Skip to main content
Dermato-endocrinology logoLink to Dermato-endocrinology
. 2013 Apr 1;5(2):239–251. doi: 10.4161/derm.22678

Putative role of HIF transcriptional activity in melanocytes and melanoma biology

Blazej Zbytek 1, Danielle L Peacock 1, Tiffany N Seagroves 1, Andrzej Slominski 1,2,*
PMCID: PMC3772912  PMID: 24194964

Abstract

Hypoxia-inducible factor-1α (HIF-1α) is a highly oxygen sensitive bHLH protein that is part of the heterodimeric HIF-1 transcription factor. Under hypoxic stress, HIF-1 activity is induced to control expression of multiple downstream target genes, including vascular endothelial growth factor (VEGF). The normal epidermis exists in a constant mild hypoxic microenvironment and constitutively expresses HIF-1α and HIF-2α. Expression of HIF-1α and/or HIF-2α has been suggested to correlate with the increased malignant potential of melanocytes, therefore, failures of melanoma therapies may be partially linked to high HIF activity. Notably, melanomas that have the V600E BRAF mutation exhibit increased HIF-1α expression. We have utilized a bioinformatics approach to identify putative hypoxia response elements (HREs) in a set of genes known to participate in the process of melanogenesis (includingTRPM1, SLC45A2, HRAS, C-KIT, PMEL and CRH). While some of the mechanistic links between these genes and the HIF pathway have been previously explored, others await further investigation. Although agents targeting HIF activity have been proposed as novel treatment modalities for melanoma, there are currently no clinical trials in progress to test their efficacy in melanoma.

Keywords: HIF-1, keratinocytes, melanocytes, melanoma, oncogenesis, skin

Introduction: Overview of the biology of the Hypoxia-Inducible Factors (HIFs)

The hypoxia-inducible factors HIF-1α and HIF-2α are oxygen-responsive basic helix-loop-helix (bHLH)-PAS domain proteins expressed in all metazoan organisms. They are key components of the HIF-1 or HIF-2 transcription factors when they heterodimerize with their common subunit, the aryl hydrocarbon receptor nuclear translocator (ARNT, also known as HIF-1β). Whereas ARNT is constitutively expressed, the HIF-α subunits are stabilized in response to decreased oxygen tension, increased reactive oxygen species (ROS) levels or growth factor activated intracellular signaling pathways.1 Hydroxylation of the HIF-α subunits promotes their physical interaction with the von Hippel-Lindau protein (VHL), which targets HIFα subunits for destruction by the 26S proteasome. Under hypoxic conditions, ROS are also generated that, in turn, cause the oxidation of iron from the ferrous to ferric state in the catalytic centers of prolyl hydroxylases (PHD), thereby leading to impaired PHD activity, loss of VHL interaction and stabilization of the HIF-α subunits. Both HIF-1 and HIF-2 bind to hypoxia response elements (HREs) in DNA and recruit the co-activators CREB binding protein (CBP) and p300, which are required for maximal induction of downstream target genes.

The HIF transcription factors regulate the transcription of hundreds of genes in a cell type-specific manner, as some tissues are more dependent on HIF-1 rather than HIF-2 activity and vice versa.2 In addition, gene knockdown, chromatin immunoprecipitation and sequencing approaches (ChIP-seq) have each identified genes that can be regulated by both HIF-1 and HIF-2, or that are preferentially regulated by either HIF-1 or HIF-2. Therefore, control of HIF activity is highly cell type and microenvironment context dependent.3-5

Some of the classic HIF target genes include glucose transporters 1 and 4 (GLUT1 and GLUT4), multiple enzymes in the glycolytic pathway, including lactate dehydrogenase A (LDHA) and phosphoglycerate kinase 1 (PGK1), VEGF, other angiogenic factors, growth factors, transferrin receptor 1, chemokine receptors (CXCR4) and lysyl oxidase (LOX).6-10 HIF-1α is overexpressed in a variety of human cancers, including melanoma.11 HIF-1α activation in human cancers also promotes several hallmarks of malignancy. For example, HIF activity stimulates immortalization through control of telomerase and phosphoglycerate mutase (PGM) expression, an enrichment of a progenitor/stem cell phenotype, de-differentiation via regulation of inhibitor of DNA binding 2 (ID2) and genetic instability (via regulation of MutS homologs 2 and 6). In addition, HIF activity controls neo-angiogenesis (VEGF), cellular metabolism, autocrine growth factor expression [including insulin-like growth factor 2 (IGF2)], transforming growth factor α (TGFα), invasion/metastasis (e.g., via E-cadherin and LOX), as well as chemotherapy resistance via regulation of the ABC transporters ABCB1 and ABCG2.12

HIF-1 expression and the role of HIF-1α in normal keratinocytes and non-melanoma skin cancers

The majority of our knowledge about the hypoxic response in the skin is derived from studies on HIF-1α, therefore, this review will primarily focus on the role of HIF-1 in the skin and in melanoma. Often quoted to be body’s largest immune/endocrine organ (comprising ~15% of total body weight and covering an average surface area of ~2 sq. m), the skin is a source of multiple peptide and cytokine mediators that act not only locally but also systemically.13,14 On the other hand, the cells in the epidermis and dermis also respond to systemic mediators delivered by the vasculature by modifying their phenotype to better preserve body homeostasis.13,14 Of note, the epidermis does not produce its own vasculature, being nourished instead through the exchange of substances via the most superficial regions of the papillary capillaries.15 The dermal vasculature forms superficial and deep dermal plexuses that are connected by the straight collaterals, whereas the superficial plexus sends papillary loops toward the skin surface. The presence and absence of local circulation are reflected in the oxygenation levels of the dermis and the epidermis. Although the human dermis is well-vascularized and therefore, well-oxygenated, the epidermis is mildly hypoxic (~5% O2), with portions of some sebaceous glands and hair follicles shown to be moderately to severely hypoxic.16 Hypoxic conditions in the epidermis lead to readily detectable levels of the HIFs, including HIF-1α and HIF-2α by immunostaining.16-18

The main cellular constituent of the epidermis, the keratinocytes, are either replicating in the basal layer (~50%) or are undergoing differentiating as they migrate toward the epidermal surface (~50%). The keratinocytes of subsequent layers gradually flatten to form a solid, cornified layer that is desquamated. The intermediate filaments, cytokeratins, are the most important structural elements of the keratinocyte, whereas the cornified layer of skin is comprised of various cross-linked proteins and lipids.

UV (UV) radiation, in particular UVB exposure, causes damage of keratinocytes,19 leading to cell cycle arrest to allow DNA repair pathways to be activated, or to apoptosis when UV damage is beyond repair. In sun-damaged skin, the apoptotic keratinocytes are known as sunburn cells. A current dogma in the field is that the main pathway linking UVB exposure to DNA damage is tumor protein 53 (p53)-dependent.20 However, it has been shown previously that keratinocytes with inactive p53 still undergo UVB-induced apoptosis.21 One example of p53-independent induction of apoptosis in keratinocytes is stimulation of cyclin-dependent kinase inhibitor 1 (p21) by HIF-1α. p21 is a pivotal regulator of the G1/S cell cycle checkpoint.22 The other example of p53-independent pro-apoptotic pathway induced by UV is mediated by HIF-1α and phorbol-12-myristate-13-acetate-induced protein 1 (NOXA). HIF-1α expression, induced through mitogen activated kinase p38 (p38 MAPK) and C-jun terminal kinase (JNK) activation, also stimulates NOXA, a protein that targets the anti-apoptotic protein myeloid cell leukemia sequence 1 (MCL-1) for degradation. In the absence of MCL-1, B-cell lymphoma 2 (BCL-2)–associated X (BAX) mediates mitochondrial outer membrane permeabilization, leading to apoptosis.23

Although there is evidence that HIF-1α positively correlates with increased microvessel density, poor survival or decreased response to therapeutic intervention in various epithelial neoplasms, such as non-small cell lung cancers, breast cancers, cervical cancers and cancers of head and neck area,24-27 the clinical significance of the expression and/or cellular localization of HIF-1α in epidermal neoplasms, such as squamous cancers and basal cell carcinomas has not been studied extensively. There are a few reports that suggest indirect involvement of HIF-1α in skin cancers since VEGF, a major direct target gene of HIF-1α, is expressed in squamous cell carcinomas28 and papillomas.29 Non-melanoma skin cancers present in renal transplant recipients have also been shown to have high numbers of VEGF-positive lymphocytes.30

Expression and role of HIF-1α in normal melanocytes and melanoma

The melanocytes are derived from neural crest cells primarily located at the dermo-epidermal junction. Their density in the skin varies among different parts of the body from 1 in 4 to 1 in 10 basal keratinocytes. Melanin, a protective pigment produced by these cells, is transferred from melanocytes through dendrites to approximately 36 neighboring keratinocytes by the process of apocopation.15 Melanin not only absorbs UV irradiation, but also serves as a scavenger of ROS and miscellaneous chemical compounds.31,32 Melanocytes also play sensory and regulatory functions within the epidermis33,34 and hair follicle35,36 with intermediates of melanogenesis serving as bioregulators37,38

Melanoma is a malignant neoplasm derived from melanocytes. Constitutive expression of HIF-1α has been shown in melanomas by immunohistochemical staining of tissue sections, by quantitative PCR, by western blotting and by immunonofluorescent staining of cultured cells.39-41 In one study, the expression of HIF-1α in patient melanomas was detectable in the majority of samples (87.6%), but expression was not correlated with any clinicopathological variables, including patient prognosis or survival.40 In another study, the expression of HIF-1α was increased at each step of progression compared with the previous phase. It was higher in vertical growth phase than in horizontal growth phase and the highest in the metastatic melanomas.41 In addition, overexpression of HIF-1α promoted soft colony formation and invasion through Matrigel. Moreover, in a separate study, it was observed that increased HIF-1α expression was present not only in the melanoma cells but also in melanoma-associated stromal cells, including pericytes.42 Finally, increased expression of HIF-1α and HIF-2α was found to be correlated with VEGF expression.43 In this particular study, HIF-2α, rather than HIF-1α, had a stronger association with VEGF expression and with poor prognosis in univariate and multivariate analyses. Of note, in these studies, Breslow’s thickness had prognostic value only in univariate analysis. Although the authors suggested that HIF-2α is a better prognosticator than HIF-1α, the overall cumulative data generated to date point to a contribution of both HIF-1α and HIF-2α, as well as VEGF, in melanoma progression and/or metastasis.

In accordance with data reported in literature and discussed in this review, we also detected by immunohistochemistry the expression of HIF-1α in malignant melanoma, with a trend of higher expression in melanoma cells vs. benign nevi (Fig. 1). HIF-1α staining was localized in both the nucleus and the cytoplasm of melanoma cells. Interestingly, an increased expression of nuclear HIF-1α was also observed in the keratinocytes in the epidermis that was colonized by melanoma cells in comparison to expression observed in the epidermis distant from melanoma cells.

graphic file with name de-5-239-g1.jpg

Figure 1. HIF-1α expression in an invasive malignant melanoma arising in melanocytic nevus. (A) H&E-stained section of the nevus side of the lesion. (B) H&E-stained section of the melanoma side of the same lesion. (C) HIF-1α localization in the section containing the nevus side. (D) HIF-1α localization in the sections containing the melanoma side. Arrows indicate melanoma cells with cytoplasmatic localization of HIF-1α signal. Asterisks indicate keratinocytes with nuclear HIF-1α positivity. Diamonds indicate the benign nevus cells with negligible to negative HIF-1α staining. H&E staining was performed with standard methodology using a Leica tissue stainer. HIF-1α immunostaining was performed using an anti-HIF1A rabbit polyclonal antibody (HPA001275, 1:50, Sigma) and detected using the using iVIEW DAB system with a VENTANA automatic stainer.

