Open Access

Role of ROS‑mediated autophagy in melanoma (Review)

  • Authors:
    • Xuebing Zhang
    • Huaijun Li
    • Chengxiang Liu
    • Xingxing Yuan
  • View Affiliations

  • Published online on: August 9, 2022     https://doi.org/10.3892/mmr.2022.12819
  • Article Number: 303
  • Copyright: © Zhang et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Melanoma is the most aggressive form of skin cancer with the poorest prognosis and its pathogenesis has yet to be fully elucidated. As key factors that regulate cellular homeostasis, both reactive oxygen species (ROS) and autophagy are involved in the development of melanoma, from melanomagenesis to progression and drug resistance. However, the interaction between ROS and autophagy in the etiology and treatment of melanoma is not well characterized. The present review examined the production of ROS and the role of oxidative stress in melanoma, and summarized the role of ROS‑mediated autophagy in melanomagenesis and melanoma cell fate decision following treatment with various anticancer drugs. The present findings may lead to a better understanding of the pathogenesis and progression of melanoma, and suggest promising treatment options for this disease.

Introduction

Reactive oxygen species (ROS) mainly refer to superoxide anion (•O2-), hydroxyl radical (•OH), hydrogen peroxide (H2O2) and singlet oxygen, which are derived from the metabolism of O2. These chemically reactive molecules are generated as metabolites of oxidative reactions in mitochondria, the nicotinamide adenine dinucleotide phosphate (NADPH) oxidase system and endoplasmic reticulum (ER) (1). In order to counteract ROS overproduction, multiple antioxidant enzymes exist to maintain redox homeostasis, including superoxide dismutase, catalase, and glutathione peroxidase (2). ROS are capable of reacting with proteins, lipids and nucleic acids, as well as regulate signaling pathways that are involved in various cellular processes (3). Thus, a state of equilibrium between oxidants and reductants is required for physiological processes, such as cell proliferation, differentiation and survival. However, the imbalance in ROS generation and removal causes excessive accumulation of ROS and thus oxidative stress, resulting in detrimental oxidative DNA damage, which is regarded as a major factor for carcinogenesis (4). Controlling ROS metabolism is beneficial for the prevention and treatment of cancer.

Autophagy is an evolutionarily conserved and lysosome-dependent catabolic process whereby cytoplasmic components, such as damaged organelles, protein aggregates and lipid droplets, are degraded and further recycled in autophagosomes for the maintenance of cellular homeostasis (5). In response to ROS-mediated oxidative damage, autophagy can be induced as an adaptive reaction to remove ROS and oxidative biomolecules and organelles, thus alleviating oxidative stress (6). The interaction between ROS and autophagy has been widely investigated in various cancer types, as it participates in tumorigenesis, metastasis and chemoresistance (7). As one of the most aggressive forms of cancer, melanoma is a type of skin cancer with a high metastasis potential and poor survival rate. In comparison with other solid tumors, ROS levels are particularly abundant in melanoma, which is considered a ROS-driven tumor (8). However, the crosstalk between ROS and autophagy in the development of melanoma remains elusive. During tumor progression or chemotherapy-induced stress, obsolete organelles and useless proteins are recycled by autophagy to foster cancer cell growth and chemoresistance. Understanding the role of autophagy in the development of melanoma is of great significance, and autophagy regulation represents a new potential therapeutic target in this disease. Chemotherapeutic agents both induce oxidative damage and autophagy via generation of amounts of ROS. However, cytoprotective autophagy limits the chemotherapy efficacy. Moreover, ROS-mediated autophagy is induced in diverse melanoma chemotherapies where it exerts double functions: i) Promoting cell survival, which is a mechanism of chemoresistance (9); and ii) triggering autophagic cell death, which improves antitumor efficacy (10). The aim of the present review was to emphasize the role of ROS-mediated autophagy in melanomagenesis and melanoma treatments, as well as discuss the therapeutic potential of combination chemotherapy with autophagy-regulating agents according to the functional status of ROS-induced autophagy in patients with melanoma.

In order to summarize the role of ROS-mediated autophagy in melanoma, a PubMed search was performed in April 2022. Articles containing the following key words were considered for inclusion: ‘reactive oxygen species’ (or ‘ROS’ or ‘oxidative stress’) AND ‘autophagy’ AND ‘melanoma’ (or ‘melanomagenesis’). Relevant articles were also identified from a manual search of reference lists within those included. The abstracts of identified articles were screened and classified for inclusion in the review. A total of 60 original articles that have been published in a peer-reviewed journal and written in English were included.

ROS generation in melanoma

Ultraviolet (UV) radiation is a major risk factor for melanoma development. UV can transduce its electromagnetic energy into chemical, hormonal, and neural signals upon absorption, thus regulating homeostatic activity, including activation of the central nervous system and endocrine glands through neural transmission or chemical messengers, which exerts systemic effects on patients with melanoma (11). The oncogenic effect of UV on skin is induced by ROS-mediated DNA damage (12). UV penetrates the epidermis and dermis of the skin and is absorbed by various biomolecules, which generates high levels of ROS in skin cells (13). By tracing the luminescence of singlet oxygen, high ROS levels can be detected during UV exposure both in vitro and in vivo (14). Photosensitizer molecules facilitate ROS production. Vitamins, well-known endogenous photosensitizers, whose chemical structure is altered upon absorption of UV radiation, are susceptible to producing ROS via photosensitized reactions (15). ROS oxidizes fatty acids in cell membranes, lipoproteins, and other lipid-containing molecules, and ultimately leads to impairment of cellular structures and functions. Oxidized fatty acids also produce ROS under UV irradiation and become strong photosensitizers under continuous UV irradiation (16). Photosensitizers such as flavins, urocanic acids, and cholesterols have been shown to produce ROS by absorbing the energy of UV radiation (17). ROS accumulation in the skin changes the absorption of those molecules, which in turn increase ROS production (18). Thus, an increase of ROS initiates a vicious cycle that amplifies the UV-mediated damaging effects on skin cells. Furthermore, UV radiation induces the activation of ROS-producing enzymes. Valencia and Kochevar revealed that UV activated NADPH oxidase (NOX)1 to generate ROS in human keratinocytes, which stimulates prostaglandin E2 synthesis and contributes to skin injury (19). Notably, UV radiation has been revealed to improve the expression of sestrin2 in melanocytes and melanoma cells, which inhibits the antioxidant response factor nuclear factor erythroid 2-related factor 2 (Nrf2) and further aggravates ROS production (20). Excessive ROS produced by UV radiation is considered to cause skin damage and ultimately tumorigenesis, mainly via oxidative stress-induced DNA damage (21).

