Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 May 4.
Published in final edited form as: J Invest Dermatol. 2014 Feb 27;134(6):1512–1518. doi: 10.1038/jid.2014.65

Melanocytes as Instigators and Victims of Oxidative Stress

L Denat 1,, AL Kadekaro 2,, L Marrot 1, S Leachman 3, ZA Abdel-Malek 2,*
PMCID: PMC4418514  NIHMSID: NIHMS561714  PMID: 24573173

Abstract

Epidermal melanocytes are particularly vulnerable to oxidative stress due to the pro-oxidant state generated during melanin synthesis, and to intrinsic antioxidant defences that are compromised in pathologic conditions. Melanoma is thought to be oxidative stress-driven, and melanocyte death in vitiligo is thought to be instigated by a highly pro-oxidant state in the epidermis. We review the current knowledge about melanin and the redox state of melanocytes, how paracrine factors help counteract oxidative stress, the role of oxidative stress in melanoma initiation and progression and in melanocyte death in vitiligo, and how this knowledge can be harnessed for melanoma and vitiligo treatment.

Introduction

Oxidative stress results from overproduction of pro-oxidant species in cells, and/or reduction of cellular antioxidant capacity, and can damage nucleic acids, lipids, and proteins, leading to mutagenesis or cell death (Sander et al., 2004). Reactive oxygen species (ROS) are produced by mitochondria and peroxisomes during normal cellular metabolic processes. The ROS production may be accentuated under pathologic conditions, such as inflammation and cancer, as well as, upon exposure to exogenous factors, such as ultraviolet rays (UV), or chemicals (Klaunig and Kamendulis, 2004; Klaunig et al., 2009; Sander et al., 2004; Zhang et al., 1997). Skin is the largest organ that interfaces with the environment, and a major source of ROS that are induced by sun exposure. Epidermal melanocytes are particularly vulnerable to excessive ROS production due to their specialized function: melanin synthesis, which is stimulated by sun exposure, during the process of tanning, and by inflammation that results in postinflammatory hyperpigmentation (Figure 1). Oxidative stress can disrupt the homeostasis of melanocytes, compromising their survival or leading to their malignant transformation (Casp et al., 2002a; Fried and Arbiser, 2008; Gavalas et al., 2006; Govindarajan et al., 2002; Guan et al., 2008; Picardo et al., 1996a; Schallreuter et al., 1999).

Figure 1.

Figure 1

Induction of ROS by endogenous and exogenous sources and antioxidant defences that restore normal redox state in melanocytes.

Melanin and the redox state of melanocytes

Melanin synthesis involves oxidation reactions and superoxide anion (O2) and hydrogen peroxide (H2O2) generation, which subject melanocytes to oxidative stress (Koga et al., 1992; Simon et al., 2009). Confinement of melanin synthesis to melanosomes protects other cellular components from oxidative damage. Tyrosinase, the rate-limiting enzyme for melanin synthesis, oxidizes tyrosine to dopa, and dopa to dopaquinone, a specific orthoquinone that can react with nucleophilic compounds such as thiols or amino groups. The catalytic activity of tyrosinase results in the generation of O2 (Koga et al., 1992; Tomita et al., 1984). Dopaquinone is converted into dopachrome through a redox exchange. After spontaneous decarboxylation, dopachrome generates either dihydroxyindole (5,6-DHI), which is oxidized into indole quinone, or produces dihydroxyindole carboxylic acid (5,6-DHICA) after tautomerisation by tyrosinase-related protein 2 (TRP2), and 5,6-DHICA is then converted into the corresponding quinone. Moreover, TRP2 protects against oxidative stress by increasing glutathione levels, and reducing the toxicity of quinones and DNA damage induced by free radicals (Michard et al., 2008). The redox cycling from indoles to quinones generates ROS (Nappi and Vass, 1996). Polymerization of these reactive quinones finally leads to the formation of the brown/black eumelanin. The red-yellow pheomelanin differs from eumelanin in that it has a higher ratio of sulfur to quinones, and its synthesis involves the generation of cysteinyl-dopa (instead of dopa), which is converted into benzothiazine derivatives. These differences account for the higher sunlight-induced pro-oxidant property of pheomelanin compared to eumelanin.

In the skin, the balance between the pro- and antioxidant properties of melanin are mainly determined by the relative eumelanin and pheomelanin contents, the levels of melanin intermediates, the concentrations of reactive metals within the melanosome microenvironment (Di Donato et al., 2002; Liu et al., 2005). There are conflicting reports about the role of melanin or melanin intermediates as pro- or antioxidants. Constitutive pigmentation is reported to correlate directly with catalase activity in cultured human melanocytes, and with the levels of thioredoxin reductase in human skin (Maresca et al., 2008). Generation of H2O2 in response to UV correlates inversely with constitutive pigmentation, suggesting an anti-oxidant effect of melanin(Song et al., 2009). In comparison to keratinocytes, the induction of 8-hydroxydeoxyguanosine (8-OHdG), a major form of oxidative DNA damage, and expression of several base-excision repair (BER) genes are higher in melanocytes (Mouret et al., 2012). Paradoxically, cultured human melanocytes with high melanin content are reported to be more vulnerable to UVA-induced, but less susceptible to hydrogen peroxide-induced oxidative DNA damage than their counterparts with low melanin content (Hoogduijn et al., 2004; Wang et al., 2010). Stimulation of melanogenesis in human melanocytes or mouse melanoma cells is reported to increase UVA-induced DNA damage (Kvam and Tyrrell, 1999; Marrot et al., 1999; Wenczl et al., 1998). In contrast, stimulation of melanogenesis in cultured human melanocytes by α-melanocortin (α-MSH) increases the activity and protein levels of catalase, and markedly reduces UV-induced H2O2 generation (Maresca et al., 2008; Song et al., 2009). In human melanoma cells, increased pigmentation protects against UV- or hydrogen peroxide- induced mitochondrial DNA damage (Swalwell et al., 2011). The controversy about the pro-oxidant versus the antioxidant effects of melanin and its intermediates is fuelled by reports using purified melanin or melanin intermediates exogenously added to cultured cells or naked DNA (Kipp and Young, 1999; Kovacs et al., 2012; Tomita et al., 1984). Although these data support the oxidative nature of melanin, the experimental conditions used are unlikely to be physiologically relevant, since melanin is normally confined in melanosomes.