Significance of HIF-1α in current melanoma treatments

As of today there is no effective treatment of melanoma. Various modalities have been attempted but they typically do not increase survival of the patients to any significant degree.44-48 There are attempts to explain why some of these treatments fail,47,49,50 however, the recent Bayesian network meta-analysis suggest that overall survival time of patients receiving ipilimumab is expected to be better than with other therapies during management of pretreated patients with unresectable stage III or IV melanoma.51 Interestingly, there is evidence that stimulation of the hypoxic response is partially responsible for attenuation of therapeutic responses. Isolated limb-perfusion treatment and agents inducing ROS are two examples of scenarios in which HIF-1α plays a role in limiting therapeutic efficacy in either approach. Agents targeting BRAF (v-raf murine sarcoma viral oncogene homolog B1) represent another approach for which the inhibition of HIF activity may be part of the mechanism responsible for killing melanoma cells.

Isolated limb-perfusion is a treatment modality used for isolated, unresectable regional in-transit disease. Tumor necrosis factor α (TNF-α) or other therapeutic agents are administered in such a way that delivery is restricted to local tissues, without being distributed in the whole body.52 In an isolated limb-perfusion mouse model, in which melanoma cells were engrafted into either ischemic or non-ischemic limbs, tissue ischemia was shown to be a major stimulant of increased HIF-1α activity.53 In this study, ischemia induced HIF-1α activity and increased the number of new vessels in tumors and the invasiveness of the engrafted melanomas. The expression of HIF-1α, metalloproteinases and VEGF was higher in the ischemic group than in the non-ischemic controls.53 TNF-α has been previously shown to stimulate both HIF-1α and VEGF expression, which ultimately results in the promotion of blood vessel formation. Therefore, induction of HIF activity by TNFα may be involved in the potential mechanism of failure of this treatment modality.54

Dacarbasine is an Food and Drug Administration (FDA)-approved agent utilized to treat advanced melanoma.46 It primarily acts through the generation of ROS.55 ROS are important mediators of HIF-1α activity in melanoma. Increased production of ROS in melanoma cells stimulates expression of HIF-1α.39 In addition, melanoma cells respond to hypoxic conditions by the mitochondrial release of ROS that is generated through the electron transfer chain complex III, leading to stabilization and activation of HIF-1α. Furthermore, induction of ROS with subsequent increased expression of HIF-1α leads to activation of the proto-oncogene MET (hepatocyte growth factor receptor). In turn, MET expression increases the spread of melanoma cells on extracelullar matrix, demonstrating increased motility, invasion, growth of metastatic colonies and the formation of vessel-like structures.56

Vemurafenib, an inhibitor of BRAF, is also being used with limited success for treatment of advanced melanoma.45,47 The BRAF (V600E) mutation is the most common mutation found in cutaneous melanomas.57 Data collected through microarray analyses of melanomas and melanocytic cell cultures has shown that HIF1A mRNA levels were significantly increased in lesions containing the V600E mutation.58 Transfection of the BRAF(V600E) mutant into BRAF wild type (WT) melanoma cells also increased their hypoxic tolerance. Pharmacologic inhibition of BRAF by the compound BAY 43–9006 results in decreased HIF-1α expression, primarily due to reduced protein stability, without affecting the production of the HIF-1α protein.58 From this study, the authors concluded that the oncogenic activity of the V600E BRAF mutation in melanoma may be partially mediated directly through the HIF pathway.

Exogenous compounds affecting HIF-1α in melanoma

The variety of novel anticancer agents that inhibit HIF-1α and HIF-2α activity has been recently reviewed by Semenza.59 These drugs act through a variety of mechanisms. Aminoflavone, the active component of AFP-464, a drug currently studied in phase I trials for breast and kidney cancer treatment, inhibits both HIF1A mRNA and HIF-1α protein expression.59 Translation of HIF-1α is inhibited by a variety of compounds, including the mammalian target of rapamycin (mTOR) inhibitors, such as rapamycin, cardiac glycosides, such as digoxin, topoisomerases inhibitors, such as topotecan, microtubules inhibitors, such astaxotere and, even specific oligonucleotides(ENZ-2968).59 Several other compounds induce HIF-1α protein degradation. This class of drugs encompasses the heat shock protein (HSP) inhibitors (17-AAG; 17-allylamino-17-demethoxygeldanamycin), antioxidants (ascorbate), a thioredoxin inhibitor (PX-2), a histone deacetylase (LAQ824), G-rich oligonucleotides, berberine, Se-methylselenocysteine, and a commonly utilized guanylate cyclase inhibitor (YC-1).59 The antimicrobial drug acriflavine was shown by the Semenza laboratory to heterodimerize with HIF, thus, effectively blocking its transcriptional function. Anthracyclines and echinomycin block binding of the HIFs to DNA, leading to diminished expression of HIF-driven genes. Finally, the proteasome inhibitor nortezomib interferes with the domain of HIF-1α that binds to the co-activator p300, attenuating the hypoxic transcriptional response. This drug has been recently approved for treatment of hematopoietic malignancies.59

There are several reports in which various inhibitors described above have been tested in melanoma models. In addition, a variety of other inhibitors commonly used for melanoma treatment were also found to affect the HIFs indirectly. For example, MONCPT, a topoisomerase I inhibitor, decreases B16F10 melanoma metastases in C57BL/6 mice by decreasing ERK (extracellular regulated kinase) phosphorylation, HIF-1α expression and the secretion of VEGF and metalloproteinase 9.60 The imidazoacridinone C-1311 acts as both a DNA-reactive topoisomerase II and receptor tyrosine kinase inhibitor, and represses HIF-1 activity and downstream VEGF expression. In a cell-free system, C-1311 prevented HIF-1α from binding to an oligonucleotide encompassing a canonical HRE, but did not directly interfere with HIF-1α protein production. In mice bearing murine melanoma B16-F10 cells, C-1311-induced significant downregulation of VEGF, resulting in a 70% reduction of angiogenesis.61 Combined treatment with Nutlin-3, RITA (reactivation of p53 and induction of tumor cell apoptosis) and topotecan inhibited HIF-1α and the delayed growth of B16-F10 melanoma cells.62 A complex consisting of transferrin, which binds to transferrin receptor, with a short hairpin RNA (shRNA) targeting HIF1A inhibited growth of melanoma tumors in the A375 MM xenograft model.63 In MDA-MB-435 cells, which share features of breast cancer and melanoma cells, a small molecule inhibitor of HIF-1α known as DJ12 inhibited expression of HIF target genes by blocking transactivation of the HRE response element.64 Telomere homolog oligonucleotides (T-oligos) act through Wermer and ataxia teleangiectasia mutated (ATM) kinase proteins and inhibit the activity of HIF-1, leading to decreased expression of VEGF, VEGF receptor 2, angiopoieitin-1, angiopoieitin-2. These oligos also decreased tumor vascularity in melanoma xenografts in immunocompromised mice.65 Vernolide-A inhibited radiation-induced tumor angiogenesis in B16F-10 melanoma cells in C57BL/6 mice through regulation of HIF-1α, metalloproteinases 2 and 9 and VEGF.66 It has been shown that berberine has anti-angiogenic effects in melanoma and it possibly acts through HIF-1α, VEGF and nitric oxide.67 Selenium inhibits metastatic potential of murine melanoma cells and acts through inhibition of HIF-1α, VEGF and interleukin 18.68 Vitamin K3 (menadione) attenuates self-ubiquitination of the seven in absentia homolog 2 (Siah2), increases expression of PHD3, thereby leading to decreased activity of HIF-1α and phosphorylated extracellular signal-regulated kinases. Vitamin K3 also decreased the growth of melanoma xenografts.69 Lenalidomide, used for patients with myelodysplasia with a chromosome 5q abnormality, has anti-angiogenic effects by acting on cadherin 5, β-catenin and cluster of differentiation 31 (CD31), and was found to decrease significantly lung metastasis by B16-F10 cells in mice.70

Role of HIF-1α in the control of the behavior of melanoma cells

HIF-1α regulates the expression of several genes that control motility, invasion, homing to distant organs, interactions of tumor cells with endothelial cells and neo-angiogenesis in variety of tumor types. Some of the key genes relevant to melanoma will be discussed here. First, HIF-1α is a positive regulator of connective tissue growth factor (CTGF) activity, which modulates function of the transforming growth factor β (TGFß) family, including TGFß-1, -2, -3 and bone morphogenic protein 4 and 7 (BMP4 and 7).71 These proteins are known to be important in melanoma cell invasion, spreading and motility. Another hypoxia-regulated target that contributes to melanoma is BCL-2, which acts not only as an inhibitor of apoptosis, but also stimulates formation of vessel-like structures. Specifically, under hypoxic conditions BCL-2 increases HIF-1 activity and the expression of VEGF in melanoma cells.72 The effects of BCL-2 on VEGF expression were shown to be dependent on its Bcl-2 homology (BH)4, but not its BH1 or BH2 domains,73 suggesting that the BCL-2 domains engaged in angiogenesis are distinct from those that promote apoptosis.

Hypoxia is also pro-angiogenic in melanoma since it negatively regulates expression of a class of axonal sprouting inhibitors known as class 3 semaphorins (SEMA3AF). This family of proteins functions as tumor suppressors in a variety of tumor types, including melanomas, and functionally interacts with the neuropilin-2 (NRP2) receptors. Hypoxia inhibits NRP2 expression by melanoma cells. Silencing of HIF1A abrogates hypoxia-induced NRP2 repression,74 whereas overexpression of HIF-1α directly inhibits NRP2 promoter activity. Loss of NRP2 expression in tumor cells inhibited SEMA3F-dependent activities, such as depolymerization of F-actin, and inhibition of tumor cell migration. On the other hand, loss of NRP2 expression in tumor cells increased VEGF protein levels in conditioned media. This increase in VEGF protein levels promoted paracrine signaling in endothelial cells, including VEGF receptor-2 phosphorylation, and the activation of downstream signaling proteins such as mitogen activated protein kinase p44/42 and p38. In addition, the elevated VEGF levels induced migration and sprouting of endothelial cells, two key steps of tumor angiogenesis in vivo. Thus, HIF-dependent inhibition of NRP2 increases melanoma invasiveness and angiogenesis.74

The factor-inhibiting HIF (FIH) protein is another key modulator of HIF activity. Its action is via interactions with the C-terminal activation domain (C-TAD), of the hypoxia-inducible factors. Silencing of FIH severely reduced cell proliferation and in vivo tumor growth in A375 melanoma cells. Also, silencing of FIH significantly increased both the total and phosphorylated forms of the tumor suppressor p53, which led to an increase in the cell cycle inhibitor p21. Thus, FIH activity is essential for tumor growth through the suppression of the p53-p21 axis, the major barrier that prevents cancer progression.75