Compared with those of keratinocytes and fibroblasts, both melanocytes and melanoma cells exhibit higher basal levels of ROS (22). Suppressing melanin synthesis in melanocytes by N phenylthiourea alleviates intracellular ROS (23). These findings indicate that the process of synthesis of melanin is an essential source of ROS. Further mechanistic investigations revealed that the melanosome and its melanin contents were associated with oxidative reactions, where superoxide anion and H2O2 were produced (24). Melanin has both photoprotective and phototoxic properties based on diverse conditions. The biosynthesis of melanin requires a series of oxidoreduction reactions and consumes oxygen, which generates cytotoxic intermediates, such as free radicals. Under physiological conditions, the process of melanin synthesis is limited within the boundaries of melanosomes, and plays a protective role against UV-induced carcinogenesis; however, this process can be dysregulated with cytotoxic intermediates leaking outside melanosomes under pathological conditions, thus contributing to the malignant transformation of melanocytes (25). Melanin exerts a double role in determining ROS levels: It absorbs UV radiation and thus mitigates UV-induced ROS production in melanocytes and keratinocytes; however, melanocytes are in a state of pro-oxidation during melanin synthesis, and become more predisposed to intracellular ROS accumulation and carcinogenesis (23). One explanation could be that the pro-oxidant activity is different between the two forms of melanin in the skin, namely the reddish-yellow pheomelanin and the brown-black eumelanin. L-tyrosine and L-dihydroxyphenylalanine serve as substrates for melanin pigmentation, which is regulated by transcriptional factors including tyrosinase and tyrosinase-related proteins, and signaling pathways involving cAMP and protein kinase C (26). In addition, metal cations such as Mn+2 and Cu+2 stimulate L-dihydroxyphenylalanine auto-oxidation to melanin (27). Melanocytes with high pheomelanin content become pro-oxidant and thus generate ROS upon exposure to UV radiation and metal ions (28,29), while this effect can be prevented by eumelanin, if present in sufficient quantity (30). Thus, the enhanced ratio of pheomelanin to eumelanin in isolated melanosomes leads to mutagenesis through enhanced ROS production (30). Furthermore, eumelanin expression is regulated by the melanocortin 1 receptor gene (MC1R) signaling, which is responsible for ROS scavenging and DNA repair (31). MC1R inactivation results in elevated ROS production and compromised DNA repair, thereby leading to increased risk of carcinogenesis (32). However, increased melanin pigmentation is associated with shorter overall survival and disease-free survival time in patients with melanoma, and inhibition of melanin synthesis improves the radiotherapeutic response (33). The forms of melanin that are responsible for the negative effect of melanin on melanoma therapy remain to be investigated.

The cellular ROS pool in melanocytes and melanoma cells can be also derived from NOX family enzymes. Among them, NOX1, NOX4, and NOX5 have been demonstrated to be expressed in the melanocytic lineage (34). ROS generation induced by these NOX isoforms is required for the proliferation and malignant transformation of melanoma cells. NOX1 was shown to be upregulated in melanoma cell lines and is activated to produce ROS by the mutation of N-RAS, an oncogene, which is involved in melanoma progression (35). NOX4 expression is significantly higher in melanoma tumors compared with that in primary tumors, indicating its oncogenic role in melanoma. Govindaraja et al (36) revealed that NOX4-generated ROS activated by AKT, a serine-threonine kinase highly expressed in melanoma, facilitated the transformation of radial growth to vertical growth that was required for the invasive and metastatic phenotype. In addition, ROS produced by NOX4 promote cell survival through activating the focal adhesion kinase pathway, maintaining cell adhesion and viability (37). Of note, in cultured melanoma cells under hypoxic conditions, NOX4 is involved in the generation of ROS and cell apoptosis mediated by α-melanocyte-stimulating hormone (α-MSH), an inducer of melanin (38). A possible explanation for this proapoptotic role of NOX4 is that NOX4-dependent ROS production sensitizes melanoma cells to TNF-related apoptosis-inducing ligand TRAIL-induced apoptosis via Bax phosphorylation (39). In addition, α-MSH activates its downstream signal transducer melanocyte-inducing transcription factor to stimulate NOX4 gene expression and further drive ROS generation, ultimately suppressing melanin synthesis (40). Silencing of NOX4 expression in melanoma cells attenuates ROS production and thereby suppresses cell growth and tumorigenicity in vivo by regulating G2-M cell cycle progression, suggesting that NOX4-generated ROS elicit a transformation in the phenotype of melanoma cells (41). Furthermore, NOX5 is overexpressed in melanoma, and affects cell proliferation through the ROS-mediated hypoxia-inducible factor (HIF)-1α and p27Kip1 signaling pathways (42). Accordingly, these results indicate that NOX-induced ROS production is associated with melanomagenesis and melanoma progression. Differences in signals and resultant effects involving these ROS-producing NOX enzymes merit further investigation.

In summary, the main contributors to ROS generation in melanoma cells include UV radiation, melanin synthesis and NOX family enzymes (Fig. 1). Excessive ROS leads to oxidative stress, which enhances the melanomagenesis and progression of melanoma.

Oxidative stress and melanomagenesis

Excessive ROS causes oxidative stress, which is recognized as the initiator and promoter of melanoma. The main mechanism by which ROS overproduction promotes melanomagenesis has been well established, and involves the induction of oxidative DNA damage and mutagenesis. The by-products of DNA damage include 8-hydroxydeoxyguanosine (8-OHdG), cyclobutane pyrimidine dimers, pyrimidone adducts, DNA strand breaks, and DNA crosslinks, which result in genomic instability and transcriptional silencing (43). 8-OHdG, a major form of oxidative DNA damage, is regarded as a premutagenic DNA lesion and is highly expressed in melanocytes compared with its levels in keratinocytes (44). Similarly, the expression level of 8-OHdG is lower in patients with melanoma, who have a significantly longer survival time (45). These findings indicate that ROS-mediated DNA oxidation exacerbates a malignant phenotype, and has been implicated in the poor prognosis of patients with melanoma.

Oxidative stress also promotes the occurrence and development of melanoma via genotoxicity. On one hand, it causes oncogene activation. For example, mutations of the BRAF oncogene, occurring in ~50% of melanoma cases, are induced by oxidative stress (46). In addition, ROS stabilize the expression of HIF-1α, a transcriptional regulator of the hypoxic response, to activate the Met protooncogene, which drives the proliferation and metastasis of melanoma cells, and induces angiogenesis (47). Increased expression of HIF-1α can also originate from induction of melanogenesis in melanoma cells, and further regulates cellular metabolism and the behavior of cancer cells (48). Another oncogene, RAC1, which is associated with an increased risk of melanoma, can be activated by high levels of ROS to accelerate the migration and invasion of B16 melanoma cells; however, these effects are weakened by the suppression of ROS-mediated RAC1 activation (49). Notably, tumor suppressor genes compromised in melanoma aggravate oxidative stress. Jenkins et al (50) reported that depletion of p16 expression contributed to marked increases in ROS levels in cultured human melanocytes, which triggers oxidative DNA damage. Additionally, silencing large tumor suppressor kinase 1, a tumor suppressor in melanoma, was demonstrated to lead to enhanced oxidative stress, which is highly engaged in melanoma growth (51). Thus, the loss of tumor suppressor genes increases the susceptibility of melanocytes to oxidative stress and expedites carcinogenesis. On the other hand, oxidative stress mediates epigenetic modifications to induce melanomagenesis. Molognoni et al (52) revealed that ROS increased DNA methyltrasferase 1 and DNA hypermethylation, which led to Ras activation and malignant transformation in melanoma via activation of the ERK signaling pathway.