Activation of antioxidant defenses in melanocytes by paracrine factors

The homeostasis of epidermal human melanocytes is maintained primarily by a complex paracrine network consisting of growth factors and cytokines synthesized by epidermal keratinocytes and dermal fibroblasts, and modulated by UV. The keratinocyte-derived endothelin-1 is a potent mitogen and melanogenic factor that reduces H2O2 generation and apoptosis in UV-irradiated human melanocytes (Imokawa et al., 1992; Kadekaro et al., 2005; Tada et al., 1998). The melanocortins α-MSH and adrenocorticotropic hormone (ACTH) are synthesized by keratinocytes and melanocytes, and stimulate eumelanin synthesis as well as melanocyte survival and proliferation by binding and activating the melanocortin 1 receptor (MC1R). The MC1R is a Gs protein-coupled receptor expressed on the cell surface of melanocytes. Treatment of cultured human melanocytes with α-MSH results in rapid reduction in the generation of H2O2 in response to UV exposure, consistent with earlier findings by Haycock et al. (Haycock et al., 2000; Kadekaro et al., 2005; Kadekaro et al., 2010; Song et al., 2009). Additionally, α-MSH increases the protein and activity levels of catalase, and counteracts the inhibitory effect of UV on this enzyme (Song et al., 2009). Subsequently, treatment with α-MSH reduces the induction of 8-oxodG and enhances its repair in UV-irradiated melanocytes, and also reduces oxidative DNA damage induced by H2O2 (Kadekaro et al., 2012; Song et al., 2009). The antioxidant effects of α-MSH require binding and activation of MC1R, are absent in melanocytes expressing loss-of-function MC1R, and are inhibited by agouti signaling protein, the physiological MC1R antagonist (Song et al., 2009). These results establish the significance of the activated MC1R in protection of melanocytes from oxidative stress.

Activation of p53 is an important mechanism by which the activated MC1R reduces oxidative stress in melanocytes. It is noteworthy that p53 regulates pigmentation by increasing the expression of tyrosinase in human melanocytes, and pro-opiomelanocortin, the precursor for melanocortins, in mouse keratinocytes (Cui et al., 2007). Activation of the MC1R by α-MSH binding augments the UV-induced accumulation of p53 in human melanocytes by increasing phosphorylation of p53 on Ser15. Treatment with α-MSH also increases the levels of the BER enzymes OGG1 and APE-1 by a p53-dependent mechanism (Kadekaro et al., 2012).

Additionally, activation of MC1R by α-MSH regulates intracellular redox status by up regulating the expression of antioxidant genes, including heme oxygenase-1 (HO-1), ferritin, and peroxiredoxin-1 (Kadekaro et al., 2010; Song et al., 2009). α-MSH activates a number of transcription factors known to regulate the redox state of melanocytes. In normal human melanocytes and melanoma cells, the redox sensor APE-1 is a target of Mitf, the master regulator of melanocyte survival and function (Liu et al., 2009). Treatment of human melanocytes with α-MSH up regulates Mitf as well as APE-1 (Kadekaro et al., 2012; Kadekaro et al., 2005). Melanocytes also express Nrf-2, an important transcription factor that up regulates the expression of genes for phase II detoxification enzymes, and its main target HO-1 (Jain et al., 2010; Jian et al., 2011; Kaspar et al., 2009; Marrot et al., 2008; Taguchi et al., 2011). Additionally, α-MSH increases the expression of Nrf-2 gene and its target genes HO-1, γ-glutamylcysteine-synthetase, and glutathione-S-transferase Pi in cultured human melanocytes, and abrogates the inhibitory effects of UV on Nrf-2 and its targets (Kokot et al., 2009). Another transcription factor that is activated by α-MSH is NFκB, known to be activated by TNF-α and ROS (Haycock et al., 2000; Ichiyama et al., 1999; Manna and Aggarwal, 1998). Treatment of melanocytes with α-MSH inhibits UV-induced apoptosis by increasing the protein levels of Bcl2, a known target of NFκB and Mitf (Bohm et al., 2005; Kadekaro et al., 2005).

Significance of oxidative stress in melanoma

Sunlight is a major inducer of ROS formation in the skin, and a major contributor to skin cancer (Sander et al., 2004). Irradiation of the skin by UVA and/or UVB impairs natural antioxidant defenses, and induces high levels of ROS. Acute exposure to UV is a main etiological factor for melanomagenesis. Irradiation of cultured human melanocytes with UV (75% UVB, 25%UVA) results in rapid dose-dependent generation of H2O2 (van der Kemp et al., 2002)(van der Kemp et al., 2002), and subsequent decrease in catalase activity and protein levels, and reduced HO-1 expression (Kadekaro et al., 2012; Kadekaro et al., 2005; Kadekaro et al., 2010; Kokot et al., 2009; Song et al., 2009). Exposure of human OGG1 protein, an important BER enzyme, to UVB results in its inactivation (van der Kemp et al., 2002).

There is increasing evidence for the significance of oxidative stress in initiation and progression of melanoma. The role of oxidative stress in melanoma is supported by the findings that mutations in several melanoma-associated genes result from, or exacerbate, oxidative stress. The activating V600EBRAF mutation, a somatic mutation commonly expressed in nevi and melanoma, may be oxidative stress-induced (Landi et al., 2006). In melanocytes, p16 is an important regulator of oxidative stress, and its depletion in cultured human melanocytes significantly increases ROS levels (Jenkins et al., 2011). Melanocytes are more sensitive to p16 depletion than either keratinocytes or fibroblasts, which may impart the association of p16 mutations with melanoma. Loss of PTEN is associated with melanoma progression, presumably due to increased superoxide anion resulting from sustained activation of Akt (Govindarajan et al., 2007). Loss-of-function alleles of the MC1R that are associated with increased melanoma risk cause sustained oxidative stress in human melanocytes due to inability to respond to α-MSH (Kadekaro et al., 2010). In addition, oxidative stress can impair nucleotide excision repair, the main repair pathway for UV-induced DNA photoproducts, via lipid peroxidation products that inactivate DNA repair enzymes (Feng et al., 2006; Feng et al., 2004). Null polymorphisms of GSTM1 and GSTT1 that belong to the glutathione S-transferase family of antioxidant genes, have been associated with high risk of melanoma, especially in subjects with history of sunburns in childhood (Fortes et al., 2011). One SNP in the glutathione S-transferase gene GSTP1, which reduces the activity of the enzyme, has been associated with melanoma susceptibility, and with further increase in melanoma risk when co-expressed with MC1R variant alleles (Ibarrola-Villava et al., 2012). These results strongly suggest that oxidative stress is a driver of melanomagenesis (Cassidy et al., 2013).