Identification of putative HIF-dependent target genes implicated in melanogenesis and melanocyte functions

Melanogenesis, a process of transformation of L-phenylalanine and L-tyrosine to melanin pigment, is a highly regulated within the melanocyte (for most comprehensive reviews see refs. 31, 7678). Melanocyte development is controlled by v-kit Hardy-Zuckerman 4 feline sarcoma viral oncogene homolog (C-Kit) and endothelins [acting through endothelin receptors B (EDNRB)]. Tyrosinase (TYR) is a key enzyme responsible for transformation of L-tyrosine into subsequent melanogenic intermediates that finally form polymers known as eumelanin and pheomelanin. Tyrosinase related proteins (TYRPs) stabilize and regulate activity of tyrosinase or have their own enzymatic activity. It was also proposed that these melanogenesis related proteins have additional regulatory functions aside of their role as enzymes.38,79,80 Interestingly, active melanogenesis can affect intracellular metabolism including glycolysis and pentose phosphate pathway.81-84 These properties make melanogenesis an extremely interesting pathway to study its interaction with HIF within melanoma tumor environment.37

The function of the melanogenic pathway is prominently regulated by proopiomelanocortin (POMC)-derived melanocyte stimulating hormone (MSH) and ACTH acting through MSH receptor (MC1R).31,85 The main transcription factors involved in the process are microphtalmia associated transcription factor (MITF) and Sry-related HMG box (SOX) 10. Corticotropin releasing hormone (CRH) and related peptide urocortin acting through CRH receptors stimulating production of POMC derived peptides by skin cells.86-89 Other genes that affect pigmentation or melanocyte function are: SOX5, SOX9, SOX18, TRPM1, SLC45A2, GPCR143, BRAF, HRAS, INK4/p16, MLANA, PMEL, VDR, CYP27B1 and CYP27A1.31,34,76 These genes will be introduced and described in detail below.

Identification of High Matrix Score HREs in Melanogenesis Genes

In order to identify potential direct links between HIF activity and melanocyte physiology, we have searched the promoter regions of genes involved in regulating melanogenesis for the presence of consensus HRE sites that may bind the HIF-1 and/or HIF-2 transcription factors. The results of this bioinformatics-based approach are presented in Table 1. A total of 33 genes that encode proteins regulating melanocyte behavior were analyzed for HREs. Of these genes, six were identified as having highly significant matrix scores for the presence of an HRE (p < 0.001) (Fig. 2). These included genes important in regulation of melanin synthesis (SLC45A2,90,91 PMEL92), in melanocyte genesis during embryonal development and differentiation (TRPM1)93 and (c-KIT)94, a master regulator of systemic and local responses to stress (CRH)86,89 and an important oncogene (HRAS).95

Table 1. Summary of identification of putative HREs in genes regulating melanocyte and melanoma behavior.

Gene Accession Number Number of HREs upstream of Start Site Number of HREs after start site Location after start site Genes identified with a strong matrix score
POMC
NM_000939
1 (-248)
2 (209 and 383)
First Intron
 
MC1R
NM_002386
2 (-1372 and -1799)
1 (143)
First Exon
 
MITF
NM_198159
3 (-198, -950 and -1699)
0
-
 
SOX10
NM_006941
2 (-387 and -640)
0
-
 
SOX5 (transcript variant 1)
NM_006940
1 (-720)
0
-
 
SOX5 (transcript variant 2)
NM_152989
2 (-1218 and -1577)
0
-
 
SOX9
NM_000346
0
0
-
 
SOX18
NM_018419
1 (-1240)
0
-
 
EDNRB (transcript variant 1)
NM_000115
1 (-1422)
2 (28 and 220)
First Exon
 
EDNRB (transcript variant 2)
NM_003991
2 (-128 and -1007)
0
-
 
TRPM1
NM_002420
3 (-1074, -1103 and -1265)
0
-
Strong Matrix Scores, p = 4.73E-05
SLC45A2
NM_016180
7 (from -237 to -1326)
0
-
Strong Matrix Scores, p = 4.25E-10
GPCR143
NM_000273
2 (-806 and -1661)
2 (230 and 446)
First Intron
 
BRAF
NM_004333
0
0
-
 
HRAS
NM_005343
4 (-36, -1322, -1684 and -1811)
1 (472)
First Intron
Strong Matrix Scores, p = 4.32E-10
INK4/p16
NM_000077
1 (-1216)
1 (148)
First Intron
 
c-Kit
NM_000222
4 (-142, -404, -687 and -726)
0
-
Strong Matrix Scores, p = 1.93E-04
TYR
NM_000372
0
0
-
 
TYRP1
NM_000550
0
1 (126)
First Exon
 
TYRP2
NM_001922
0
1 (231)
First Exon
 
MLANA
NM_005511
1 (-721)
0
-
 
PMEL
NM_006928
2 (-1011 and -1086)
0
-
Strong Matrix Scores, p = 4.32E-04
VDR
NM_000376
0
0
-
 
CYP27B1
NM_000785
0
1 (248)
First Intron
 
CYP27A1
NM_000784
1 (-993)
2 (346 and 400)
First Intron
 
CRH
NM_000756
0
3 (260, 280 and 338)
First Exon
Strong Matrix Scores, p = 3.81E-05
CRHR1
NM_004382
0
0
-
 
CRHR2
NM_001883
1 (-476)
0
-
 
UCN
NM_003353
1 (-311)
1 (161)
First Exon
 
UCN2
NM_033199
4 (-524, -542, -1557 and -1875
0
-
 
UCN3 NM_053049 1 (-20) 1 (134) First Exon  

The promoter regions of genes known to regulate melanocyte activity and melanoma were scanned for the presence of putative HIF response elements (HREs) using the publicly available web server, Transcription Factor Matrix (TFM) Explorer.156-158 A 2,500 base pair sequence (2,000 base pairs upstream of the start site and 500 base pairs after the start site) of each gene was scanned using weight matrices derived from JASPER and TRANSFAC® for the HRE sequence 5′-RCGTG-3′, where R is any purine (A or G.) Using this approach, a few genes were identified as having a significant matrix scores for the presence of an HRE (p < 0.001). The remaining genes were either not considered by these parameters to be significant for the presence of a consensus HRE site, or no HRE was identified in the scanned sequence submitted for analysis. A summary of all of the analyzed genes and the locations of the HREs identified in the database search are presented.

graphic file with name de-5-239-g2.jpg

Figure 2. Overview of the position and localization of putative hypoxia response elements (HREs) in genes involved in the regulation of melanogenesis. The number ‘0’ refers to the transcriptional start site of the gene. Positive numbers indicate specific base pairs located after the start site, and negative numbers indicate specific base pairs before the start site. The red line refers to the first exon of the gene and the black and gray dashed lines refer to the first intron. White boxes represent the location of the HRE identified in the gene. The forward and reverse arrows refer to the HRE direction, i.e., the forward or reverse strand, respectively. (A) Reviews the HREs identified in genes that produced strong matrix scores during database analysis (p < 0.001). (B) Lists the HREs identified in key genes involved in the regulation of the melanocyte differentiation program.

SLC45A2 (solute carrier family 45 member 2) is also known as a membrane-associated transporter protein (MATP) or melanoma antigen AIM1.90,91 It is believed that its main role is to regulate the transport of melanogenic proteins into melanosomes, where melanogenesis in vivo takes place. It regulates skin, hair and eye pigmentation and mutations in this gene are known to be the cause of oculocutaneous albinism type 4.91 Expression of the SLC45A2 gene is regulated by MITF.96 It is also considered a melanoma associated gene.97 There are seven putative HREs located upstream of the transcription start site from -237 to -1326 nucleotides.

PMEL17 (also known as GP100/SILV) is the human homolog of mouse silver locus,92 which may function in the structure of the melanosome,98 its morphogenesis99 and in the synthesis of 5,6-dihydroxyindole-2-carboxylic acid (DHICA).100 It is transcriptionally regulated by MITF in both normal and malignant melanocytes.101 There are 2 HREs located at -1011 and -1086 positions upstream of the transcription start site.

TRPM1 (melastatin) is a member of the transient receptor potential (TRP) cation channel superfamily playing important roles in a variety of cellular processes.102 TRPM1 serves as the tumor suppressor gene in melanoma;93 its expression correlates with terminal melanocytic differentiation and loss of expression has been identified as an important diagnostic and prognostic marker for primary cutaneous melanoma.102-105 Its gene codes two transcripts, the TRPM1 channel protein is encoded by the exons93 and miR-211 is encoded in one of the introns.106,107 Because of the above observations, we have proposed a dual role for TRPM1 protein to regulate melanogenesis and Ca2+ homeostasis, whereas the expression of miR-211 may be linked to its tumor suppressor function.102 Of note, three HREs are located at the -1074, -1103 and -1265 nucleotides of the promoter region of the TRPM1 gene, suggesting that the HIF pathway may play a regulatory role in both TRPM1 mRNA and miR-211 synthesis.

CRH is the central regulator of the hypothalamic-pituitary-adrenal (HPA) axis, the main organizer of the body’s response to stress. Briefly, stress induces the hypothalamic production and release of CRH, which in the anterior pituitary activates CRH receptor 1 (CRHR1) to stimulate adrenocorticotropin (ACTH) release with POMC expression.108,109 In addition, CRH regulates behavioral, autonomic, endocrine, reproductive, cardiovascular, gastro-intestinal, immune and metabolic functions both centrally and in the periphery.108,110 CRH is also expressed in the skin,111 where it regulates local homeostasis directly (reviewed in refs. 85 and 89) or indirectly through activation of the local POMC system (reviewed in ref. 112), and its expression is stimulated by UVB.113-115 The direct effects include regulation of cell proliferation, differentiation and immune interactions (reviewed in refs. 86 and 89), defining it as a novel growth factor/pleiotropic cytokine,116 while indirect effects include chain of reactions secondary to CRH induced expression of POMC in melanocytes88,117 and other skin cells87,118 or other signaling pathways as proposed.119 Therefore, CRH appears to be one of the major players in the regulation of skin homeostasis13 and melanocytic functions.34,88,116,117,120

Three HREs sites were identified at positions 260, 280 and 338 within the first exon (Table 1; Fig. 2). Of interest, we noted that the CRH peptide is coded by exon 285 but that that no HRE site was detected in the CRHR1 regulatory regions, indicating pleiotropic regulation at the level of the ligand but not the receptor (CRH also serves as the ligand for CRHR2). Interestingly, another ligand for CRHR1, urocortin 1 (UCN), also contains two HREs, one being upstream at the -311 with second located within exon 1 at +161 (Table 1). Furthermore, two other related peptides (UCN2 and UCN3) that act on CRHR2 but not CRHR1 had three (-524, -542, -1557 and -1875) and two (-20 and +134) HREs, with CRHR2 itself having one HRE (-476) (Table 1). This is consistent with a complexicity of CRH/UCN signaling in regulation of melanogenesis and hair growth as we have discussed previously.89,120 In addition, the presence of putative HREs in the CRH gene may explain of the observation of the hypoxia-induced release of CRH121,122 and the described phenomenon of ischemia123 or anemia-induced activity of the HPA axis.124