Furthermore, oxidative stress favors the development of melanoma by the activation of specific signaling pathways. For instance, ROS-mediated oxidation of light chain 8 (LC8), a multifunctional protein of the dynein motor complex, activates the NF-κΒ signaling pathway due to the impaired ability of LC8 to bind to the NF-κB component I-κBα, suppressing its phosphorylation by IKK and ultimately leading to NF-κB activation (53). This effect facilitates melanoma progression by its antiapoptotic effects and creates an inflammatory microenvironment (54). Other signaling pathways, particularly the PI3K/AKT, MAPK/ERK and Nrf2 signaling pathways, are implicated in the initiation and progression of melanoma (35,43).

Melanogenesis is an essential process in melanoma cells. By interacting with hormones, neuropeptides and vitamin D, these cells participate in steroidogenesis and sex hormone conversion, which is required for the production of melanin (55). Local metabolites, such as HIF-1α, are upregulated during this process, and their accumulation can activate HIF-1-related pathways to affect the progression of melanoma (48). Melanosomes act as vital boundaries of the process of melanin synthesis, and restrict cytotoxic intermediates to leak into the surrounding environment, which influences the behavior of melanoma cells, triggering tumorigenesis (56). The induction of melanogenesis is associated with changes in glucose metabolism in melanoma cells (57). UV irradiation and UV-induced ROS also cause a persistent increase in glucose consumption, which is accompanied by increased glycolysis in these cells, and promotes the invasion of melanoma (58). In addition, intermediates of melanogenesis including quinones, semiquinones, quinonimines, and ROS produced during this process exert immunosuppressive functions in melanoma (59). The activity of T lymphocytes and cytokines, such as IL-1, IL-6, TNF-α, and IL-10, is suppressed during this process, and inhibition of melanogenesis increases the lymphocyte-mediated killing effect on melanoma cells (60). Mitochondrial ROS also promote inflammatory cytokines, including TGF-β and IL-13, which induce the activation and M2 polarization of macrophages, thus leading to an immunosuppressive microenvironment and metastatic behaviors in melanoma (61). Therefore, ROS-induced oxidative stress can facilitate melanoma progression through the promotion of glucose metabolism and immunosuppression.

ROS-mediated oxidative stress possibly leads to melanomagenesis through two potentially important effects: Genotoxicity and nongenotoxicity (Fig. 2). The genotoxic effect is induced by oxidative stress-mediated damage to DNA, which causes genetic and epigenetic changes in melanoma-related genes. The nongenotoxic effect is enhanced by the activation of specific signaling pathways that influence numerous cellular processes linked to carcinogenesis and melanoma progression. Moreover, oxidative stress triggers glucose metabolism and immunosuppression and further accelerates melanoma evasion.

Autophagy as a response to alleviate oxidative stress

A complex interaction between ROS and autophagy exists. In response to various stimuli, ROS and autophagy can regulate each other through a variety of signaling pathways, thus determining cell fate, which is largely dependent on the quantity of ROS produced and the antioxidant ability of cells (62). Under hypoxic conditions, large amounts of ROS are accumulated in cancer cells, which subsequently stimulate autophagy (63). As the major species of ROS, H2O2 is generated in these starved tumor cells as a result of PI3K activation, which induces the formation of autophagosomes through oxidation of autophagy-related gene (ATG)4 (64). In addition, both O2 and H2O2 trigger autophagy by AMP-activated protein kinase activation and subsequent mTOR inhibition, as well as through transcriptional regulation of ATGs such as p62 and Beclin 1 (BECN1) (65,66). Consequently, autophagy functions as an antioxidant defense mechanism to mitigate ROS damage to cells and maintains cellular homeostasis by engulfing oxidized substances (67). However, autophagy may provide nutrients and a favorable environment for tumor progression via the degradation of ROS-damaged organelles and proteins (62). In this context, ROS-mediated autophagy may play double roles in melanomagenesis and melanoma progression.

As aforementioned, UV radiation is a major resource for ROS generation in melanoma. Exposure to UV light increases the expression of p62 in an ROS-dependent manner, which involves Nrf2 activity to counteract oxidative stress (68). Deletion of ATG7 in a BRAFV600E/PTEN null model of melanoma was revealed to be associated with enhanced oxidative stress and senescence, which prevent melanoma tumorigenesis (69). These findings indicate that oxidative stress-induced autophagy exerts a tumor-promoting role. Autophagy is also essential for the survival of cancer cells under starvation, which further promotes melanoma metastasis (70). Moreover, in melanoma cells in response to hypoxia and reoxygenation treatment, the persisted accumulation of intracellular ROS is accompanied by increased autophagy (71). Suppression of autophagy markedly accelerates cell death induced by cycling hypoxia via increased ROS generation (72). Therefore, excessive ROS could induce autophagy to protect the survival of melanoma cell under stress conditions. In addition, BRAFV600E has been demonstrated to increase the ER stress response, which subsequently activates cytoprotective autophagy (73). Alterations in the oxidative environment of the ER cause the generation of ER stress-induced ROS (74). For example, NOX4, an essential ROS-producing NOX enzyme in melanoma, can be activated to generate ROS during the ER stress response (75). This process is responsible for high basal autophagy during melanomagenesis. Mechanistically, the JNK-mediated phosphorylation of Bcl-2 and Bcl-xl releases BECN1; meanwhile, tribbles homolog 3 inactivates mTOR signaling, triggering the autophagy process (76). Thus, melanoma cells employ autophagy as an adaptive mechanism to moderate oxidative stress.

In conclusion, autophagy plays a prominent regulatory role in tumor cell proliferation by counteracting ROS-mediated oxidative stress and maintaining cellular homeostasis. However, ROS overproduction prolongs the activation of autophagy and its excessive induction may culminate in autophagic cell death. In this regard, ROS could switch autophagic cell survival to death. Thus, clarification of the role of ROS-induced autophagy in melanoma progression may provide a promising therapeutic strategy for this disease.

Role of ROS-induced autophagy in melanoma treatment

Autophagy as a protective role to ameliorate ROS-induced apoptosis

Several studies have reported that various anticancer agents are employed to treat melanoma through ROS production and ROS-mediated apoptosis (77,78). Kalantuboside B, a natural bufadienolide derivative, has been demonstrated to enhance intracellular ROS levels, which induce apoptosis and autophagy; moreover, apoptosis was revealed to be potentiated by the autophagy inhibitors chloroquine and 3-methyladenine in A2058 melanoma cells and xenografts, suggesting that autophagy plays a protective role in melanoma (9). Further mechanistic evaluation revealed that this natural agent triggered ROS-induced autophagy via activating the ERK signaling pathway and downregulating the calcium-dependent p53 signaling (9). Dihydromyricetin, another natural compound, was also found to promote autophagy to alleviate cell apoptosis through the ROS-NF-κB signaling pathway in human melanoma cells (79). In fact, autophagy serves as a modulator of ROS generation and contributes to cell survival under stress. Esomeprazole, a proton pump inhibitor, was revealed to induce ROS accumulation and ROS-mediated cell death by mitochondrial dysfunction and involvement of NADPH oxidase, while it also elicited early autophagy to reduce its cytotoxicity against melanoma (80). Similarly, shikonin, a botanical anticancer drug, was demonstrated to facilitate ER stress-mediated apoptosis by increasing the generation of ROS, which was accompanied by prosurvival autophagy through activation of the p38 signaling pathway in A375 melanoma cells (81). These findings indicate that autophagy represents an adaptive survival response to overcome drug-induced cellular stress and cytotoxicity. In addition, Zn(II) phthalocyanine photodynamic therapy, an emerging therapy for melanoma, causes ROS-mediated oxidative stress that further activates both apoptosis and autophagy; however, suppression of autophagy strengthens phototoxicity and prevents apoptotic cell death (82). Furthermore, combination of photodynamic therapy with the natural agent curcumin was shown to aggravate oxidative stress-mediated cell apoptosis and facilitate the formation of autophagosomes, which favor ROS-damaged melanoma cells to escape apoptosis (83). Therefore, these studies suggest that these anti-melanoma therapies induce a protective autophagy to eliminate ROS, which compromises ROS-mediated apoptosis.