There is increasing evidence for aberrant redox state in melanoma. Melanocytes derived from melanoma patients display increased sensitivity to oxidizing agents due to endogenous antioxidant imbalance (Grammatico et al., 1998; Meyskens et al., 2001; Picardo et al., 1996b; Picardo et al., 1999). Melanoma tumor cells have higher intracellular levels of O2 compared to normal melanocytes, and aberrantly activate the transcription factors NF-κB and AP-1 (Meyskens et al., 2001). Moreover, melanoma tumor cells express higher levels of neuronal nitric oxide synthase, thus generate higher levels of nitric oxide than normal melanocytes, and this increase correlates with the disease stage in melanoma (Yang et al., 2013). The significance of oxidative stress in melanoma is further supported by the finding that the antioxidant N-acetylcysteine inhibits tumor formation in the HGF-survivin melanoma mouse model (Cotter et al., 2007), and selective inhibitors of neuronal nitric oxide synthase inhibit melanoma cell growth and metastatic potential (Yang et al., 2013). Accordingly, antioxidants are being considered for prevention, as well as treatment of melanoma.

The association of aberrant melanin synthesis with oxidative stress and melanoma has been investigated by several research teams. Dysplastic nevi that are precursors for melanoma have increased ROS, and high pheomelanin, sulfur, iron, and calcium levels, and DNA damage (Pavel et al., 2004; Salopek et al., 1991; Smit et al., 2008). Noonan et al.(2012) reported that frequency of UVA-induced melanoma tumors in HGF mice increases with skin pigmentation via an oxidative process involving melanin photoreactivity (Noonan et al., 2012). Conversely, tumor formation in HGF mice is inhibited by the antioxidant N-acetylcysteine (Cotter et al., 2007). In human skin, UVA-induced pigmentation was found to lack photoprotective properties (Miyamura et al., 2011), indicating that exposure to UVA (e.g. in tanning beds) is not a safe practice. Recently, Mitra et al.(2012) observed that recessive yellow mice, with loss of function mc1r, and co-expressing activating BRAFv600E mutation develop more invasive melanoma tumors than their albino counterparts, and that pheomelanin results in oxidative DNA damage (Mitra et al., 2012). They concluded that oxidative DNA damage resulting from pheomelanin synthesis is causal for melanoma, independently of UV exposure. How these findings apply to human pigmentation and melanoma deserves to be investigated, since human melanocytes synthesize both, eumelanin and pheomelanin, unlike recessive yellow mouse melanocytes that only synthesize pheomelanin. The ratio of these pigments should determine the overall effects on the redox state of melanocytes particularly upon UV exposure. Also, eumelanin and pheomelanin and their intermediates might differ chemically in human vs. mouse melanocytes, which might impact their pro- or antioxidant properties. Given that eumelanin is a scavenger of ROS (Meredith and Sarna, 2006), it can be concluded that reduction of eumelanin, as in individuals with fair skin, or absence of eumelanin, as in recessive yellow mice, potentiates melanoma risk by increasing the vulnerability of melanocytes to oxidative stress.

Oxidative stress and loss of melanocytes in vitiligo

Vitiligo is a depigmentary disease that occurs in approximately 0.5% of the world population, and is characterized by loss of melanocytes in the epidermis by an autoimmune mechanism (Spritz, 2013; Taieb and Picardo, 2007). However, there is strong evidence for the role of oxidative stress as a key factor in the onset and progression of the disease. Increased sensitivity of melanocytes from vitiligo patients to UVB-induced cell death as compared to normal melanocytes was attributed to their compromised capacity to cope with increased oxidative stress (Jimbow et al., 2001). Further evidence supports the exaggerated sensitivity of melanocytes from non-lesional vitiligo skin to chemical or physical oxidative stress (Boissy and Manga, 2004; Maresca et al., 1997). Vitiligo patients are known to have very high levels of H2O2 (1 mM) and peroxynitrite in their epidermis, concomitant with reduced levels and activity of catalase, which affects the immune response (Maresca et al., 1997; Schallreuter et al., 1999; Schallreuter et al., 2012; Schallreuter et al., 1991). High levels of H2O2 inactivate and reduce the levels of methionine sulfoxide reductase (MSR) A and B, and thioredoxin/thioredoxin reductase, thus contributing to oxidative stress and melanocyte death in vitiligo (Schallreuter et al., 2008; Zhou et al., 2009). Also, high levels of H2O2 in the epidermis are found to oxidize proopiomelanocortin-derived bioactive peptides ACTH and α-MSH, both of which have antioxidant and survival effects on human melanocytes, and this effect can be mitigated by treatment with pseudocatalase (Kadekaro et al., 2005; Kadekaro et al., 2010; Spencer et al., 2007). These findings suggest that the pro-oxidant state of vitiligo skin is causal for melanocyte death.

The transcription factor Nrf-2 is implicated in the pathogenesis of vitiligo. An allelic variant of the Nrf-2 gene, A−650, is thought to be a risk factor for vitiligo (Guan et al., 2008). More recently, Natarajan et al.(2010) reported increased transcript levels of Nrf-2, as well as its targets NQO-1, γ-glutamyl cysteine ligase catalytic and modulatory subunits (GCLC and GCLM, respectively) in vitiligo lesional epidermis, as compared to non-lesional skin (Natarajan et al., 2010). However, induction of Nrf-2, and its target genes HO-1, NQO-1, GCLC, and GCLM by the electrophilic compounds curcumin and santalol is evident in non-lesional, but not in lesional vitiligo skin, further confirming the disruption of redox homeostasis in vitiligo (Natarajan et al., 2010). Treatment of vitiligo patients with PUVA increases the expression of the Nrf-2 target HO-1 in the skin (Elassiuty et al., 2011). Comparison of cultured non-lesional vitiligo melanocytes to their normal counterpart shows that the former exhibit greater induction of HO-1 than the latter in response to exposure to UVA or the phenolic compound 4-Tertuary butylphenol, demonstrating increased sensitivity of vitiligo-derived melanocytes to oxidative stress.

In vitiligo, oxidative stress-induced death of melanocytes is exacerbated by abnormal levels and/or activities of other antioxidant and BER enzymes. Catalase allelic variants have been associated with vitiligo, and the levels of several antioxidant enzymes, such as catalase, glutathione peroxidase, and glutathione reductase have been found to be altered in vitiligo, which account for sustained high levels of hydrogen peroxide in the epidermis (Casp et al., 2002b; Gavalas et al., 2006; Park et al., 2006). Salem et al. (2009) showed that in both lesional and non-lesional vitiligo skin, the levels of the BER enzymes OGG1, APE1, and DNA polymerase β are increased (Salem et al., 2009). In addition to high levels of hydrogen peroxide, high levels of inducible nitric oxide synthase (iNos) in lesional and non-lesional skin, and increased 8-oxoG in the skin and plasma of vitiligo patients, can be detected, further indicating generalized oxidative stress in vitiligo.