C-kit is important for development of melanocytes and their migration from the neural crest to their final destinations.125 It is also responsible for piebaldism126 and also has a known role in mediating the development of mucosal, acral and melanomas that arise in sun-damaged skin.127 Four HREs were identified at the -142, -404, -687 and -726 positions within the promoter region of the C-kit gene. Notably, HIF-1α may be an important part of the signaling transduction pathways that lead to the activation of c-kit in several malignancies. In small cell lung cancers and neuroblastoma, the stimulation of c-kit leads to increased expression of VEGF that mediates angiogenesis.128,129 In pancreatic cancers, c-kit expression results in increased invasiveness.130 In studies of various melanoma cell lines having the most common c-kit mutations, HIF-1α was found to be essential for their proliferative potential, such that mutated melanoma cells proliferated only in hypoxic conditions or in the presence of overexpressed HIF-1α. Finally, the RAS (rat sarcoma viral oncogene homolog)/RAF (murine sarcoma viral oncogene homolog)/MEK (mitogen-activated protein kinase kinase)/ERK pathways were engaged in this process.131

Harvey ras (HRAS) belongs along with KRAS and NRAS to the RAS family. These proteins operate as GTPases and trigger MAPK-dependent pathways.132 HRAS is responsible for the development of Noonan, Costello and cardio-facio-cutaneous syndromes.132 Homozygous single nucleotide polymorphism for rs 12628C portends a potential risk for bladder, colon, gastrointestinal, oral, and thyroid carcinoma and also for melanoma.95 Of note, amplification of HRAS is characteristic of Spitz nevi.133 There are five HREs from positions -2000 to +500 of the HRAS gene (Fig. 2; Table 1). H-ras also mediates the levels of HIF-1α in response to vitamin D3 in breast cancer cells. This effect on HIF-1α expression is mediated at the transcriptional level in the absence of HRAS, but in the presence of HRAS, HIF-1α expression is mediated via proteosomal degradation.134

Additional Melanogenesis Genes with Putative HRE Sites

POMC and MC1Rs are important in the regulation of melanogenesis (reviewed in refs. 31 and 85). Specifically, binding of α-MSH (one of the derivatives of POMC) to MC1R stimulates cAMP in the melanocyte that, in turn, leads to activation of downstream transcription factors. Of note, variants of MCR1 are important in the determination of skin types, hair color, response to tanning and propensity to develop melanoma.31,135 One HRE site was found to be located at the -248 position and two HREs were found to be located at the +209 and +383 positions of POMC gene. There are three HREs identified at the -1799, -1372 and +143 positions of MC1R gene. HIF1-α has been shown to control POMC expression in the hypothalamus of mice in the setting of POMC-expressing neurons responsible for hypothalamic glucose sensing.136 In addition, in renal carcinomas, in which VHL is frequently mutated, leading to overexpression of the HIFs, overexpression of POMC has also been observed, being mediated by Nurr77.137

MITF is considered to be a master regulator of melanogenesis on transcriptional level, as reviewed in reference 31. MITF regulates the expression of TYR, TYRP1 and TYRP2 genes. Mutation of MITF leads to the development of variant IIA of Waardenburg syndrome.138 Three putative HRE sites are located at -198, -950 and -1699 positions of MITF gene (Table 1). A hypoxic environment leads to inhibition of expression of MITF in melanoma cells, which, in turn, increases their malignant potential. This is mediated through HIF-1α and the basic helix-loop-helix family, member e40 (BHLHB2) pathway.139 Although MITF is a key factor in melanoma, the effects of hypoxia on MITF downregulation are not melanocyte-cell specific and can also be observed in human mast cell lines and human clear cell sarcoma.140

The Sry-related HMG box (SOX) group of transcription factors has also been implicated in the normal physiological development of a wide array of cellular functions from maintaining stem cell progeny, to lineage restriction, and in the determination of cell fate. SOX10 plays a role in melanocyte development, SOX9 affects pigment formation and SOX5 and SOX18 play a more regulatory role.141 SOX10 is known to be mutated in variant IV of Waardenburg syndrome.142 SOX genes are integral in the transcriptional regulation of the MITF.143 In metastatic melanomas, three SOX10 gene mutations have been found, two of which generate frame shift mutations before the DNA binding domain, leading to truncated proteins and inactivated SOX10 transcriptional activity. The third mutation causes a frame shift mutation at the C-terminus and this mutant SOX10 has reduced activity on the dopachrome tautomerase promoter, causing a loss of heterozygosity of SOX10 within the tumor.144 SOX10 is expressed not just in melanoma cells but also in congenital nevi.145 It is responsible for preservation of oncogenic properties of human melanoma cells.145 Two putative HREs are located at the -387 and -640 of the promoter region of the SOX10 gene. One HRE is located at the -720 promoter region of the SOX5 gene (variant 1) and two located at -1218 and -1577 of the SOX5 gene (variant 2). One HRE is located at the -1240 of the promoter region of the SOX18 gene. No HREs were identified as associated with the SOX9 gene. The relationship between SOX and HIF-1α activities is apparent in the process of new cartilage formation under hypoxic conditions sinceHIF-1α was found to be essential for the activation of SOX9, which led to the formation of chondrocytes from primitive mesenchymal cells.146

Several genes analyzed for HREs are related to structure of melanocytes and to the enzymatic pathway of the melanogenesis. TYR, TYRP1 and TYRP2 are of the most prominence here. Of note, no possible HRE sites were identified in the promoter of TYR and only a singular putative HRE region in the first exon was identified in either TYRP1 (at +126) orTYRP2 (at +231), which might suggest an indirect hypoxic regulation of TYR, which contains several HREs (Fig. 2; Table 1). In a C. s elegans model, HIF-1 was found to increase expression of TYR-2 that in turn downregulated the effect of CEP-1, which is equivalent to mammalian p53.147 Since p53 is a major stimulant of apoptosis, the anti-apoptotic effect of HIF-1 mediated by in part by the key genes of the melanogenesis machinery has obvious connotations for melanoma. Indeed, if expression of TYRP-2 is turned off in melanoma, the cells will undergo apoptosis.147

Endothelin 3 and EDNRB are important in migration of melanocytes during embryonal development, their proper positioning in the body and survival.148 Mutation of EDNRB is present at the of Waardenburg syndrome IV that is characterized by absent neural networks in the colon.148 One putative HRE is located at the -1422 of the promoter region and two HREs were identified at +28 and +220 after start site of the EDNRB (variant 1) gene. In addition, two HREs were identified at the -128 and -1007 of the promoter region of the EDNRB (variant 2) gene. Tumor growth, vessel formation and invasiveness of melanoma cells defined by expression of VEGF, COX, prostaglandin E2 production is increased by endothelin 3 binding to its receptor. This behavior is mediated by HIF-1α.149

BRAF is a serine-threonine kinase that affects MAP pathway. It is considered to the most commonly mutated gene in melanoma. Vemurafenib targeting mutated BRAF was recently FDA approved for the treatment of advanced melanomas.150 Of note, no possible HRE sites were identified in the promoter of BRAF. Resistance to vemurafenib is being increasingly recognized and is subject of ongoing studies. Vemirafenib-resistant melanomas respond to treatment with vemurafenib with an increase in HIF-1α levels, which thus serves as one of the mechanism of their resistance.151 Indeed it is thought that HIF-1α is responsible for pro-proliferative effects of BRAF V600E mutation.58

Finally, G protein coupled-receptor 143 (GPCR143) is involved in the process of pigmentation of eyes and skin. Mutation of GPCR143 leads to ocular albinism.152 Two putative HREs were identified at the -806 and -1661 of the promoter region and two HRE sites were identified at +230 and 446 after start site of the GPCR143 gene. Constitutive activation of one of GPCRs is known to lead to increased expression of VEGF through activation of HIF-1α.153 Cyclin-dependent kinase inhibitor 2A (p16) is an important inhibitor of cell cycle. p16 is mutated in familial melanoma.154 One HRE was identified at the -1216 of the promoter region and one at + 148 after start site of the p16 gene. p16 forms a physical complex with HIF-1α and modifies expression of VEGF in breast cancer cells.155

Summary and Perspective

The hypoxic transcriptional response mediated by HIF-1 and HIF-2 is activated in response to decreased oxygen levels, increased ROS or growth factor stimulation. One of the classical hypoxic response target genes, VEGF, is a key mediator of tumor neo-angiogenesis. Unlike many other organs in the body, the human epidermis exists in a constant mild hypoxic state and constitutively expresses the HIFs, including HIF-1α and HIF-2α. In keratinocytes, HIF-1α blocks the cell cycle and stimulates apoptosis by stimulating p53. In addition, HIF-1α and HIF-2α have been shown to be overexpressed in melanoma, with expression positively correlating directly with the expression of VEGF in melanoma cells, with melanoma progression and with metastatic potential. Hypoxia also stimulates expression of the CTGF, BCL-2 and SEMA3AF protein. Moreover, increased HIF-1α expression and activity seems to be one of the common explanation described as why various melanoma treatments failed (including in the isolated-limb perfusion model, and in response to treatment with dacarbazine and vemurafenib). Since drugs targeting the HIF pathway are already being studied in humans for treatment of other malignancies, their use for melanoma treatment is highly probable in the future. More importantly, our novel analysis of several genes involved in melanogenesis and melanocyte behavior has revealed that the majority of the key players in these processes contain putative functional HREs and, therefore, may be direct transcriptional targets of the HIF transcription factors. These putative direct target genes include in particular TRPM1, SLC45A2, HRAS, C-KIT, PMEL and CRH but also POMC, MC1R, MITF, SOX5, 10 and 18, EDNRB, TYRP1 and 2, MLANA, CYP27B1 and A1, CRHR2, UCN, UCN2 and 3. Melanocyte development and melanin production is a multi-step process that engages multiple players. Most connections between HIF-1α and those genes have been previously studied in non-skin and non-melanoma models. However, in the skin, HIF-1α has a role in the proliferation of melanomas with C-KIT mutations and HIF-1α participates in the downregulation of MITF expression in hypoxic environment. HIF-1α also mediates increased tumorogenic potential of melanoma cells upon binding of endothelin 3 to EDNRB.

We have shown in agreement with prior reports that melanoma cells express HIF-1α in both the nucleus and the cytoplasm. Surrounding keratinocytes adjacent to invading melanoma cells also have increased nuclear expression of HIF-1α. Although the significance of the cytoplasmic localization of HIF-1α is not clear, cytoplasmic localized HIF-1α might play other roles in skin biology than its well-established role as a transcription factor in the nucleus. These mechanisms remain to be elucidated. In summary, there is great potential for additional basic and clinical research to reveal the potential cooperative relationship of HIF activity and key melanogenesis genes in controlling melanocyte physiology, melanoma behavior and therapeutic resistance to warrant that the HIF pathway is an attractive pharmacological target for the treatment of melanoma.

Acknowledgments

We thank Michael Lee, MD for the excellent contribution to this work he has made during his one month elective rotation in the dermatopathology program at the University of Tennessee Health Science Center. The writing of this review was in part supported by NIH grants 1R01AR056666–01A2 to AS and by 1R01CA138488 to TS, and dermatopathology fellowship program at the UTHSC. We also thank Dr. Meiyun Fan at UTHSC for her assistance in training Danielle Peacock to perform database searches to identify HREs in various genes of interest.