Autophagic cell death exacerbates ROS-induced death

A variety of anticancer drugs mediate autophagic melanoma cell death by increasing the production of ROS. Liang et al (84) revealed that squalene synthase (SQS) III, a derivative of Schima crenata Korth. with antitumor activities, induced apoptosis and autophagic cell death in the human melanoma cell line A375. These effects were reversed by the ROS scavenger N-acetylcysteine, indicating that SQS III-induced autophagic cell death resulted from ROS generation, which was further demonstrated to be a messenger to inhibit the AKT/mTOR signaling pathway, which is a negative regulator of the autophagy process (84). These findings demonstrated that drug-mediated ROS caused melanoma cell death by regulating autophagy-related signaling pathways. A previous study has reported that dimethylacrylshikonin, isolated from the roots of Boraginaceae, triggered the loss of mitochondrial membrane potential and thus ROS accumulation in melanoma cells, leading to autophagic cell death by mitochondrial dysfunction (85). A similar study revealed that inhibition of dihydrolipoyl dehydrogenase, a mitochondrial oxidoreductase enzyme, contributed to ROS overproduction and alteration of mitochondrial energy metabolism, and thus triggered autophagic melanoma cell death (86). Therefore, ROS generation attributed to mitochondrial dysfunction appears to be a major contributor of autophagic cell death. Additionally, berberine-photodynamic therapy activated ER stress, which contributed to a marked increase in ROS, and thus enhanced autophagy to facilitate apoptosis in human melanoma cells (87). Photodynamic therapy-induced ROS accumulation was sufficient to elicit oxidative stress, which further mediated autophagy and consequently inhibited cell proliferation in B16F10 melanoma cells (10). Furthermore, the bis(phenylidenebenzeneamine)-1-disulfide (88), graveoline (89) and terfenadine (90) have been found to induce autophagic cell death in melanoma cells through increasing ROS production.

Collectively, ROS-induced autophagy elicited by anticancer agents exerts double functions in the treatment of melanoma, either via autophagic cell survival or death (Fig. 1). It has been reported that low doses of nitrogen-doped titanium dioxide, a photodynamic therapy for melanoma, stimulate a protective autophagy flux response, whereas therapeutic doses impair autophagy and induce necroptosis via ROS production (91). A possible explanation is that the functional effects of ROS-induced autophagy on cancer cell death are dependent on the type of oxidative stress, as well as the quantity and location of the ROS produced (92). It should be noted that anticancer drug-induced ROS also inhibit cytoprotective autophagy to promote apoptosis, and that blocking autophagy fails to affect ROS generation (78,89), suggesting that ROS serve as messengers to modulate autophagy and thus determine the cell fate. In this context, further clarifying the role of anticancer drug-induced ROS and the resultant autophagy is essential for melanoma treatment.

Future directions

As a ROS-driven tumor, melanoma is susceptible to ROS production and thus to oxidative stress. The main resources of ROS in melanocytes comprises UV radiation, melanin synthesis and activation of NOX. Excessive ROS generation causes oxidative damage to melanocytes, including oxidative DNA damage, which is responsible for gene mutations and epigenetic alterations, as well as aberrant activation of signaling pathways that are involved in cell proliferation and differentiation, ultimately leading to melanomagenesis and melanoma progression. Autophagy is elicited as an adaptive response to remove ROS and oxidative components, which maintains genetic stability and inhibits carcinogenesis; however, once a tumor is formed, autophagy is induced to provide metabolic demands and nutrients for tumor progression. Furthermore, the sustained activation of autophagy may contribute to cell death, which is also known as autophagic cell death (Fig. 1).

The crosstalk between ROS and autophagy in melanomagenesis remains unclear. Understanding the mechanism of ROS-regulated autophagy in melanoma will provide new therapeutic strategies for this disease. In addition, it is essential to investigate whether anticancer drug-induced autophagy promotes cell survival or facilitates cell death. Agents that are cytoprotective autophagy-inducing drugs combined with autophagy inhibitors and autophagic cell death-mediating agents combined with autophagy activators are both effective therapeutic strategies for melanoma. Therefore, clarification of the functional status of ROS-induced autophagy is crucial in melanoma treatment. Of note, prolonged autophagy causes the removal of excessive ROS and the degradation of dysfunctional mitochondria, which leads to reductive stress. This novel reductive stress mechanism of cell death offers a novel approach for cancer therapy, including transfer-hydrogenation catalysts (9395). Further studies on the role of reductive stress in melanoma may provide insights into the treatment of this disease.

Acknowledgements

Not applicable.

Funding

The present study was funded by the Traditional Chinese Medicine research projects of Heilongjiang Province (grant no. ZHY2020-041) and the Doctoral Program of Heilongjiang Province (grant no. LBH-Z21218).

Availability of data and materials

Data sharing is not applicable to this article, as no datasets were generated or analyzed during the current study.

Authors' contributions

XZ performed the literature search. XZ, HL and CL wrote the manuscript. XY planned and supervised the writing of the present review article. Data authentication is not applicable. All authors read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Saikolappan S, Kumar B, Shishodia G, Koul S and Koul HK: Reactive oxygen species and cancer: A complex interaction. Cancer Lett. 452:132–143. 2019. View Article : Google Scholar : PubMed/NCBI

2 

Sarmiento-Salinas FL, Perez-Gonzalez A, Acosta-Casique A, Ix-Ballote A, Diaz A, Treviño S, Rosas-Murrieta NH, Millán-Perez-Peña L and Maycotte P: Reactive oxygen species: Role in carcinogenesis, cancer cell signaling and tumor progression. Life Sci. 284:1199422021. View Article : Google Scholar : PubMed/NCBI

3 

Venza I, Venza M, Visalli M, Lentini G, Teti D and d'Alcontres FS: ROS as regulators of cellular processes in melanoma. Oxid Med Cell Longev. 2021:12086902021. View Article : Google Scholar : PubMed/NCBI

4 

Pizzimenti S, Ribero S, Cucci MA, Grattarola M, Monge C, Dianzani C, Barrera G and Muzio G: Oxidative stress-related mechanisms in melanoma and in the acquired resistance to targeted therapies. Antioxidants (Basel). 10:19422021. View Article : Google Scholar : PubMed/NCBI