Targeting oxidative stress pathways for treatment of melanoma and vitiligo

A major benefit to understanding redox-related mechanisms occurring in healthy and diseased melanocytes is the capacity to harness these pathways for effective, targeted therapies and prevention measures. Repigmentation of depigmented skin of Vitiligo, characterized by high levels of epidermal hydrogen peroxide and peroxynitrite, is achieved by reducing hydrogen peroxide, such as with application of narrow band UVB-activated pseudocatalase (Schallreuter et al., 2013). For treatment of melanoma, characterized by aberrant redox state, two different strategies were proposed (Fruehauf and Meyskens, 2007). The first strategy is to use agents that increase ROS scavenging to reduce melanoma tumor growth via inhibiting hydrogen peroxide signaling, which mediates the proliferative effects of growth factors and inhibits the activity of protein tyrosine phosphatases, such as PTEN. Over expression of superoxide dismutase, glutathione peroxidase, or catalase reduces tumor cell growth (Finch et al., 2006; Liu et al., 2006; Venkataraman et al., 2005). There is increasing evidence for the efficacy of antioxidants as chemopreventative agents that inhibit melanoma onset and progression. Administration of the antioxidant NAC or selenium delays the onset of UV-induced melanoma tumors (Cassidy et al., 2013; Cotter et al., 2007). Honikiol, a potent scavenger of superoxide and peroxyl radicals, inhibits melanoma cell growth in vitro (Dikalov et al., 2008; Mannal et al., 2011). Selective inhibitors of nitric oxide reduce melanoma cell growth and metastasis (Yang et al., 2013). Selenium, which increases glutathione peroxidase activity and the levels of glutathione, also decreases the size of human melanoma xenografts in vivo, and inhibits the growth of human melanoma cells in vitro (Cassidy et al., 2013). Treatment of human melanoma cell lines with cAMP inducers, such as α-MSH, inhibits their proliferation, due in part to inhibition of oxidative stress (Lyons et al., 2013). The second strategy is to treat melanoma tumors with agents that trigger apoptosis by compromising ROS scavenging. Such agents include butathionine sulfoximine, which depletes GSH, and disulfiram, which inhibits copper, zinc superoxide dismutase, both of which inhibit melanoma cell proliferation in vitro (Cen et al., 2004; Fruehauf et al., 1998). Additionally, quercetin, motexafin gadolinium, melphalan and cisplatin, which inhibit thioredoxin, are effective in killing cancer cells (Hashemy et al., 2006; Lu et al., 2006; Witte et al., 2005). Resveratrol, known to inhibit APE-1/Ref-1 endonuclease activity sensitizes melanoma cells to DNA alkylating agents (Yang et al., 2005). Combined therapy utilizing some of the above agents can synergistically inhibit melanoma tumor growth. Understanding the complexity of oxidative stress pathways in pigmentation production, melanocyte proliferation and malignant transformation has enormous potential to expand our armamentarium of clinically effective compounds and offer enormous promise for patients suffering from pigmentary disorders and melanoma.

Figure 2.

Figure 2

Generation of ROS by the various steps in the melanin synthetic pathway.

Acknowledgments

Supported in part by R01 ES009110 (for ZAM) and R01 ES017561 (for ZAM and SL).

Abbreviations

ROS

Reactive oxygen species

UV

Ultraviolet radiation

DHICA

Dihydroxyindole carboxylic acid

TRP-2

Tyrosinase-related protein-2

8-OxodG

8-oxodeoxyguanosine

BER

Base excision repair

ACTH

Adrenocorticotropic hormone

MC1R

Melanocortin 1 receptor

HO-1

heme Oxygenase-1

NAC

N-acetylcysteine

iNOS

inducible nitric oxide synthase

Footnotes

Conflict of interest: The authors declare no conflict of interest.