Footnotes

References

  • 1.Semenza GL. Hypoxia-inducible factor 1 (HIF-1) pathway. Sci STKE. 2007;2007:cm8. doi: 10.1126/stke.4072007cm8. [DOI] [PubMed] [Google Scholar]
  • 2.Sowter HM, Raval RR, Moore JW, Ratcliffe PJ, Harris AL. Predominant role of hypoxia-inducible transcription factor (Hif)-1alpha versus Hif-2alpha in regulation of the transcriptional response to hypoxia. Cancer Res. 2003;63:6130–4. [PubMed] [Google Scholar]
  • 3.Schödel J, Oikonomopoulos S, Ragoussis J, Pugh CW, Ratcliffe PJ, Mole DR. High-resolution genome-wide mapping of HIF-binding sites by ChIP-seq. Blood. 2011;117:e207–17. doi: 10.1182/blood-2010-10-314427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Hu J, Stiehl DP, Setzer C, Wichmann D, Shinde DA, Rehrauer H, et al. Interaction of HIF and USF signaling pathways in human genes flanked by hypoxia-response elements and E-box palindromes. Mol Cancer Res. 2011;9:1520–36. doi: 10.1158/1541-7786.MCR-11-0090. [DOI] [PubMed] [Google Scholar]
  • 5.Stiehl DP, Bordoli MR, Abreu-Rodríguez I, Wollenick K, Schraml P, Gradin K, et al. Non-canonical HIF-2α function drives autonomous breast cancer cell growth via an AREG-EGFR/ErbB4 autocrine loop. Oncogene. 2012;31:2283–97. doi: 10.1038/onc.2011.417. [DOI] [PubMed] [Google Scholar]
  • 6.Staller P, Sulitkova J, Lisztwan J, Moch H, Oakeley EJ, Krek W. Chemokine receptor CXCR4 downregulated by von Hippel-Lindau tumour suppressor pVHL. Nature. 2003;425:307–11. doi: 10.1038/nature01874. [DOI] [PubMed] [Google Scholar]
  • 7.Rolfs A, Kvietikova I, Gassmann M, Wenger RH. Oxygen-regulated transferrin expression is mediated by hypoxia-inducible factor-1. J Biol Chem. 1997;272:20055–62. doi: 10.1074/jbc.272.32.20055. [DOI] [PubMed] [Google Scholar]
  • 8.Haase VH. Hypoxia-inducible factors in the kidney. Am J Physiol Renal Physiol. 2006;291:F271–81. doi: 10.1152/ajprenal.00071.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Forsythe JA, Jiang BH, Iyer NV, Agani F, Leung SW, Koos RD, et al. Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1. Mol Cell Biol. 1996;16:4604–13. doi: 10.1128/mcb.16.9.4604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Erler JT, Bennewith KL, Nicolau M, Dornhöfer N, Kong C, Le QT, et al. Lysyl oxidase is essential for hypoxia-induced metastasis. Nature. 2006;440:1222–6. doi: 10.1038/nature04695. [DOI] [PubMed] [Google Scholar]
  • 11.Zhong H, De Marzo AM, Laughner E, Lim M, Hilton DA, Zagzag D, et al. Overexpression of hypoxia-inducible factor 1alpha in common human cancers and their metastases. Cancer Res. 1999;59:5830–5. [PubMed] [Google Scholar]
  • 12.Semenza GL. Evaluation of HIF-1 inhibitors as anticancer agents. Drug Discov Today. 2007;12:853–9. doi: 10.1016/j.drudis.2007.08.006. [DOI] [PubMed] [Google Scholar]
  • 13.Slominski AT, Zmijewski MA, Skobowiat C, Zbytek B, Slominski RM, Steketee JD. Sensing the environment: regulation of local and global homeostasis by the skin’s neuroendocrine system. Adv Anat Embryol Cell Biol. 2012;212:v–, vii, 1-115. doi: 10.1007/978-3-642-19683-6_1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Slominski A, Wortsman J. Neuroendocrinology of the skin. Endocr Rev. 2000;21:457–87. doi: 10.1210/er.21.5.457. [DOI] [PubMed] [Google Scholar]
  • 15.Fitzpatrick TBEA, Wolff K, Freedberg IM, Austen KF. Dermatology in General Medicine. New York: McGraw Hill, 1993. [Google Scholar]
  • 16.Evans SM, Schrlau AE, Chalian AA, Zhang P, Koch CJ. Oxygen levels in normal and previously irradiated human skin as assessed by EF5 binding. J Invest Dermatol. 2006;126:2596–606. doi: 10.1038/sj.jid.5700451. [DOI] [PubMed] [Google Scholar]
  • 17.Bedogni B, Powell MB. Skin hypoxia: a promoting environmental factor in melanomagenesis. Cell Cycle. 2006;5:1258–61. doi: 10.4161/cc.5.12.2810. [DOI] [PubMed] [Google Scholar]
  • 18.Boutin AT, Weidemann A, Fu Z, Mesropian L, Gradin K, Jamora C, et al. Epidermal sensing of oxygen is essential for systemic hypoxic response. Cell. 2008;133:223–34. doi: 10.1016/j.cell.2008.02.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Taichman LB, Setlow RB. Repair of ultraviolet light damage to the DNA of cultured human epidermal keratinocytes and fibroblasts. J Invest Dermatol. 1979;73:217–9. doi: 10.1111/1523-1747.ep12514242. [DOI] [PubMed] [Google Scholar]
  • 20.de Gruijl FR. p53 mutations as a marker of skin cancer risk: comparison of UVA and UVB effects. Exp Dermatol. 2002;11(Suppl 1):37–9. doi: 10.1034/j.1600-0625.11.s.1.9.x. [DOI] [PubMed] [Google Scholar]
  • 21.Chaturvedi V, Sitailo LA, Qin JZ, Bodner B, Denning MF, Curry J, et al. Knockdown of p53 levels in human keratinocytes accelerates Mcl-1 and Bcl-x(L) reduction thereby enhancing UV-light induced apoptosis. Oncogene. 2005;24:5299–312. doi: 10.1038/sj.onc.1208650. [DOI] [PubMed] [Google Scholar]
  • 22.Cho YS, Bae JM, Chun YS, Chung JH, Jeon YK, Kim IS, et al. HIF-1alpha controls keratinocyte proliferation by up-regulating p21(WAF1/Cip1) Biochim Biophys Acta. 2008;1783:323–33. doi: 10.1016/j.bbamcr.2007.11.017. [DOI] [PubMed] [Google Scholar]
  • 23.Nys K, Maes H, Dudek AM, Agostinis P. Uncovering the role of hypoxia inducible factor-1α in skin carcinogenesis. Biochim Biophys Acta. 2011;1816:1–12. doi: 10.1016/j.bbcan.2011.02.001. [DOI] [PubMed] [Google Scholar]
  • 24.Andersen S, Donnem T, Al-Saad S, Al-Shibli K, Stenvold H, Busund LT, et al. Correlation and coexpression of HIFs and NOTCH markers in NSCLC. Anticancer Res. 2011;31:1603–6. [PubMed] [Google Scholar]
  • 25.Raza S, Kornblum N, Kancharla VP, Baig MA, Singh AB, Kalavar M. Emerging therapies in the treatment of locally advanced squamous cell cancers of head and neck. Recent Pat Anticancer Drug Discov. 2011;6:246–57. doi: 10.2174/157489211795328477. [DOI] [PubMed] [Google Scholar]
  • 26.Noordhuis MG, Eijsink JJ, Roossink F, de Graeff P, Pras E, Schuuring E, et al. Prognostic cell biological markers in cervical cancer patients primarily treated with (chemo)radiation: a systematic review. Int J Radiat Oncol Biol Phys. 2011;79:325–34. doi: 10.1016/j.ijrobp.2010.09.043. [DOI] [PubMed] [Google Scholar]
  • 27.Kim NS, Kang YJ, Jo JO, Kim HY, Oh YR, Kim YO, et al. Elevated expression of thymosin β4, vascular endothelial growth factor (VEGF), and hypoxia inducible factor (HIF)-1α in early-stage cervical cancers. Pathol Oncol Res. 2011;17:493–502. doi: 10.1007/s12253-010-9327-x. [DOI] [PubMed] [Google Scholar]
  • 28.Strieth S, Hartschuh W, Pilz L, Fusenig NE. Angiogenic switch occurs late in squamous cell carcinomas of human skin. Br J Cancer. 2000;82:591–600. doi: 10.1054/bjoc.1999.0969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Scortegagna M, Martin RJ, Kladney RD, Neumann RG, Arbeit JM. Hypoxia-inducible factor-1alpha suppresses squamous carcinogenic progression and epithelial-mesenchymal transition. Cancer Res. 2009;69:2638–46. doi: 10.1158/0008-5472.CAN-08-3643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Mackenzie KA, Miller AP, Hock BD, Gardner J, Simcock JW, Roake JA, et al. Angiogenesis and host immune response contribute to the aggressive character of non-melanoma skin cancers in renal transplant recipients. Histopathology. 2011;58:875–85. doi: 10.1111/j.1365-2559.2011.03845.x. [DOI] [PubMed] [Google Scholar]
  • 31.Slominski A, Tobin DJ, Shibahara S, Wortsman J. Melanin pigmentation in mammalian skin and its hormonal regulation. Physiol Rev. 2004;84:1155–228. doi: 10.1152/physrev.00044.2003. [DOI] [PubMed] [Google Scholar]
  • 32.Wood JM, Jimbow K, Boissy RE, Slominski A, Plonka PM, Slawinski J, et al. What’s the use of generating melanin? Exp Dermatol. 1999;8:153–64. doi: 10.1111/j.1600-0625.1999.tb00365.x. [DOI] [PubMed] [Google Scholar]
  • 33.Slominski A, Paus R, Schadendorf D. Melanocytes as “sensory” and regulatory cells in the epidermis. J Theor Biol. 1993;164:103–20. doi: 10.1006/jtbi.1993.1142. [DOI] [PubMed] [Google Scholar]
  • 34.Slominski A. Neuroendocrine activity of the melanocyte. Exp Dermatol. 2009;18:760–3. doi: 10.1111/j.1600-0625.2009.00892.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Slominski A, Paus R. Melanogenesis is coupled to murine anagen: toward new concepts for the role of melanocytes and the regulation of melanogenesis in hair growth. J Invest Dermatol. 1993;101(Suppl):90S–7S. doi: 10.1111/1523-1747.ep12362991. [DOI] [PubMed] [Google Scholar]
  • 36.Slominski A, Wortsman J, Plonka PM, Schallreuter KU, Paus R, Tobin DJ. Hair follicle pigmentation. J Invest Dermatol. 2005;124:13–21. doi: 10.1111/j.0022-202X.2004.23528.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Slominski A, Zmijewski MA, Pawelek J. L-tyrosine and L-dihydroxyphenylalanine as hormone-like regulators of melanocyte functions. Pigment Cell Melanoma Res. 2012;25:14–27. doi: 10.1111/j.1755-148X.2011.00898.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Slominski A, Paus R. Are L-tyrosine and L-dopa hormone-like bioregulators? J Theor Biol. 1990;143:123–38. doi: 10.1016/S0022-5193(05)80292-9. [DOI] [PubMed] [Google Scholar]
  • 39.Kuphal S, Winklmeier A, Warnecke C, Bosserhoff AK. Constitutive HIF-1 activity in malignant melanoma. Eur J Cancer. 2010;46:1159–69. doi: 10.1016/j.ejca.2010.01.031. [DOI] [PubMed] [Google Scholar]