5 

Morishita H and Mizushima N: Diverse cellular roles of autophagy. Annu Rev Cell Dev Biol. 35:453–475. 2019. View Article : Google Scholar : PubMed/NCBI

6 

Catalani E, Giovarelli M, Zecchini S, Perrotta C and Cervia D: Oxidative stress and autophagy as key targets in melanoma cell fate. Cancers (Basel). 13:57912021. View Article : Google Scholar : PubMed/NCBI

7 

Gao L, Loveless J, Shay C and Teng Y: Targeting ROS-Mediated crosstalk between autophagy and apoptosis in cancer. Adv Exp Med Biol. 1260:1–12. 2020. View Article : Google Scholar : PubMed/NCBI

8 

Fried L and Arbiser JL: The reactive oxygen-driven tumor: Relevance to melanoma. Pigment Cell Melanoma Res. 21:117–122. 2008. View Article : Google Scholar : PubMed/NCBI

9 

Hseu YC, Cho HJ, Gowrisankar YV, Thiyagarajan V, Chen XZ, Lin KY, Huang HC and Yang HL: Kalantuboside B induced apoptosis and cytoprotective autophagy in human melanoma A2058cells: An in vitro and in vivo study. Free Radic Biol Med. 143:397–411. 2019. View Article : Google Scholar : PubMed/NCBI

10 

Santos GMP, Oliveira SCPS, Monteiro JCS, Fagnani SR, Sampaio FP, Correia NA, Crugeira PJL and Pinheiro ALB: ROS-induced autophagy reduces B16F10 melanoma cell proliferative activity. Lasers Med Sci. 33:1335–1340. 2018. View Article : Google Scholar : PubMed/NCBI

11 

Slominski AT, Zmijewski MA, Plonka PM, Szaflarski JP and Paus R: How UV Light Touches the Brain and Endocrine System Through Skin, and Why. Endocrinology. 159:1992–2007. 2018. View Article : Google Scholar : PubMed/NCBI

12 

von Thaler AK, Kamenisch Y and Berneburg M: The role of ultraviolet radiation in melanomagenesis. Exp Dermatol. 19:81–88. 2010. View Article : Google Scholar : PubMed/NCBI

13 

Terra VA, Souza-Neto FP, Pereira RC, Silva TN, Costa AC, Luiz RC, Cecchini R and Cecchini AL: Time-dependent reactive species formation and oxidative stress damage in the skin after UVB irradiation. J Photochem Photobiol B. 109:34–41. 2012. View Article : Google Scholar : PubMed/NCBI

14 

Baumler W, Regensburger J, Knak A, Felgentrager A and Maisch T: UVA and endogenous photosensitizers-the detection of singlet oxygen by its luminescence. Photochem Photobiol Sci. 11:107–117. 2012. View Article : Google Scholar : PubMed/NCBI

15 

Knak A, Regensburger J, Maisch T and Baumler W: Exposure of vitamins to UVB and UVA radiation generates singlet oxygen. Photochem Photobiol Sci. 13:820–829. 2014. View Article : Google Scholar : PubMed/NCBI

16 

Regensburger J, Maisch T, Knak A, Gollmer A, Felgentraeger A, Lehner K and Baeumler W: UVA irradiation of fatty acids and their oxidized products substantially increases their ability to generate singlet oxygen. Phys Chem Chem Phys. 15:17672–17680. 2013. View Article : Google Scholar : PubMed/NCBI

17 

Baier J, Maisch T, Maier M, Engel E, Landthaler M and Baumler W: Singlet oxygen generation by UVA light exposure of endogenous photosensitizers. Biophys J. 91:1452–1459. 2006. View Article : Google Scholar : PubMed/NCBI

18 

Regensburger J, Knak A, Maisch T, Landthaler M and Baumler W: Fatty acids and vitamins generate singlet oxygen under UVB irradiation. Exp Dermatol. 21:135–139. 2012. View Article : Google Scholar : PubMed/NCBI

19 

Valencia A and Kochevar IE: Nox1-based NADPH oxidase is the major source of UVA-induced reactive oxygen species in human keratinocytes. J Invest Dermatol. 128:214–222. 2008. View Article : Google Scholar : PubMed/NCBI

20 

Zhao B, Shah P, Qiang L, He TC, Budanov A and He YY: Distinct Role of Sesn2 in Response to UVB-Induced DNA Damage and UVA-Induced Oxidative Stress in Melanocytes. Photochem Photobiol. 93:375–381. 2017. View Article : Google Scholar : PubMed/NCBI

21 

Cadet J, Douki T and Ravanat JL: Oxidatively generated damage to cellular DNA by UVB and UVA radiation. Photochem Photobiol. 91:140–155. 2015. View Article : Google Scholar : PubMed/NCBI

22 

Meyskens FL Jr, McNulty SE, Buckmeier JA, Tohidian NB, Spillane TJ, Kahlon RS and Gonzalez RI: Aberrant redox regulation in human metastatic melanoma cells compared to normal melanocytes. Free Radic Biol Med. 31:799–808. 2001. View Article : Google Scholar : PubMed/NCBI

23 

Jenkins NC and Grossman D: Role of melanin in melanocyte dysregulation of reactive oxygen species. Biomed Res Int. 2013:9087972013. View Article : Google Scholar : PubMed/NCBI

24 

Arslanbaeva LR and Santoro MM: Adaptive redox homeostasis in cutaneous melanoma. Redox Biol. 37:1017532020. View Article : Google Scholar : PubMed/NCBI

25 

Slominski RM, Sarna T, Plonka PM, Raman C, Brozyna AA and Slominski AT: Melanoma, melanin, and melanogenesis: The Yin and Yang relationship. Front Oncol. 12:8424962022. View Article : Google Scholar : PubMed/NCBI

26 

Slominski A, Tobin DJ, Shibahara S and Wortsman J: Melanin pigmentation in mammalian skin and its hormonal regulation. Physiol Rev. 84:1155–1228. 2004. View Article : Google Scholar : PubMed/NCBI

27 

Slominski A, Zmijewski MA and Pawelek J: L-tyrosine and L-dihydroxyphenylalanine as hormone-like regulators of melanocyte functions. Pigment Cell Melanoma Res. 25:14–27. 2012. View Article : Google Scholar : PubMed/NCBI

28 

Panzella L, Leone L, Greco G, Vitiello G, D'Errico G, Napolitano A and d'Ischia M: Red human hair pheomelanin is a potent pro-oxidant mediating UV-independent contributory mechanisms of melanomagenesis. Pigment Cell Melanoma Res. 27:244–252. 2014. View Article : Google Scholar : PubMed/NCBI

29 

Panzella L, Szewczyk G, d'Ischia M, Napolitano A and Sarna T: Zinc-induced structural effects enhance oxygen consumption and superoxide generation in synthetic pheomelanins on UVA/visible light irradiation. Photochem Photobiol. 86:757–764. 2010. View Article : Google Scholar : PubMed/NCBI

30 

Shain AH and Bastian BC: From melanocytes to melanomas. Nat Rev Cancer. 16:345–358. 2016. View Article : Google Scholar : PubMed/NCBI