References

  1. Bohm M, Wolff I, Scholzen TE, et al. alpha-Melanocyte-stimulating hormone protects from ultraviolet radiation-induced apoptosis and DNA damage. J Biol Chem. 2005;280:5795–802. doi: 10.1074/jbc.M406334200. [DOI] [PubMed] [Google Scholar]
  2. Boissy RE, Manga P. On the etiology of contact/occupational vitiligo. Pigment Cell Res. 2004;17:208–14. doi: 10.1111/j.1600-0749.2004.00130.x. [DOI] [PubMed] [Google Scholar]
  3. Casp CB, She JX, McCormack WT. Genetic association of the catalase gene (CAT) with vitiligo susceptibility. Pigment Cell Research. 2002a;15:62–6. doi: 10.1034/j.1600-0749.2002.00057.x. [DOI] [PubMed] [Google Scholar]
  4. Casp CB, She JX, McCormack WT. Genetic association of the catalase gene (CAT) with vitiligo susceptibility. Pigment Cell Res. 2002b;15:62–6. doi: 10.1034/j.1600-0749.2002.00057.x. [DOI] [PubMed] [Google Scholar]
  5. Cassidy PB, Fain HD, Cassidy JP, Jr, et al. Selenium for the prevention of cutaneous melanoma. Nutrients. 2013;5:725–49. doi: 10.3390/nu5030725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Cen D, Brayton D, Shahandeh B, et al. Disulfiram facilitates intracellular Cu uptake and induces apoptosis in human melanoma cells. J Med Chem. 2004;47:6914–20. doi: 10.1021/jm049568z. [DOI] [PubMed] [Google Scholar]
  7. Cotter MA, Thomas J, Cassidy P, et al. N-acetylcysteine protects melanocytes against oxidative stress/damage and delays onset of ultraviolet-induced melanoma in mice. Clin Cancer Res. 2007;13:5952–8. doi: 10.1158/1078-0432.CCR-07-1187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Cui R, Widlund HR, Feige E, et al. Central role of p53 in the suntan response and pathologic hyperpigmentation. Cell. 2007;128:853–64. doi: 10.1016/j.cell.2006.12.045. [DOI] [PubMed] [Google Scholar]
  9. Di Donato P, Napolitano A, Prota G. Metal ions as potential regulatory factors in the biosynthesis of red hair pigments: a new benzothiazole intermediate in the iron or copper assisted oxidation of 5-S-cysteinyldopa. Biochim Biophys Acta. 2002;1571:157–66. doi: 10.1016/s0304-4165(02)00212-x. [DOI] [PubMed] [Google Scholar]
  10. Dikalov S, Losik T, Arbiser JL. Honokiol is a potent scavenger of superoxide and peroxyl radicals. Biochemical pharmacology. 2008;76:589–96. doi: 10.1016/j.bcp.2008.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Elassiuty YE, Klarquist J, Speiser J, et al. Heme oxygenase-1 expression protects melanocytes from stress-induced cell death: implications for vitiligo. Exp Dermatol. 2011;20:496–501. doi: 10.1111/j.1600-0625.2010.01232.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Feng Z, Hu W, Marnett LJ, et al. Malondialdehyde, a major endogenous lipid peroxidation product, sensitizes human cells to UV- and BPDE-induced killing and mutagenesis through inhibition of nucleotide excision repair. Mutat Res. 2006;601:125–36. doi: 10.1016/j.mrfmmm.2006.06.003. [DOI] [PubMed] [Google Scholar]
  13. Feng Z, Hu W, Tang MS. Trans-4-hydroxy-2-nonenal inhibits nucleotide excision repair in human cells: a possible mechanism for lipid peroxidation-induced carcinogenesis. Proc Natl Acad Sci U S A. 2004;101:8598–602. doi: 10.1073/pnas.0402794101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Finch JS, Tome ME, Kwei KA, et al. Catalase reverses tumorigenicity in a malignant cell line by an epidermal growth factor receptor pathway. Free Radic Biol Med. 2006;40:863–75. doi: 10.1016/j.freeradbiomed.2005.10.036. [DOI] [PubMed] [Google Scholar]
  15. Fortes C, Mastroeni S, Boffetta P, et al. Polymorphisms of GSTM1 and GSTT1, sun exposure and the risk of melanoma: a case-control study. Acta Derm Venereol. 2011;91:284–9. doi: 10.2340/00015555-1078. [DOI] [PubMed] [Google Scholar]
  16. Fried L, Arbiser JL. The reactive oxygen-driven tumor: relevance to melanoma. Pigment Cell Melanoma Res. 2008;21:117–22. doi: 10.1111/j.1755-148X.2008.00451.x. [DOI] [PubMed] [Google Scholar]
  17. Fruehauf JP, Meyskens FL., Jr Reactive oxygen species: a breath of life or death? Clin Cancer Res. 2007;13:789–94. doi: 10.1158/1078-0432.CCR-06-2082. [DOI] [PubMed] [Google Scholar]
  18. Fruehauf JP, Zonis S, al-Bassam M, et al. Melanin content and downregulation of glutathione S-transferase contribute to the action of L-buthionine-S-sulfoximine on human melanoma. Chemico-biological interactions. 1998;111–112:277–305. doi: 10.1016/s0009-2797(97)00167-1. [DOI] [PubMed] [Google Scholar]
  19. Gavalas NG, Akhtar S, Gawkrodger DJ, et al. Analysis of allelic variants in the catalase gene in patients with the skin depigmenting disorder vitiligo. Biochem Biophys Res Commun. 2006;345:1586–91. doi: 10.1016/j.bbrc.2006.05.063. [DOI] [PubMed] [Google Scholar]
  20. Govindarajan B, Klafter R, Miller MS, et al. Reactive oxygen-induced carcinogenesis causes hypermethylation of p16(Ink4a) and activation of MAP kinase. Mol Med. 2002;8:1–8. [PMC free article] [PubMed] [Google Scholar]
  21. Govindarajan B, Sligh JE, Vincent BJ, et al. Overexpression of Akt converts radial growth melanoma to vertical growth melanoma. J Clin Invest. 2007;117:719–29. doi: 10.1172/JCI30102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Grammatico P, Maresca V, Roccella F, et al. Increased sensitivity to peroxidizing agents is correlated with an imbalance of antioxidants in normal melanocytes from melanoma patients. Exp Dermatol. 1998;7:205–12. doi: 10.1111/j.1600-0625.1998.tb00325.x. [DOI] [PubMed] [Google Scholar]
  23. Guan CP, Zhou MN, Xu AE, et al. The susceptibility to vitiligo is associated with NF-E2-related factor2 (Nrf2) gene polymorphisms: a study on Chinese Han population. Exp Dermatol. 2008;17:1059–62. doi: 10.1111/j.1600-0625.2008.00752.x. [DOI] [PubMed] [Google Scholar]
  24. Hashemy SI, Ungerstedt JS, Zahedi Avval F, et al. Motexafin gadolinium, a tumor-selective drug targeting thioredoxin reductase and ribonucleotide reductase. J Biol Chem. 2006;281:10691–7. doi: 10.1074/jbc.M511373200. [DOI] [PubMed] [Google Scholar]
  25. Haycock JW, Rowe SJ, Cartledge S, et al. α-Melanocyte-stimulating hormone reduces impact of proinflammatory cytokine and peroxide-generated oxidative stress on keratinocyte and melanoma cell lines. Journal of Biological Chemistry. 2000;275:15629–36. doi: 10.1074/jbc.275.21.15629. [DOI] [PubMed] [Google Scholar]