  • 40.Valencak J, Kittler H, Schmid K, Schreiber M, Raderer M, Gonzalez-Inchaurraga M, et al. Prognostic relevance of hypoxia inducible factor-1alpha expression in patients with melanoma. Clin Exp Dermatol. 2009;34:e962–4. doi: 10.1111/j.1365-2230.2009.03706.x. [DOI] [PubMed] [Google Scholar]
  • 41.Mills CN, Joshi SS, Niles RM. Expression and function of hypoxia inducible factor-1 alpha in human melanoma under non-hypoxic conditions. Mol Cancer. 2009;8:104. doi: 10.1186/1476-4598-8-104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Treviño-Villarreal JH, Cotanche DA, Sepúlveda R, Bortoni ME, Manneberg O, Udagawa T, et al. Host-derived pericytes and Sca-1+ cells predominate in the MART-1- stroma fraction of experimentally induced melanoma. J Histochem Cytochem. 2011;59:1060–75. doi: 10.1369/0022155411428078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Giatromanolaki A, Sivridis E, Kouskoukis C, Gatter KC, Harris AL, Koukourakis MI. Hypoxia-inducible factors 1alpha and 2alpha are related to vascular endothelial growth factor expression and a poorer prognosis in nodular malignant melanomas of the skin. Melanoma Res. 2003;13:493–501. doi: 10.1097/00008390-200310000-00008. [DOI] [PubMed] [Google Scholar]
  • 44.Zbytek B, Carlson JA, Granese J, Ross J, Mihm MC, Jr., Slominski A. Current concepts of metastasis in melanoma. Expert Rev Dermatol. 2008;3:569–85. doi: 10.1586/17469872.3.5.569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Livingstone E, Zimmer L, Vaubel J, Schadendorf D. Current advances and perspectives in the treatment of advanced melanoma. J Dtsch Dermatol Ges. 2012;10:319–25. doi: 10.1111/j.1610-0387.2012.07895.x. [DOI] [PubMed] [Google Scholar]
  • 46.Crosby T, Fish R, Coles B, Mason MD. Systemic treatments for metastatic cutaneous melanoma. Cochrane Database Syst Rev. 2000:CD001215. doi: 10.1002/14651858.CD001215. [DOI] [PubMed] [Google Scholar]
  • 47.Hauschild A, Grob JJ, Demidov LV, Jouary T, Gutzmer R, Millward M, et al. Dabrafenib in BRAF-mutated metastatic melanoma: a multicentre, open-label, phase 3 randomised controlled trial. Lancet. 2012;380:358–65. doi: 10.1016/S0140-6736(12)60868-X. [DOI] [PubMed] [Google Scholar]
  • 48.Nikolaou VA, Stratigos AJ, Flaherty KT, Tsao H. Melanoma: new insights and new therapies. J Invest Dermatol. 2012;132:854–63. doi: 10.1038/jid.2011.421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Slominski A, Zbytek B, Slominski R. Inhibitors of melanogenesis increase toxicity of cyclophosphamide and lymphocytes against melanoma cells. Int J Cancer. 2009;124:1470–7. doi: 10.1002/ijc.24005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Slominski A, Paus R, Mihm MC. Inhibition of melanogenesis as an adjuvant strategy in the treatment of melanotic melanomas: selective review and hypothesis. Anticancer Res. 1998;18(5B):3709–15. [PubMed] [Google Scholar]
  • 51.Dequen P, Lorigan P, Jansen JP, van Baardewijk M, Ouwens MJ, Kotapati S. Systematic Review and Network Meta-Analysis of Overall Survival Comparing 3 mg/kg Ipilimumab with Alternative Therapies in the Management of Pretreated Patients with Unresectable Stage III or IV Melanoma. Oncologist. 2012 doi: 10.1634/theoncologist.2011-0427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Krementz ET, Carter RD, Sutherland CM, Muchmore JH, Ryan RF, Creech O., Jr. Regional chemotherapy for melanoma. A 35-year experience. Ann Surg. 1994;220:520–34, discussion 534-5. doi: 10.1097/00000658-199410000-00010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Sun B, Zhang D, Zhang S, Zhang W, Guo H, Zhao X. Hypoxia influences vasculogenic mimicry channel formation and tumor invasion-related protein expression in melanoma. Cancer Lett. 2007;249:188–97. doi: 10.1016/j.canlet.2006.08.016. [DOI] [PubMed] [Google Scholar]
  • 54.Jing Y, Ma N, Fan T, Wang C, Bu X, Jiang G, et al. Tumor necrosis factor-alpha promotes tumor growth by inducing vascular endothelial growth factor. Cancer Invest. 2011;29:485–93. doi: 10.3109/07357907.2011.597812. [DOI] [PubMed] [Google Scholar]
  • 55.Pourahmad J, Amirmostofian M, Kobarfard F, Shahraki J. Biological reactive intermediates that mediate dacarbazine cytotoxicity. Cancer Chemother Pharmacol. 2009;65:89–96. doi: 10.1007/s00280-009-1007-8. [DOI] [PubMed] [Google Scholar]
  • 56.Comito G, Calvani M, Giannoni E, Bianchini F, Calorini L, Torre E, et al. HIF-1α stabilization by mitochondrial ROS promotes Met-dependent invasive growth and vasculogenic mimicry in melanoma cells. Free Radic Biol Med. 2011;51:893–904. doi: 10.1016/j.freeradbiomed.2011.05.042. [DOI] [PubMed] [Google Scholar]
  • 57.Dong J, Phelps RG, Qiao R, Yao S, Benard O, Ronai Z, et al. BRAF oncogenic mutations correlate with progression rather than initiation of human melanoma. Cancer Res. 2003;63:3883–5. [PubMed] [Google Scholar]
  • 58.Kumar SM, Yu H, Edwards R, Chen L, Kazianis S, Brafford P, et al. Mutant V600E BRAF increases hypoxia inducible factor-1alpha expression in melanoma. Cancer Res. 2007;67:3177–84. doi: 10.1158/0008-5472.CAN-06-3312. [DOI] [PubMed] [Google Scholar]
  • 59.Semenza GL. Hypoxia-inducible factors: mediators of cancer progression and targets for cancer therapy. Trends Pharmacol Sci. 2012;33:207–14. doi: 10.1016/j.tips.2012.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Yang XC, Tu CX, Luo PH, Zhu H, Zhu DF, Wu HH, et al. Antimetastatic activity of MONCPT in preclinical melanoma mice model. Invest New Drugs. 2010;28:800–11. doi: 10.1007/s10637-009-9323-8. [DOI] [PubMed] [Google Scholar]
  • 61.Paradziej-Łukowicz J, Skwarska A, Peszyńska-Sularz G, Brillowska-Dąbrowska A, Konopa J. Anticancer imidazoacridinone C-1311 inhibits hypoxia-inducible factor-1α (HIF-1α), vascular endothelial growth factor (VEGF) and angiogenesis. Cancer Biol Ther. 2011;12:586–97. doi: 10.4161/cbt.12.7.15980. [DOI] [PubMed] [Google Scholar]
  • 62.de Lange J, Ly LV, Lodder K, Verlaan-de Vries M, Teunisse AF, Jager MJ, et al. Synergistic growth inhibition based on small-molecule p53 activation as treatment for intraocular melanoma. Oncogene. 2012;31:1105–16. doi: 10.1038/onc.2011.309. [DOI] [PubMed] [Google Scholar]
  • 63.Liu Y, Tao J, Li Y, Yang J, Yu Y, Wang M, et al. Targeting hypoxia-inducible factor-1alpha with Tf-PEI-shRNA complex via transferrin receptor-mediated endocytosis inhibits melanoma growth. Mol Ther. 2009;17:269–77. doi: 10.1038/mt.2008.266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Jones DT, Harris AL. Identification of novel small-molecule inhibitors of hypoxia-inducible factor-1 transactivation and DNA binding. Mol Cancer Ther. 2006;5:2193–202. doi: 10.1158/1535-7163.MCT-05-0443. [DOI] [PubMed] [Google Scholar]
  • 65.Coleman C, Levine D, Kishore R, Qin G, Thorne T, Lambers E, et al. Inhibition of melanoma angiogenesis by telomere homolog oligonucleotides. J Oncol. 2010;2010:928628. doi: 10.1155/2010/928628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Pratheeshkumar P, Kuttan G. Vernolide-A inhibits radiation-induced hypoxia-mediated tumor angiogenesis by regulating HIF-1α, MMP-2, MMP-9, and VEGF. J Environ Pathol Toxicol Oncol. 2011;30:139–51. doi: 10.1615/JEnvironPatholToxicolOncol.v30.i2.50. [DOI] [PubMed] [Google Scholar]
  • 67.Hamsa TP, Kuttan G. Antiangiogenic activity of berberine is mediated through the downregulation of hypoxia-inducible factor-1, VEGF, and proinflammatory mediators. Drug Chem Toxicol. 2012;35:57–70. doi: 10.3109/01480545.2011.589437. [DOI] [PubMed] [Google Scholar]
  • 68.Song H, Kim J, Lee HK, Park HJ, Nam J, Park GB, et al. Selenium inhibits migration of murine melanoma cells via down-modulation of IL-18 expression. Int Immunopharmacol. 2011;11:2208–13. doi: 10.1016/j.intimp.2011.10.002. [DOI] [PubMed] [Google Scholar]
  • 69.Shah M, Stebbins JL, Dewing A, Qi J, Pellecchia M, Ronai ZA. Inhibition of Siah2 ubiquitin ligase by vitamin K3 (menadione) attenuates hypoxia and MAPK signaling and blocks melanoma tumorigenesis. Pigment Cell Melanoma Res. 2009;22:799–808. doi: 10.1111/j.1755-148X.2009.00628.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lu L, Payvandi F, Wu L, Zhang LH, Hariri RJ, Man HW, et al. The anti-cancer drug lenalidomide inhibits angiogenesis and metastasis via multiple inhibitory effects on endothelial cell function in normoxic and hypoxic conditions. Microvasc Res. 2009;77:78–86. doi: 10.1016/j.mvr.2008.08.003. [DOI] [PubMed] [Google Scholar]
  • 71.Braig S, Wallner S, Junglas B, Fuchshofer R, Bosserhoff AK. CTGF is overexpressed in malignant melanoma and promotes cell invasion and migration. Br J Cancer. 2011;105:231–8. doi: 10.1038/bjc.2011.226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Trisciuoglio D, Gabellini C, Desideri M, Ragazzoni Y, De Luca T, Ziparo E, et al. Involvement of BH4 domain of bcl-2 in the regulation of HIF-1-mediated VEGF expression in hypoxic tumor cells. Cell Death Differ. 2011;18:1024–35. doi: 10.1038/cdd.2010.175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Trisciuoglio D, Gabellini C, Desideri M, Ziparo E, Zupi G, Del Bufalo D. Bcl-2 regulates HIF-1alpha protein stabilization in hypoxic melanoma cells via the molecular chaperone HSP90. PLoS ONE. 2010;5:e11772. doi: 10.1371/journal.pone.0011772. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 74.Coma S, Shimizu A, Klagsbrun M. Hypoxia induces tumor and endothelial cell migration in a semaphorin 3F- and VEGF-dependent manner via transcriptional repression of their common receptor neuropilin 2. Cell Adh Migr. 2011;5:266–75. doi: 10.4161/cam.5.3.16294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Pelletier J, Dayan F, Durivault J, Ilc K, Pécou E, Pouysségur J, et al. The asparaginyl hydroxylase factor-inhibiting HIF is essential for tumor growth through suppression of the p53-p21 axis. Oncogene. 2012;31:2989–3001. doi: 10.1038/onc.2011.471. [DOI] [PubMed] [Google Scholar]