31 

Swope VB and Abdel-Malek ZA: MC1R: Front and center in the bright side of dark eumelanin and DNA repair. Int J Mol Sci. 19:26672018. View Article : Google Scholar : PubMed/NCBI

32 

Nasti TH and Timares L: MC1R, eumelanin and pheomelanin: Their role in determining the susceptibility to skin cancer. Photochem Photobiol. 91:188–200. 2015. View Article : Google Scholar : PubMed/NCBI

33 

Brozyna AA, Jozwicki W, Roszkowski K, Filipiak J and Slominski AT: Melanin content in melanoma metastases affects the outcome of radiotherapy. Oncotarget. 7:17844–17853. 2016. View Article : Google Scholar : PubMed/NCBI

34 

Liu-Smith F, Dellinger R and Meyskens FL Jr: Updates of reactive oxygen species in melanoma etiology and progression. Arch Biochem Biophys. 563:51–55. 2014. View Article : Google Scholar : PubMed/NCBI

35 

Meierjohann S: Oxidative stress in melanocyte senescence and melanoma transformation. Eur J Cell Biol. 93:36–41. 2014. View Article : Google Scholar : PubMed/NCBI

36 

Govindarajan B, Sligh JE, Vincent BJ, Li M, Canter JA, Nickoloff BJ, Rodenburg RJ, Smeitink JA, Oberley L, Zhang Y, et al: Overexpression of Akt converts radial growth melanoma to vertical growth melanoma. J Clin Invest. 117:719–729. 2007. View Article : Google Scholar : PubMed/NCBI

37 

Ribeiro-Pereira C, Moraes JA, Souza Mde J, Laurindo FR, Arruda MA and Barja-Fidalgo C: Redox modulation of FAK controls melanoma survival-role of NOX4. PLoS One. 9:e994812014. View Article : Google Scholar : PubMed/NCBI

38 

Liu GS, Wu JC, Tsai HE, Dusting GJ, Chan EC, Wu CS and Tai MH: Proopiomelanocortin gene delivery induces apoptosis in melanoma through NADPH oxidase 4-mediated ROS generation. Free Radic Biol Med. 70:14–22. 2014. View Article : Google Scholar : PubMed/NCBI

39 

Quast SA, Berger A and Eberle J: ROS-dependent phosphorylation of Bax by wortmannin sensitizes melanoma cells for TRAIL-induced apoptosis. Cell Death Dis. 4:e8392013. View Article : Google Scholar : PubMed/NCBI

40 

Liu GS, Peshavariya H, Higuchi M, Brewer AC, Chang CW, Chan EC and Dusting GJ: Microphthalmia-associated transcription factor modulates expression of NADPH oxidase type 4: A negative regulator of melanogenesis. Free Radic Biol Med. 52:1835–1843. 2012. View Article : Google Scholar : PubMed/NCBI

41 

Yamaura M, Mitsushita J, Furuta S, Kiniwa Y, Ashida A, Goto Y, Shang WH, Kubodera M, Kato M, Takata M, et al: NADPH oxidase 4 contributes to transformation phenotype of melanoma cells by regulating G2-M cell cycle progression. Cancer Res. 69:2647–2654. 2009. View Article : Google Scholar : PubMed/NCBI

42 

Antony S, Jiang G, Wu Y, Meitzler JL, Makhlouf HR, Haines DC, Butcher D, Hoon DS, Ji J, Zhang Y, et al: NADPH oxidase 5 (NOX5)-induced reactive oxygen signaling modulates normoxic HIF-1α and p27Kip1 expression in malignant melanoma and other human tumors. Mol Carcinog. 56:2643–2662. 2017. View Article : Google Scholar : PubMed/NCBI

43 

Xian D, Lai R, Song J, Xiong X and Zhong J: Emerging Perspective: Role of Increased ROS and Redox Imbalance in Skin Carcinogenesis. Oxid Med Cell Longev. 2019:81273622019. View Article : Google Scholar : PubMed/NCBI

44 

Mouret S, Forestier A and Douki T: The specificity of UVA-induced DNA damage in human melanocytes. Photochem Photobiol Sci. 11:155–162. 2012. View Article : Google Scholar : PubMed/NCBI

45 

Murtas D, Piras F, Minerba L, Ugalde J, Floris C, Maxia C, Demurtas P, Perra MT and Sirigu P: Nuclear 8-hydroxy-2′-deoxyguanosine as survival biomarker in patients with cutaneous melanoma. Oncol Rep. 23:329–335. 2010.PubMed/NCBI

46 

Landi MT, Bauer J, Pfeiffer RM, Elder DE, Hulley B, Minghetti P, Calista D, Kanetsky PA, Pinkel D and Bastian BC: MC1R germline variants confer risk for BRAF-mutant melanoma. Science. 313:521–522. 2006. View Article : Google Scholar : PubMed/NCBI

47 

Comito G, Calvani M, Giannoni E, Bianchini F, Calorini L, Torre E, Migliore C, Giordano S and Chiarugi P: HIF-1α stabilization by mitochondrial ROS promotes Met-dependent invasive growth and vasculogenic mimicry in melanoma cells. Free Radic Biol Med. 51:893–904. 2011. View Article : Google Scholar : PubMed/NCBI

48 

Slominski A, Kim TK, Brozyna AA, Janjetovic Z, Brooks DL, Schwab LP, Skobowiat C, Jóźwicki W and Seagroves TN: The role of melanogenesis in regulation of melanoma behavior: Melanogenesis leads to stimulation of HIF-1α expression and HIF-dependent attendant pathways. Arch Biochem Biophys. 563:79–93. 2014. View Article : Google Scholar : PubMed/NCBI

49 

Park SJ, Kim YT and Jeon YJ: Antioxidant dieckol downregulates the Rac1/ROS signaling pathway and inhibits Wiskott-Aldrich syndrome protein (WASP)-family verprolin-homologous protein 2 (WAVE2)-mediated invasive migration of B16 mouse melanoma cells. Mol Cells. 33:363–369. 2012. View Article : Google Scholar : PubMed/NCBI

50 

Jenkins NC, Liu T, Cassidy P, Leachman SA, Boucher KM, Goodson AG, Samadashwily G and Grossman D: The p16(INK4A) tumor suppressor regulates cellular oxidative stress. Oncogene. 30:265–274. 2011. View Article : Google Scholar : PubMed/NCBI

51 

Kazimierczak U, Dondajewska E, Zajaczkowska M, Karwacka M, Kolenda T and Mackiewicz A: LATS1 Is a Mediator of Melanogenesis in Response to Oxidative Stress and Regulator of Melanoma Growth. Int J Mol Sci. 22:31082021. View Article : Google Scholar : PubMed/NCBI

52 

Molognoni F, de Melo FH, da Silva CT and Jasiulionis MG: Ras and Rac1, frequently mutated in melanomas, are activated by superoxide anion, modulate Dnmt1 level and are causally related to melanocyte malignant transformation. PLoS One. 8:e819372013. View Article : Google Scholar : PubMed/NCBI

53 

Lingappan K: NF-κB in Oxidative Stress. Curr Opin Toxicol. 7:81–86. 2018. View Article : Google Scholar : PubMed/NCBI