  26. Hoogduijn MJ, Cemeli E, Ross K, et al. Melanin protects melanocytes and keratinocytes against H2O2-induced DNA strand breaks through its ability to bind Ca2+ Exp Cell Res. 2004;294:60–7. doi: 10.1016/j.yexcr.2003.11.007. [DOI] [PubMed] [Google Scholar]
  27. Ibarrola-Villava M, Martin-Gonzalez M, Lazaro P, et al. Role of glutathione S-transferases in melanoma susceptibility: association with GSTP1 rs1695 polymorphism. Br J Dermatol. 2012;166:1176–83. doi: 10.1111/j.1365-2133.2012.10831.x. [DOI] [PubMed] [Google Scholar]
  28. Ichiyama T, Campbell IL, Furukawa S, et al. Autocrine alpha-melanocyte-stimulating hormone inhibits NF-kappaB activation in human glioma. J Neurosci Res. 1999;58:684–9. [PubMed] [Google Scholar]
  29. Imokawa G, Yada Y, Miyagishi M. Endothelins secreted from human keratinocytes are intrinsic mitogens for human melanocytes. Journal of Biological Chemistry. 1992;267:24675–80. [PubMed] [Google Scholar]
  30. Jain A, Lamark T, Sjottem E, et al. p62/SQSTM1 is a target gene for transcription factor NRF2 and creates a positive feedback loop by inducing antioxidant response element-driven gene transcription. J Biol Chem. 2010;285:22576–91. doi: 10.1074/jbc.M110.118976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Jenkins NC, Liu T, Cassidy P, et al. The p16(INK4A) tumor suppressor regulates cellular oxidative stress. Oncogene. 2011;30:265–74. doi: 10.1038/onc.2010.419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Jian Z, Li K, Liu L, et al. Heme oxygenase-1 protects human melanocytes from H2O2-induced oxidative stress via the Nrf2-ARE pathway. J Invest Dermatol. 2011;131:1420–7. doi: 10.1038/jid.2011.56. [DOI] [PubMed] [Google Scholar]
  33. Jimbow K, Chen H, Park JS, et al. Increased sensitivity of melanocytes to oxidative stress and abnormal expression of tyrosinase-related protein in vitiligo. British Journal of Dermatology. 2001;144:55–65. doi: 10.1046/j.1365-2133.2001.03952.x. [DOI] [PubMed] [Google Scholar]
  34. Kadekaro AL, Chen J, Yang J, et al. Alpha-melanocyte-stimulating hormone suppresses oxidative stress through a p53-mediated signaling pathway in human melanocytes. Mol Cancer Res. 2012;10:778–86. doi: 10.1158/1541-7786.MCR-11-0436. [DOI] [PubMed] [Google Scholar]
  35. Kadekaro AL, Kavanagh R, Kanto H, et al. α-Melanocortin and endothelin-1 activate anti-apoptotic pathways and reduce DNA damage in human melanocytes. Cancer Res. 2005;65:4292–9. doi: 10.1158/0008-5472.CAN-04-4535. [DOI] [PubMed] [Google Scholar]
  36. Kadekaro AL, Leachman S, Kavanagh RJ, et al. Melanocortin 1 receptor genotype: an important determinant of the damage response of melanocytes to ultraviolet radiation. Faseb J. 2010;24:3850–60. doi: 10.1096/fj.10-158485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Kaspar JW, Niture SK, Jaiswal AK. Nrf2:INrf2 (Keap1) signaling in oxidative stress. Free Radic Biol Med. 2009;47:1304–9. doi: 10.1016/j.freeradbiomed.2009.07.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Kipp C, Young AR. The soluble eumelanin precursor 5,6-dihydroxyindole-2-carboxylic acid enhances oxidative damage in human keratinocyte DNA after UVA irradiation. Photochemistry and Photobiology. 1999;70:191–8. [PubMed] [Google Scholar]
  39. Klaunig JE, Kamendulis LM. The role of oxidative stress in carcinogenesis. Annu Rev Pharmacol Toxicol. 2004;44:239–67. doi: 10.1146/annurev.pharmtox.44.101802.121851. [DOI] [PubMed] [Google Scholar]
  40. Klaunig JE, Kamendulis LM, Hocevar BA. Oxidative stress and oxidative damage in carcinogenesis. Toxicol Pathol. 2009;38:96–109. doi: 10.1177/0192623309356453. [DOI] [PubMed] [Google Scholar]
  41. Koga S, Nakano M, Terokubota S. Generation of superoxide during the enzymatic action of tyrosinase. Archives of Biochemistry and Biophysics. 1992;292:570–5. doi: 10.1016/0003-9861(92)90032-r. [DOI] [PubMed] [Google Scholar]
  42. Kokot A, Metze D, Mouchet N, et al. Alpha-melanocyte-stimulating hormone counteracts the suppressive effect of UVB on Nrf2 and Nrf-dependent gene expression in human skin. Endocrinology. 2009;150:3197–206. doi: 10.1210/en.2008-1315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Kovacs D, Flori E, Maresca V, et al. The eumelanin intermediate 5,6-dihydroxyindole-2-carboxylic acid is a messenger in the cross-talk among epidermal cells. J Invest Dermatol. 2012;132:1196–205. doi: 10.1038/jid.2011.457. [DOI] [PubMed] [Google Scholar]
  44. Kvam E, Tyrrell RM. The role of melanin in the induction of oxidative DNA base damage by ultraviolet A irradiation of DNA or melanoma cells. J Invest Dermatol. 1999;113:209–13. doi: 10.1046/j.1523-1747.1999.00653.x. [DOI] [PubMed] [Google Scholar]
  45. Landi MT, Bauer J, Pfeiffer RM, et al. MC1R germline variants confer risk for BRAF-mutant melanoma. Science. 2006;313:521–2. doi: 10.1126/science.1127515. [DOI] [PubMed] [Google Scholar]
  46. Liu F, Fu Y, Meyskens FL., Jr MiTF regulates cellular response to reactive oxygen species through transcriptional regulation of APE-1/Ref-1. J Invest Dermatol. 2009;129:422–31. doi: 10.1038/jid.2008.255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Liu J, Du J, Zhang Y, et al. Suppression of the malignant phenotype in pancreatic cancer by overexpression of phospholipid hydroperoxide glutathione peroxidase. Human gene therapy. 2006;17:105–16. doi: 10.1089/hum.2006.17.105. [DOI] [PubMed] [Google Scholar]
  48. Liu Y, Hong L, Wakamatsu K, et al. Comparison of structural and chemical properties of black and red human hair melanosomes. Photochem Photobiol. 2005;81:135–44. doi: 10.1562/2004-08-03-RA-259.1. [DOI] [PubMed] [Google Scholar]
  49. Lu J, Papp LV, Fang J, et al. Inhibition of Mammalian thioredoxin reductase by some flavonoids: implications for myricetin and quercetin anticancer activity. Cancer Res. 2006;66:4410–8. doi: 10.1158/0008-5472.CAN-05-3310. [DOI] [PubMed] [Google Scholar]
  50. Lyons J, Bastian BC, McCormick F. MC1R and cAMP signaling inhibit cdc25B activity and delay cell cycle progression in melanoma cells. Proc Natl Acad Sci U S A. 2013;110:13845–50. doi: 10.1073/pnas.1201917110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Manna SK, Aggarwal BB. Alpha-melanocyte-stimulating hormone inhibits the nuclear transcription factor NF-kappa B activation induced by various inflammatory agents. J Immunol. 1998;161:2873–80. [PubMed] [Google Scholar]
