2019;15(8):1676–84.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  296. Guo P, Wang S, Liang W, Wang W, Wang H, Zhao M, et al. Salvianolic acid B reverses multidrug resistance in HCT-8/VCR human colorectal cancer cells by increasing ROS levels. Mol Med Rep. 2017;15(2):724–30.

    Article  PubMed  CAS  Google Scholar 

  297. Kong L, Wang X, Zhang K, Yuan W, Yang Q, Fan J, et al. Gypenosides synergistically enhances the anti-tumor effect of 5-fluorouracil on colorectal cancer In Vitro and In Vivo: a role for oxidative stress-mediated DNA damage and p53 activation. PLoS ONE. 2015;10(9): e0137888.

    Article  PubMed  PubMed Central  Google Scholar 

  298. Shi H, Xiong L, Yan G, Du S, Liu J, Shi Y. Susceptibility of cervical cancer to dihydroartemisinin-induced ferritinophagy-dependent ferroptosis. Front Mol Biosci. 2023;10:1156062.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  299. Fath M, Nasiri K, Ghasemzadeh S, Nejati S, Ghafari N, Masouleh S, et al. Thymoquinone potentiates anti-cancer effects of cisplatin in oral squamous cell carcinoma via targeting oxidative stress. Chem Biol Drug Des. 2024;103(3): e14492.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the Science and Technology Development Fund, Macau SAR (Grant No. 0098/2021/A2 and 0048/2023/AFJ).

Funding

The Science and Technology Development Fund, Macau SAR (Grant No. 0098/2021/A2 and 0048/2023/AFJ).

Author information

Authors and Affiliations

Authors

Contributions

YS wrote original manuscript. QL, YH, ZY, GL, XS (Xiaoyu Sun), XG, and YQ contributed materials information gathering. QW and TX revised, edited and supervised manuscript. XS (Xinbing Sui) initiated the study and proposed the concept. All authors contributed to the article and approved the submitted version.

Corresponding authors

Correspondence to Qibiao Wu, Tian Xie or Xinbing Sui.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

All authors consent to the publication of this work in Chinese Medicine.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sun, Y., Li, Q., Huang, Y. et al. Natural products for enhancing the sensitivity or decreasing the adverse effects of anticancer drugs through regulating the redox balance. Chin Med 19, 110 (2024). https://doi.org/10.1186/s13020-024-00982-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13020-024-00982-2

Keywords