54 

Dolcet X, Llobet D, Pallares J and Matias-Guiu X: NF-kB in development and progression of human cancer. Virchows Arch. 446:475–482. 2005. View Article : Google Scholar : PubMed/NCBI

55 

Slominski A and Wortsman J: Neuroendocrinology of the skin. Endocr Rev. 21:457–487. 2000. View Article : Google Scholar : PubMed/NCBI

56 

Slominski RM, Zmijewski MA and Slominski AT: The role of melanin pigment in melanoma. Exp Dermatol. 24:258–259. 2015. View Article : Google Scholar : PubMed/NCBI

57 

Li W, Slominski R and Slominski AT: High-resolution magic angle spinning nuclear magnetic resonance analysis of metabolic changes in melanoma cells after induction of melanogenesis. Anal Biochem. 386:282–284. 2009. View Article : Google Scholar : PubMed/NCBI

58 

Kamenisch Y, Ivanova I, Drexler K and Berneburg M: UVA, metabolism and melanoma: UVA makes melanoma hungry for metastasis. Exp Dermatol. 27:941–949. 2018. View Article : Google Scholar : PubMed/NCBI

59 

Slominski A, Paus R and Mihm MC: Inhibition of melanogenesis as an adjuvant strategy in the treatment of melanotic melanomas: Selective review and hypothesis. Anticancer Res. 18((5B)): 3709–3715. 1998.PubMed/NCBI

60 

Slominski A, Zbytek B and Slominski R: Inhibitors of melanogenesis increase toxicity of cyclophosphamide and lymphocytes against melanoma cells. Int J Cancer. 124:1470–1477. 2009. View Article : Google Scholar : PubMed/NCBI

61 

Kuo CL, Chou HY, Chiu YC, Cheng AN, Fan CC, Chang YN, Chen CH, Jiang SS, Chen NJ and Lee AY: Mitochondrial oxidative stress by Lon-PYCR1 maintains an immunosuppressive tumor microenvironment that promotes cancer progression and metastasis. Cancer Lett. 474:138–150. 2020. View Article : Google Scholar : PubMed/NCBI

62 

Li D, Ding Z, Du K, Ye X and Cheng S: Reactive oxygen species as a link between antioxidant pathways and autophagy. Oxid Med Cell Longev. 2021:55832152021.PubMed/NCBI

63 

Poillet-Perez L, Despouy G, Delage-Mourroux R and Boyer-Guittaut M: Interplay between ROS and autophagy in cancer cells, from tumor initiation to cancer therapy. Redox Biol. 4:184–192. 2015. View Article : Google Scholar : PubMed/NCBI

64 

Scherz-Shouval R, Shvets E, Fass E, Shorer H, Gil L and Elazar Z: Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. EMBO J. 38:e1018122019. View Article : Google Scholar : PubMed/NCBI

65 

Mathew R, Karp CM, Beaudoin B, Vuong N, Chen G, Chen HY, Bray K, Reddy A, Bhanot G, Gelinas C, et al: Autophagy suppresses tumorigenesis through elimination of p62. Cell. 137:1062–1075. 2009. View Article : Google Scholar : PubMed/NCBI

66 

Jain A, Lamark T, Sjøttem E, Larsen KB, Awuh JA, Øvervatn A, McMahon M, Hayes JD and Johansen T: p62/SQSTM1 is a target gene for transcription factor NRF2 and creates a positive feedback loop by inducing antioxidant response element-driven gene transcription. J Biol Chem. 285:22576–22591. 2010. View Article : Google Scholar : PubMed/NCBI

67 

Redza-Dutordoir M and Averill-Bates DA: Interactions between reactive oxygen species and autophagy: Special issue: Death mechanisms in cellular homeostasis. Biochim Biophys Acta Mol Cell Res. 1868:1190412021. View Article : Google Scholar : PubMed/NCBI

68 

Sample A, Zhao B, Wu C, Qian S, Shi X, Aplin A and He YY: The autophagy receptor adaptor p62 is Up-regulated by UVA radiation in melanocytes and in melanoma cells. Photochem Photobiol. 94:432–437. 2018. View Article : Google Scholar : PubMed/NCBI

69 

Xie X, Koh JY, Price S, White E and Mehnert JM: Atg7 overcomes senescence and promotes growth of BrafV600E-Driven melanoma. Cancer Discov. 5:410–423. 2015. View Article : Google Scholar : PubMed/NCBI

70 

Zhao Y, Wang W, Min I, Wyrwas B, Moore M, Zarnegar R and Fahey TJ III: BRAF V600E-dependent role of autophagy in uveal melanoma. J Cancer Res Clin Oncol. 143:447–455. 2017. View Article : Google Scholar : PubMed/NCBI

71 

Wang Y, Wang Y, Wu J, Wang W and Zhang Y: Oxygen partial pressure plays a crucial role in B16 melanoma cell survival by regulating autophagy and mitochondrial functions. Biochem Biophys Res Commun. 510:643–648. 2019. View Article : Google Scholar : PubMed/NCBI

72 

Rouschop KM, Ramaekers CH, Schaaf MB, Keulers TG, Savelkouls KG, Lambin P, Koritzinsky M and Wouters BG: Autophagy is required during cycling hypoxia to lower production of reactive oxygen species. Radiother Oncol. 92:411–416. 2009. View Article : Google Scholar : PubMed/NCBI

73 

Ma XH, Piao SF, Dey S, McAfee Q, Karakousis G, Villanueva J, Hart LS, Levi S, Hu J, Zhang G, et al: Targeting ER stress-induced autophagy overcomes BRAF inhibitor resistance in melanoma. J Clin Invest. 124:1406–1417. 2014. View Article : Google Scholar : PubMed/NCBI

74 

Victor P, Sarada D and Ramkumar KM: Crosstalk between endoplasmic reticulum stress and oxidative stress: Focus on protein disulfide isomerase and endoplasmic reticulum oxidase 1. Eur J Pharmacol. 892:1737492021. View Article : Google Scholar : PubMed/NCBI

75 

Santos CX, Tanaka LY, Wosniak J and Laurindo FR: Mechanisms and implications of reactive oxygen species generation during the unfolded protein response: Roles of endoplasmic reticulum oxidoreductases, mitochondrial electron transport, and NADPH oxidase. Antioxid Redox Signal. 11:2409–2427. 2009. View Article : Google Scholar : PubMed/NCBI

76 

Corazzari M, Rapino F, Ciccosanti F, Giglio P, Antonioli M, Conti B, Fimia GM, Lovat PE and Piacentini M: Oncogenic BRAF induces chronic ER stress condition resulting in increased basal autophagy and apoptotic resistance of cutaneous melanoma. Cell Death Differ. 22:946–958. 2015. View Article : Google Scholar : PubMed/NCBI

77 

Ghaemi B, Moshiri A, Herrmann IK, Hajipour MJ, Wick P, Amani A and Kharrazi S: Supramolecular insights into domino effects of simpleAg@ZnO-Induced oxidative stress in melanoma cancer cells. ACS Appl Mater Interfaces. 11:46408–46418. 2019. View Article : Google Scholar : PubMed/NCBI