  52. Mannal PW, Schneider J, Tangada A, et al. Honokiol produces anti-neoplastic effects on melanoma cells in vitro. Journal of surgical oncology. 2011;104:260–4. doi: 10.1002/jso.21936. [DOI] [PubMed] [Google Scholar]
  53. Maresca V, Enrica F, Stefania B, et al. Correlation between melanogenic and catalase activity in in vitro human melanocytes: a synergic strategy against oxidative stress. Pigment Cell & Melanoma Research. 2008;21:200–5. doi: 10.1111/j.1755-148X.2007.00432.x. [DOI] [PubMed] [Google Scholar]
  54. Maresca V, Roccella M, Roccella F, et al. Increased sensitivity to peroxidative agents as a possible pathogenic factor of melanocyte damage in vitiligo. Journal of Investigative Dermatology. 1997;109:310–3. doi: 10.1111/1523-1747.ep12335801. [DOI] [PubMed] [Google Scholar]
  55. Marrot L, Belaidi J-P, Meunier J-R, et al. The human melanocyte as a particular target for UVA radiation and an endpoint for photoprotection assessment. Photochemistry and Photobiology. 1999;69:686–93. [PubMed] [Google Scholar]
  56. Marrot L, Jones C, Perez P, et al. The significance of Nrf2 pathway in (photo)-oxidative stress response in melanocytes and keratinocytes of the human epidermis. Pigment Cell Melanoma Res. 2008;21:79–88. doi: 10.1111/j.1755-148X.2007.00424.x. [DOI] [PubMed] [Google Scholar]
  57. Meredith P, Sarna T. The physical and chemical properties of eumelanin. Pigment Cell Res. 2006;19:572–94. doi: 10.1111/j.1600-0749.2006.00345.x. [DOI] [PubMed] [Google Scholar]
  58. Meyskens FL, Jr, McNulty SE, Buckmeier JA, et al. Aberrant redox regulation in human metastatic melanoma cells compared to normal melanocytes. Free Radic Biol Med. 2001;31:799–808. doi: 10.1016/s0891-5849(01)00650-5. [DOI] [PubMed] [Google Scholar]
  59. Michard Q, Commo S, Belaidi JP, et al. TRP-2 specifically decreases WM35 cell sensitivity to oxidative stress. Free Radic Biol Med. 2008;44:1023–31. doi: 10.1016/j.freeradbiomed.2007.11.021. [DOI] [PubMed] [Google Scholar]
  60. Mitra D, Luo X, Morgan A, et al. An ultraviolet-radiation-independent pathway to melanoma carcinogenesis in the red hair/fair skin background. Nature. 2012;491:449–53. doi: 10.1038/nature11624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Miyamura Y, Coelho SG, Schlenz K, et al. The deceptive nature of UVA tanning versus the modest protective effects of UVB tanning on human skin. Pigment Cell Melanoma Res. 2011;24:136–47. doi: 10.1111/j.1755-148X.2010.00764.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Mouret S, Forestier A, Douki T. The specificity of UVA-induced DNA damage in human melanocytes. Photochem Photobiol Sci. 2012;11:155–62. doi: 10.1039/c1pp05185g. [DOI] [PubMed] [Google Scholar]
  63. Nappi AJ, Vass E. Hydrogen peroxide generation associated with the oxidations of the eumelanin precursors 5,6-dihydroxyindole and 5,6-dihydroxyindole-2-carboxylic acid. Melanoma Res. 1996;6:341–9. doi: 10.1097/00008390-199610000-00001. [DOI] [PubMed] [Google Scholar]
  64. Natarajan VT, Singh A, Kumar AA, et al. Transcriptional upregulation of Nrf2-dependent phase II detoxification genes in the involved epidermis of vitiligo vulgaris. J Invest Dermatol. 2010;130:2781–9. doi: 10.1038/jid.2010.201. [DOI] [PubMed] [Google Scholar]
  65. Noonan FP, Zaidi MR, Wolnicka-Glubisz A, et al. Melanoma induction by ultraviolet A but not ultraviolet B radiation requires melanin pigment. Nat Commun. 2012;3:884. doi: 10.1038/ncomms1893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Park HH, Ha E, Uhm YK, et al. Association study between catalase gene polymorphisms and the susceptibility to vitiligo in Korean population. Exp Dermatol. 2006;15:377–80. doi: 10.1111/j.0906-6705.2006.00423.x. [DOI] [PubMed] [Google Scholar]
  67. Pavel S, van Nieuwpoort F, van der Meulen H, et al. Disturbed melanin synthesis and chronic oxidative stress in dysplastic naevi. Eur J Cancer. 2004;40:1423–30. doi: 10.1016/j.ejca.2003.11.035. [DOI] [PubMed] [Google Scholar]
  68. Picardo M, Grammatico P, Roccella F, et al. Imbalance in the antioxidant pool in melanoma cells and normal melanocytes from patient with melanoma. Journal of Investigative Dermatology. 1996a;107:322–6. doi: 10.1111/1523-1747.ep12363163. [DOI] [PubMed] [Google Scholar]
  69. Picardo M, Grammatico P, Roccella F, et al. Imbalance in the antioxidant pool in melanoma cells and normal melanocytes from patients with melanoma. J Invest Dermatol. 1996b;107:322–6. doi: 10.1111/1523-1747.ep12363163. [DOI] [PubMed] [Google Scholar]
  70. Picardo M, Maresca V, Eibenschutz L, et al. Correlation between antioxidants and phototypes in melanocytes cultures. A possible link of physiologic and pathologic relevance. J Invest Dermatol. 1999;113:424–5. doi: 10.1046/j.1523-1747.1999.00714.x. [DOI] [PubMed] [Google Scholar]
  71. Salem MM, Shalbaf M, Gibbons NC, et al. Enhanced DNA binding capacity on up-regulated epidermal wild-type p53 in vitiligo by H2O2-mediated oxidation: a possible repair mechanism for DNA damage. Faseb J. 2009;23:3790–807. doi: 10.1096/fj.09-132621. [DOI] [PubMed] [Google Scholar]
  72. Salopek TG, Yamada K, Ito S, et al. Dysplastic nevi contain high levels of pheomelanin: quantitative comparison of pheomelanin/eumelanin levels between normal skin, common nevi and dysplstic nevi. Pigment Cell Research. 1991;4:172–9. doi: 10.1111/j.1600-0749.1991.tb00435.x. [DOI] [PubMed] [Google Scholar]
  73. Sander CS, Chang H, Hamm F, et al. Role of oxidative stress and the antioxidant network in cutaneous carcinogenesis. Int J Dermatol. 2004;43:326–35. doi: 10.1111/j.1365-4632.2004.02222.x. [DOI] [PubMed] [Google Scholar]
  74. Schallreuter KU, Moore J, Wood JM, et al. In vivo and in vitro evidence for hydrogen peroxide (H2O2) accumulation in the epidermis of patients with vitiligo and its successful removal by a UVB-activated pseudocatalase. Journal of Investigative Dermatology Symposium Proceedings. 1999;4:91–6. doi: 10.1038/sj.jidsp.5640189. [DOI] [PubMed] [Google Scholar]
  75. Schallreuter KU, Rubsam K, Gibbons NC, et al. Methionine sulfoxide reductases A and B are deactivated by hydrogen peroxide (H2O2) in the epidermis of patients with vitiligo. J Invest Dermatol. 2008;128:808–15. doi: 10.1038/sj.jid.5701100. [DOI] [PubMed] [Google Scholar]