  • 76.Plonka PM, Passeron T, Brenner M, Tobin DJ, Shibahara S, Thomas A, et al. What are melanocytes really doing all day long...? Exp Dermatol. 2009;18:799–819. doi: 10.1111/j.1600-0625.2009.00912.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Hearing VJ. Biochemical control of melanogenesis and melanosomal organization. J Investig Dermatol Symp Proc. 1999;4:24–8. doi: 10.1038/sj.jidsp.5640176. [DOI] [PubMed] [Google Scholar]
  • 78.Schallreuter KU, Kothari S, Chavan B, Spencer JD. Regulation of melanogenesis--controversies and new concepts. Exp Dermatol. 2008;17:395–404. doi: 10.1111/j.1600-0625.2007.00675.x. [DOI] [PubMed] [Google Scholar]
  • 79.Slominski A, Paus R. Towards defining receptors for L-tyrosine and L-dopa. Mol Cell Endocrinol. 1994;99:C7–11. doi: 10.1016/0303-7207(94)90001-9. [DOI] [PubMed] [Google Scholar]
  • 80.Slominski A, Moellmann G, Kuklinska E. L-tyrosine, L-dopa, and tyrosinase as positive regulators of the subcellular apparatus of melanogenesis in Bomirski Ab amelanotic melanoma cells. Pigment Cell Res. 1989;2:109–16. doi: 10.1111/j.1600-0749.1989.tb00170.x. [DOI] [PubMed] [Google Scholar]
  • 81.Scisłowski PW, Słomiński A, Bomirski A. Biochemical characterization of three hamster melanoma variants--II. Glycolysis and oxygen consumption. Int J Biochem. 1984;16:327–31. doi: 10.1016/0020-711X(84)90107-1. [DOI] [PubMed] [Google Scholar]
  • 82.Scisłowski PW, Słominski A. The role of NADP-dependent dehydrogenases in hydroxylation of tyrosine in hamster melanoma. Neoplasma. 1983;30:239–43. [PubMed] [Google Scholar]
  • 83.Scisłowski PW, Słomiński A, Bomirski A, Zydowo M. Metabolic characterization of three hamster melanoma variants. Neoplasma. 1985;32:593–8. [PubMed] [Google Scholar]
  • 84.Li W, Slominski R, Slominski AT. High-resolution magic angle spinning nuclear magnetic resonance analysis of metabolic changes in melanoma cells after induction of melanogenesis. Anal Biochem. 2009;386:282–4. doi: 10.1016/j.ab.2008.12.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Slominski A, Wortsman J, Luger T, Paus R, Solomon S. Corticotropin releasing hormone and proopiomelanocortin involvement in the cutaneous response to stress. Physiol Rev. 2000;80:979–1020. doi: 10.1152/physrev.2000.80.3.979. [DOI] [PubMed] [Google Scholar]
  • 86.Slominski A, Wortsman J, Pisarchik A, Zbytek B, Linton EA, Mazurkiewicz JE, et al. Cutaneous expression of corticotropin-releasing hormone (CRH), urocortin, and CRH receptors. FASEB J. 2001;15:1678–93. doi: 10.1096/fj.00-0850rev. [DOI] [PubMed] [Google Scholar]
  • 87.Slominski A, Zbytek B, Semak I, Sweatman T, Wortsman J. CRH stimulates POMC activity and corticosterone production in dermal fibroblasts. J Neuroimmunol. 2005;162:97–102. doi: 10.1016/j.jneuroim.2005.01.014. [DOI] [PubMed] [Google Scholar]
  • 88.Slominski A, Zbytek B, Szczesniewski A, Semak I, Kaminski J, Sweatman T, et al. CRH stimulation of corticosteroids production in melanocytes is mediated by ACTH. Am J Physiol Endocrinol Metab. 2005;288:E701–6. doi: 10.1152/ajpendo.00519.2004. [DOI] [PubMed] [Google Scholar]
  • 89.Slominski A, Zbytek B, Zmijewski M, Slominski RM, Kauser S, Wortsman J, et al. Corticotropin releasing hormone and the skin. Front Biosci. 2006;11:2230–48. doi: 10.2741/1966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Nakayama K, Fukamachi S, Kimura H, Koda Y, Soemantri A, Ishida T. Distinctive distribution of AIM1 polymorphism among major human populations with different skin color. J Hum Genet. 2002;47:92–4. doi: 10.1007/s100380200007. [DOI] [PubMed] [Google Scholar]
  • 91.Newton JM, Cohen-Barak O, Hagiwara N, Gardner JM, Davisson MT, King RA, et al. Mutations in the human orthologue of the mouse underwhite gene (uw) underlie a new form of oculocutaneous albinism, OCA4. Am J Hum Genet. 2001;69:981–8. doi: 10.1086/324340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kwon BS, Halaban R, Ponnazhagan S, Kim K, Chintamaneni C, Bennett D, et al. Mouse silver mutation is caused by a single base insertion in the putative cytoplasmic domain of Pmel 17. Nucleic Acids Res. 1995;23:154–8. doi: 10.1093/nar/23.1.154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Duncan LM, Deeds J, Hunter J, Shao J, Holmgren LM, Woolf EA, et al. Down-regulation of the novel gene melastatin correlates with potential for melanoma metastasis. Cancer Res. 1998;58:1515–20. [PubMed] [Google Scholar]
  • 94.Nishikawa S, Kusakabe M, Yoshinaga K, Ogawa M, Hayashi S, Kunisada T, et al. In utero manipulation of coat color formation by a monoclonal anti-c-kit antibody: two distinct waves of c-kit-dependency during melanocyte development. EMBO J. 1991;10:2111–8. doi: 10.1002/j.1460-2075.1991.tb07744.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Tomei S, Adams S, Uccellini L, Bedognetti D, De Giorgi V, Erdenebileg N, et al. Association between HRAS rs12628 and rs112587690 polymorphisms with the risk of melanoma in the North American population. Med Oncol. 2012 doi: 10.1007/s12032-012-0255-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Du J, Fisher DE. Identification of Aim-1 as the underwhite mouse mutant and its transcriptional regulation by MITF. J Biol Chem. 2002;277:402–6. doi: 10.1074/jbc.M110229200. [DOI] [PubMed] [Google Scholar]
  • 97.Fernandez LP, Milne RL, Pita G, Avilés JA, Lázaro P, Benítez J, et al. SLC45A2: a novel malignant melanoma-associated gene. Hum Mutat. 2008;29:1161–7. doi: 10.1002/humu.20804. [DOI] [PubMed] [Google Scholar]
  • 98.Kobayashi T, Urabe K, Orlow SJ, Higashi K, Imokawa G, Kwon BS, et al. The Pmel 17/silver locus protein. Characterization and investigation of its melanogenic function. J Biol Chem. 1994;269:29198–205. [PubMed] [Google Scholar]
  • 99.Berson JF, Harper DC, Tenza D, Raposo G, Marks MS. Pmel17 initiates premelanosome morphogenesis within multivesicular bodies. Mol Biol Cell. 2001;12:3451–64. doi: 10.1091/mbc.12.11.3451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Chakraborty AK, Platt JT, Kim KK, Kwon BS, Bennett DC, Pawelek JM. Polymerization of 5,6-dihydroxyindole-2-carboxylic acid to melanin by the pmel 17/silver locus protein. Eur J Biochem. 1996;236:180–8. doi: 10.1111/j.1432-1033.1996.t01-1-00180.x. [DOI] [PubMed] [Google Scholar]
  • 101.Du J, Miller AJ, Widlund HR, Horstmann MA, Ramaswamy S, Fisher DE. MLANA/MART1 and SILV/PMEL17/GP100 are transcriptionally regulated by MITF in melanocytes and melanoma. Am J Pathol. 2003;163:333–43. doi: 10.1016/S0002-9440(10)63657-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Guo H, Carlson JA, Slominski A. Role of TRPM in melanocytes and melanoma. Exp Dermatol. 2012;21:650–4. doi: 10.1111/j.1600-0625.2012.01565.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Carlson JA, Ross JS, Slominski AJ. New techniques in dermatopathology that help to diagnose and prognosticate melanoma. Clin Dermatol. 2009;27:75–102. doi: 10.1016/j.clindermatol.2008.09.007. [DOI] [PubMed] [Google Scholar]
  • 104.Lu S, Slominski A, Yang S-E, Sheehan C, Ross J, Carlson JA. The correlation of TRPM1 (Melastatin) mRNA expression with microphthalmia-associated transcription factor (MITF) and other melanogenesis-related proteins in normal and pathological skin, hair follicles and melanocytic nevi. J Cutan Pathol. 2010;37(Suppl 1):26–40. doi: 10.1111/j.1600-0560.2010.01504.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Deeds J, Cronin F, Duncan LM. Patterns of melastatin mRNA expression in melanocytic tumors. Hum Pathol. 2000;31:1346–56. doi: 10.1016/S0046-8177(00)80003-9. [DOI] [PubMed] [Google Scholar]
  • 106.Mazar J, DeYoung K, Khaitan D, Meister E, Almodovar A, Goydos J, et al. The regulation of miRNA-211 expression and its role in melanoma cell invasiveness. PLoS ONE. 2010;5:e13779. doi: 10.1371/journal.pone.0013779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Levy C, Khaled M, Iliopoulos D, Janas MM, Schubert S, Pinner S, et al. Intronic miR-211 assumes the tumor suppressive function of its host gene in melanoma. Mol Cell. 2010;40:841–9. doi: 10.1016/j.molcel.2010.11.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Perrin MH, Vale WW. Corticotropin releasing factor receptors and their ligand family. Ann N Y Acad Sci. 1999;885:312–28. doi: 10.1111/j.1749-6632.1999.tb08687.x. [DOI] [PubMed] [Google Scholar]
  • 109.Chrousos GP. The hypothalamic-pituitary-adrenal axis and immune-mediated inflammation. N Engl J Med. 1995;332:1351–62. doi: 10.1056/NEJM199505183322008. [DOI] [PubMed] [Google Scholar]
  • 110.Hillhouse EW, Grammatopoulos DK. The molecular mechanisms underlying the regulation of the biological activity of corticotropin-releasing hormone receptors: implications for physiology and pathophysiology. Endocr Rev. 2006;27:260–86. doi: 10.1210/er.2005-0034. [DOI] [PubMed] [Google Scholar]
  • 111.Slominski A, Ermak G, Mazurkiewicz JE, Baker J, Wortsman J. Characterization of corticotropin-releasing hormone (CRH) in human skin. J Clin Endocrinol Metab. 1998;83:1020–4. doi: 10.1210/jc.83.3.1020. [DOI] [PubMed] [Google Scholar]