78 

Chang SN, Khan I, Kim CG, Park SM, Choi DK, Lee H, Hwang BS, Kang SC and Park JG: Decursinol angelate arrest melanoma cell proliferation by initiating cell death and tumor shrinkage via induction of apoptosis. Int J Mol Sci. 22:40962021. View Article : Google Scholar : PubMed/NCBI

79 

Zhou DZ, Sun HY, Yue JQ, Peng Y, Chen YM and Zhong ZJ: Dihydromyricetin induces apoptosis and cytoprotective autophagy through ROS-NF-κB signalling in human melanoma cells. Free Radic Res. 51:517–528. 2017. View Article : Google Scholar : PubMed/NCBI

80 

Marino ML, Fais S, Djavaheri-Mergny M, Villa A, Meschini S, Lozupone F, Venturi G, Della Mina P, Pattingre S, Rivoltini L, et al: Proton pump inhibition induces autophagy as a survival mechanism following oxidative stress in human melanoma cells. Cell Death Dis. 1:e872010. View Article : Google Scholar : PubMed/NCBI

81 

Liu Y, Kang X, Niu G, He S, Zhang T, Bai Y, Li Y, Hao H, Chen C, Shou Z and Li B: Shikonin induces apoptosis and prosurvival autophagy in human melanoma A375 cells via ROS-mediated ER stress and p38 pathways. Artif Cells Nanomed Biotechnol. 47:626–635. 2019. View Article : Google Scholar : PubMed/NCBI

82 

Valli F, García Vior MC, Roguin LP and Marino J: Crosstalk between oxidative stress-induced apoptotic and autophagic signaling pathways in Zn(II) phthalocyanine photodynamic therapy of melanoma. Free Radic Biol Med. 152:743–754. 2020. View Article : Google Scholar : PubMed/NCBI

83 

Niu T, Tian Y, Mei Z and Guo G: Inhibition of autophagy enhances curcumin united light irradiation-induced oxidative stress and tumor growth suppression in human melanoma cells. Sci Rep. 6:313832016. View Article : Google Scholar : PubMed/NCBI

84 

Liang QP, Xu TQ, Liu BL, Lei XP, Hambrook JR, Zhang DM and Zhou GX: Sasanquasaponin III from Schima crenata Korth induces autophagy through Akt/mTOR/p70S6K pathway and promotes apoptosis in human melanoma A375 cells. Phytomedicine. 58:1527692019. View Article : Google Scholar : PubMed/NCBI

85 

Kretschmer N, Deutsch A, Durchschein C, Rinner B, Stallinger A, Higareda-Almaraz JC, Scheideler M, Lohberger B and Bauer R: Comparative gene expression analysis in WM164 melanoma cells revealed that β-β-Dimethylacrylshikonin Leads to ROS generation, loss of mitochondrial membrane potential, and autophagy induction. Molecules. 23:28232018. View Article : Google Scholar : PubMed/NCBI

86 

Yumnam S, Kang MC, Oh SH, Kwon HC, Kim JC, Jung ES, Lee CH, Lee AY, Hwang JI and Kim SY: Downregulation of dihydrolipoyl dehydrogenase by UVA suppresses melanoma progression via triggering oxidative stress and altering energy metabolism. Free Radic Biol Med. 162:77–87. 2021. View Article : Google Scholar : PubMed/NCBI

87 

Fang J, Huang X, Yang Y, Wang X, Liang X and Liu J: Berberine-photodynamic induced apoptosis by activating endoplasmic reticulum stress-autophagy pathway involving CHOP in human malignant melanoma cells. Biochem Biophys Res Commun. 552:183–190. 2021. View Article : Google Scholar : PubMed/NCBI

88 

Hu WP, Hsu CC, Wang YC, Senadi GC, Kuo KK, Jen JF and Wang JJ: Bis(phenylidenebenzeneamine)-1-disulfide derivatives induce autophagy in melanoma cells through a mitochondria-mediated pathway. Anticancer Res. 35:6075–6080. 2015.PubMed/NCBI

89 

Ghosh S, Bishayee K and Khuda-Bukhsh AR: Graveoline isolated from ethanolic extract of Ruta graveolens triggers apoptosis and autophagy in skin melanoma cells: A novel apoptosis-independent autophagic signaling pathway. Phytother Res. 28:1153–1162. 2014. View Article : Google Scholar : PubMed/NCBI

90 

Nicolau-Galmés F, Asumendi A, Alonso-Tejerina E, Pérez-Yarza G, Jangi SM, Gardeazabal J, Arroyo-Berdugo Y, Careaga JM, Díaz-Ramón JL, Apraiz A and Boyano MD: Terfenadine induces apoptosis and autophagy in melanoma cells through ROS-dependent and -independent mechanisms. Apoptosis. 16:1253–1267. 2011. View Article : Google Scholar : PubMed/NCBI

91 

Mohammadalipour Z, Rahmati M, Khataee A and Moosavi MA: Differential effects of N-TiO2 nanoparticle and its photo-activated form on autophagy and necroptosis in human melanoma A375 cells. J Cell Physiol. 235:8246–8259. 2020. View Article : Google Scholar : PubMed/NCBI

92 

Yun HR, Jo YH, Kim J, Shin Y, Kim SS and Choi TG: Roles of autophagy in oxidative stress. Int J Mol Sci. 21:32892020. View Article : Google Scholar : PubMed/NCBI

93 

Soldevila-Barreda JJ, Romero-Canelón I, Habtemariam A and Sadler PJ: Transfer hydrogenation catalysis in cells as a new approach to anticancer drug design. Nat Commun. 6:65822015. View Article : Google Scholar : PubMed/NCBI

94 

Sopha P, Ren HY, Grove DE and Cyr DM: Endoplasmic reticulum stress-induced degradation of DNAJB12 stimulates BOK accumulation and primes cancer cells for apoptosis. J Biol Chem. 292:11792–11803. 2017. View Article : Google Scholar : PubMed/NCBI

95 

Coverdale JPC, Romero-Canelon I, Sanchez-Cano C, Clarkson GJ, Habtemariam A, Wills M and Sadler PJ: Asymmetric transfer hydrogenation by synthetic catalysts in cancer cells. Nat Chem. 10:347–354. 2018. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

October-2022
Volume 26 Issue 4

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Zhang X, Li H, Liu C and Yuan X: Role of ROS‑mediated autophagy in melanoma (Review). Mol Med Rep 26: 303, 2022
APA
Zhang, X., Li, H., Liu, C., & Yuan, X. (2022). Role of ROS‑mediated autophagy in melanoma (Review). Molecular Medicine Reports, 26, 303. https://doi.org/10.3892/mmr.2022.12819
MLA
Zhang, X., Li, H., Liu, C., Yuan, X."Role of ROS‑mediated autophagy in melanoma (Review)". Molecular Medicine Reports 26.4 (2022): 303.
Chicago
Zhang, X., Li, H., Liu, C., Yuan, X."Role of ROS‑mediated autophagy in melanoma (Review)". Molecular Medicine Reports 26, no. 4 (2022): 303. https://doi.org/10.3892/mmr.2022.12819