  76. Schallreuter KU, Salem MA, Gibbons NC, et al. Blunted epidermal L-tryptophan metabolism in vitiligo affects immune response and ROS scavenging by Fenton chemistry, part 1: Epidermal H2O2/ONOO(−)-mediated stress abrogates tryptophan hydroxylase and dopa decarboxylase activities, leading to low serotonin and melatonin levels. Faseb J. 2012;26:2457–70. doi: 10.1096/fj.11-197137. [DOI] [PubMed] [Google Scholar]
  77. Schallreuter KU, Salem MA, Holtz S, et al. Basic evidence for epidermal H2O2/ONOO(−)-mediated oxidation/nitration in segmental vitiligo is supported by repigmentation of skin and eyelashes after reduction of epidermal H2O2 with topical NB-UVB-activated pseudocatalase PC-KUS. Faseb J. 2013;27:3113–22. doi: 10.1096/fj.12-226779. [DOI] [PubMed] [Google Scholar]
  78. Schallreuter KU, Wood JM, Berger J. Low catalase levels in the epidermis of patients with vitiligo [see comments] Journal of Investigative Dermatology. 1991;97:1081–5. doi: 10.1111/1523-1747.ep12492612. [DOI] [PubMed] [Google Scholar]
  79. Simon JD, Peles D, Wakamatsu K, et al. Current challenges in understanding melanogenesis: bridging chemistry, biological control, morphology, and function. Pigment Cell Melanoma Res. 2009;22:563–79. doi: 10.1111/j.1755-148X.2009.00610.x. [DOI] [PubMed] [Google Scholar]
  80. Smit NP, van Nieuwpoort FA, Marrot L, et al. Increased melanogenesis is a risk factor for oxidative DNA damage--study on cultured melanocytes and atypical nevus cells. Photochem Photobiol. 2008;84:550–5. doi: 10.1111/j.1751-1097.2007.00242.x. [DOI] [PubMed] [Google Scholar]
  81. Song X, Mosby N, Yang J, et al. alpha-MSH activates immediate defense responses to UV-induced oxidative stress in human melanocytes. Pigment Cell Melanoma Res. 2009;22:809–18. doi: 10.1111/j.1755-148X.2009.00615.x. [DOI] [PubMed] [Google Scholar]
  82. Spencer JD, Gibbons NC, Rokos H, et al. Oxidative stress via hydrogen peroxide affects proopiomelanocortin peptides directly in the epidermis of patients with vitiligo. J Invest Dermatol. 2007;127:411–20. doi: 10.1038/sj.jid.5700538. [DOI] [PubMed] [Google Scholar]
  83. Spritz RA. Modern vitiligo genetics sheds new light on an ancient disease. J Dermatol. 2013;40:310–8. doi: 10.1111/1346-8138.12147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Swalwell H, Latimer J, Haywood RM, et al. Investigating the role of melanin in UVA/UVB- and hydrogen peroxide-induced cellular and mitochondrial ROS production and mitochondrial DNA damage in human melanoma cells. Free Radic Biol Med. 2011;52:626–34. doi: 10.1016/j.freeradbiomed.2011.11.019. [DOI] [PubMed] [Google Scholar]
  85. Tada A, Suzuki I, Im S, et al. Endothelin-1 is a paracrine growth factor that modulates melanogenesis of human melanocytes and participates in their responses to ultraviolet radiation. Cell Growth and Differentiation. 1998;9:575–84. [PubMed] [Google Scholar]
  86. Taguchi K, Motohashi H, Yamamoto M. Molecular mechanisms of the Keap1-Nrf2 pathway in stress response and cancer evolution. Genes Cells. 2011;16:123–40. doi: 10.1111/j.1365-2443.2010.01473.x. [DOI] [PubMed] [Google Scholar]
  87. Taieb A, Picardo M. The definition and assessment of vitiligo: a consensus report of the Vitiligo European Task Force. Pigment Cell Res. 2007;20:27–35. doi: 10.1111/j.1600-0749.2006.00355.x. [DOI] [PubMed] [Google Scholar]
  88. Tomita Y, Hariu A, Kato C, et al. Radical production during tyrosinase reaction, DOPA-melanin formation, and photoirradiation of DOPA-melanin. Journal of Investigative Dermatology. 1984;82:573–6. doi: 10.1111/1523-1747.ep12261311. [DOI] [PubMed] [Google Scholar]
  89. van der Kemp PA, Blais JC, Bazin M, et al. Ultraviolet-B-induced inactivation of human OGG1, the repair enzyme for removal of 8-oxoguanine in DNA. Photochem Photobiol. 2002;76:640–8. doi: 10.1562/0031-8655(2002)076<0640:ubiioh>2.0.co;2. [DOI] [PubMed] [Google Scholar]
  90. Venkataraman S, Jiang X, Weydert C, et al. Manganese superoxide dismutase overexpression inhibits the growth of androgen-independent prostate cancer cells. Oncogene. 2005;24:77–89. doi: 10.1038/sj.onc.1208145. [DOI] [PubMed] [Google Scholar]
  91. Wang HT, Choi B, Tang MS. Melanocytes are deficient in repair of oxidative DNA damage and UV-induced photoproducts. Proc Natl Acad Sci U S A. 2010;107:12180–5. doi: 10.1073/pnas.1005244107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Wenczl E, Van der Schans GP, Roza L, et al. (Pheo)Melanin photosensitizes UVA-induced DNA damage in cultured human melanocytes. Journal of Investigative Dermatology. 1998;111:678–82. doi: 10.1046/j.1523-1747.1998.00357.x. [DOI] [PubMed] [Google Scholar]
  93. Witte AB, Anestal K, Jerremalm E, et al. Inhibition of thioredoxin reductase but not of glutathione reductase by the major classes of alkylating and platinum-containing anticancer compounds. Free Radic Biol Med. 2005;39:696–703. doi: 10.1016/j.freeradbiomed.2005.04.025. [DOI] [PubMed] [Google Scholar]
  94. Yang S, Irani K, Heffron SE, et al. Alterations in the expression of the apurinic/apyrimidinic endonuclease-1/redox factor-1 (APE/Ref-1) in human melanoma and identification of the therapeutic potential of resveratrol as an APE/Ref-1 inhibitor. Mol Cancer Ther. 2005;4:1923–35. doi: 10.1158/1535-7163.MCT-05-0229. [DOI] [PubMed] [Google Scholar]
  95. Yang Z, Misner B, Ji H, et al. Targeting nitric oxide signaling with nNOS inhibitors as a novel strategy for the therapy and prevention of human melanoma. Antioxidants & redox signaling. 2013;19:433–47. doi: 10.1089/ars.2012.4563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Zhang X, Rosenstein BS, Wang Y, et al. Identification of possible reactive oxygen species involved in ultraviolet radiation-induced oxidative DNA damage. Free Radical Biology and Medicine. 1997;23:980–5. doi: 10.1016/s0891-5849(97)00126-3. [DOI] [PubMed] [Google Scholar]
  97. Zhou Z, Li CY, Li K, et al. Decreased methionine sulphoxide reductase A expression renders melanocytes more sensitive to oxidative stress: a possible cause for melanocyte loss in vitiligo. Br J Dermatol. 2009;161:504–9. doi: 10.1111/j.1365-2133.2009.09288.x. [DOI] [PubMed] [Google Scholar]

RESOURCES