  • 112.Slominski A, Wortsman J, Tuckey RC, Paus R. Differential expression of HPA axis homolog in the skin. Mol Cell Endocrinol. 2007;265-266:143–9. doi: 10.1016/j.mce.2006.12.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Slominski A, Baker J, Ermak G, Chakraborty A, Pawelek J. Ultraviolet B stimulates production of corticotropin releasing factor (CRF) by human melanocytes. FEBS Lett. 1996;399:175–6. doi: 10.1016/S0014-5793(96)01315-4. [DOI] [PubMed] [Google Scholar]
  • 114.Zbytek B, Wortsman J, Slominski A. Characterization of a ultraviolet B-induced corticotropin-releasing hormone-proopiomelanocortin system in human melanocytes. Mol Endocrinol. 2006;20:2539–47. doi: 10.1210/me.2006-0116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Skobowiat C, Dowdy JC, Sayre RM, Tuckey RC, Slominski AT. Cutaneous hypothalamic-pituitary-adrenal axis homolog: regulation by ultraviolet radiation. Am J Physiol Endocrinol Metab. 2011;301:E484–93. doi: 10.1152/ajpendo.00217.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Slominski A, Zbytek B, Pisarchik A, Slominski RM, Zmijewski MA, Wortsman J. CRH functions as a growth factor/cytokine in the skin. J Cell Physiol. 2006;206:780–91. doi: 10.1002/jcp.20530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Zbytek B, Pfeffer LM, Slominski AT. CRH inhibits NF-kappa B signaling in human melanocytes. Peptides. 2006;27:3276–83. doi: 10.1016/j.peptides.2006.07.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Ito N, Ito T, Kromminga A, Bettermann A, Takigawa M, Kees F, et al. Human hair follicles display a functional equivalent of the hypothalamic-pituitary-adrenal axis and synthesize cortisol. FASEB J. 2005;19:1332–4. doi: 10.1096/fj.04-1968fje. [DOI] [PubMed] [Google Scholar]
  • 119.Slominski A. On the role of the corticotropin-releasing hormone signalling system in the aetiology of inflammatory skin disorders. Br J Dermatol. 2009;160:229–32. doi: 10.1111/j.1365-2133.2008.08958.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Kauser S, Slominski A, Wei ET, Tobin DJ. Modulation of the human hair follicle pigmentary unit by corticotropin-releasing hormone and urocortin peptides. FASEB J. 2006;20:882–95. doi: 10.1096/fj.05-5257com. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Chen XQ, Du JZ, Wang YS. Regulation of hypoxia-induced release of corticotropin-releasing factor in the rat hypothalamus by norepinephrine. Regul Pept. 2004;119:221–8. doi: 10.1016/j.regpep.2004.02.005. [DOI] [PubMed] [Google Scholar]
  • 122.Fan JM, Chen XQ, Jin H, Du JZ. Gestational hypoxia alone or combined with restraint sensitizes the hypothalamic-pituitary-adrenal axis and induces anxiety-like behavior in adult male rat offspring. Neuroscience. 2009;159:1363–73. doi: 10.1016/j.neuroscience.2009.02.009. [DOI] [PubMed] [Google Scholar]
  • 123.Fassbender K, Schmidt R, Mössner R, Daffertshofer M, Hennerici M. Pattern of activation of the hypothalamic-pituitary-adrenal axis in acute stroke. Relation to acute confusional state, extent of brain damage, and clinical outcome. Stroke. 1994;25:1105–8. doi: 10.1161/01.STR.25.6.1105. [DOI] [PubMed] [Google Scholar]
  • 124.Allen LH. Biological mechanisms that might underlie iron’s effects on fetal growth and preterm birth. J Nutr. 2001;131(2S-2):581S–9S. doi: 10.1093/jn/131.2.581S. [DOI] [PubMed] [Google Scholar]
  • 125.Ito M, Kawa Y, Ono H, Okura M, Baba T, Kubota Y, et al. Removal of stem cell factor or addition of monoclonal anti-c-KIT antibody induces apoptosis in murine melanocyte precursors. J Invest Dermatol. 1999;112:796–801. doi: 10.1046/j.1523-1747.1999.00552.x. [DOI] [PubMed] [Google Scholar]
  • 126.Spritz RA. Molecular basis of human piebaldism. J Invest Dermatol. 1994;103(Suppl):137S–40S. doi: 10.1038/jid.1994.25. [DOI] [PubMed] [Google Scholar]
  • 127.Smalley KS, Sondak VK, Weber JS. c-KIT signaling as the driving oncogenic event in sub-groups of melanomas. Histol Histopathol. 2009;24:643–50. doi: 10.14670/HH-24.643. [DOI] [PubMed] [Google Scholar]
  • 128.Nilsson MB, Zage PE, Zeng L, Xu L, Cascone T, Wu HK, et al. Multiple receptor tyrosine kinases regulate HIF-1alpha and HIF-2alpha in normoxia and hypoxia in neuroblastoma: implications for antiangiogenic mechanisms of multikinase inhibitors. Oncogene. 2010;29:2938–49. doi: 10.1038/onc.2010.60. [DOI] [PubMed] [Google Scholar]
  • 129.Litz J, Krystal GW. Imatinib inhibits c-Kit-induced hypoxia-inducible factor-1alpha activity and vascular endothelial growth factor expression in small cell lung cancer cells. Mol Cancer Ther. 2006;5:1415–22. doi: 10.1158/1535-7163.MCT-05-0503. [DOI] [PubMed] [Google Scholar]
  • 130.Zhang H, Zhang G, Gonzalez FJ, Park SM, Cai D. Hypoxia-inducible factor directs POMC gene to mediate hypothalamic glucose sensing and energy balance regulation. PLoS Biol. 2011;9:e1001112. doi: 10.1371/journal.pbio.1001112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Monsel G, Ortonne N, Bagot M, Bensussan A, Dumaz N. c-Kit mutants require hypoxia-inducible factor 1alpha to transform melanocytes. Oncogene. 2010;29:227–36. doi: 10.1038/onc.2009.320. [DOI] [PubMed] [Google Scholar]
  • 132.Tidyman WE, Rauen KA. Noonan, Costello and cardio-facio-cutaneous syndromes: dysregulation of the Ras-MAPK pathway. Expert Rev Mol Med. 2008;10:e37. doi: 10.1017/S1462399408000902. [DOI] [PubMed] [Google Scholar]
  • 133.van Engen-van Grunsven AC, van Dijk MC, Ruiter DJ, Klaasen A, Mooi WJ, Blokx WA. HRAS-mutated Spitz tumors: A subtype of Spitz tumors with distinct features. Am J Surg Pathol. 2010;34:1436–41. doi: 10.1097/PAS.0b013e3181f0a749. [DOI] [PubMed] [Google Scholar]
  • 134.Jiang Y, Zheng W, Teegarden D. 1α, 25-Dihydroxyvitamin D regulates hypoxia-inducible factor-1α in untransformed and Harvey-ras transfected breast epithelial cells. Cancer Lett. 2010;298:159–66. doi: 10.1016/j.canlet.2010.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Böhm M, Luger TA, Tobin DJ, García-Borrón JC. Melanocortin receptor ligands: new horizons for skin biology and clinical dermatology. J Invest Dermatol. 2006;126:1966–75. doi: 10.1038/sj.jid.5700421. [DOI] [PubMed] [Google Scholar]
  • 136.Zhang H, Zhang G, Gonzalez FJ, Park SM, Cai D. Hypoxia-inducible factor directs POMC gene to mediate hypothalamic glucose sensing and energy balance regulation. PLoS Biol. 2011;9:e1001112. doi: 10.1371/journal.pbio.1001112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Choi JW, Park SC, Kang GH, Liu JO, Youn HD. Nur77 activated by hypoxia-inducible factor-1alpha overproduces proopiomelanocortin in von Hippel-Lindau-mutated renal cell carcinoma. Cancer Res. 2004;64:35–9. doi: 10.1158/0008-5472.CAN-03-0145. [DOI] [PubMed] [Google Scholar]
  • 138.Lautenschlager NT, Milunsky A, DeStefano A, Farrer L, Baldwin CT. A novel mutation in the MITF gene causes Waardenburg syndrome type 2. Genet Anal. 1996;13:43–4. doi: 10.1016/1050-3862(95)00148-4. [DOI] [PubMed] [Google Scholar]
  • 139.Cheli Y, Giuliano S, Fenouille N, Allegra M, Hofman V, Hofman P, et al. Hypoxia and MITF control metastatic behaviour in mouse and human melanoma cells. Oncogene. 2012;31:2461–70. doi: 10.1038/onc.2011.425. [DOI] [PubMed] [Google Scholar]
  • 140.Feige E, Yokoyama S, Levy C, Khaled M, Igras V, Lin RJ, et al. Hypoxia-induced transcriptional repression of the melanoma-associated oncogene MITF. Proc Natl Acad Sci USA. 2011;108:E924–33. doi: 10.1073/pnas.1106351108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Harris ML, Baxter LL, Loftus SK, Pavan WJ. Sox proteins in melanocyte development and melanoma. Pigment Cell Melanoma Res. 2010;23:496–513. doi: 10.1111/j.1755-148X.2010.00711.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Pingault V, Bondurand N, Kuhlbrodt K, Goerich DE, Préhu MO, Puliti A, et al. SOX10 mutations in patients with Waardenburg-Hirschsprung disease. Nat Genet. 1998;18:171–3. doi: 10.1038/ng0298-171. [DOI] [PubMed] [Google Scholar]
  • 143.Elworthy S, Lister JA, Carney TJ, Raible DW, Kelsh RN. Transcriptional regulation of mitfa accounts for the sox10 requirement in zebrafish melanophore development. Development. 2003;130:2809–18. doi: 10.1242/dev.00461. [DOI] [PubMed] [Google Scholar]
  • 144.Cronin JC, Wunderlich J, Loftus SK, Prickett TD, Wei X, Ridd K, et al. Frequent mutations in the MITF pathway in melanoma. Pigment Cell &amp. Melanoma Res. 2009;22:435–44. doi: 10.1111/j.1755-148X.2009.00578.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Shakhova O, Zingg D, Schaefer SM, Hari L, Civenni G, Blunschi J, et al. Sox10 promotes the formation and maintenance of giant congenital naevi and melanoma. Nat Cell Biol. 2012;14:882–90. doi: 10.1038/ncb2535. [DOI] [PubMed] [Google Scholar]
  • 146.Robins JC, Akeno N, Mukherjee A, Dalal RR, Aronow BJ, Koopman P, et al. Hypoxia induces chondrocyte-specific gene expression in mesenchymal cells in association with transcriptional activation of Sox9. Bone. 2005;37:313–22. doi: 10.1016/j.bone.2005.04.040. [DOI] [PubMed] [Google Scholar]
  • 147.Sendoel A, Kohler I, Fellmann C, Lowe SW, Hengartner MO. HIF-1 antagonizes p53-mediated apoptosis through a secreted neuronal tyrosinase. Nature. 2010;465:577–83. doi: 10.1038/nature09141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Yaar M, Park HY. Melanocytes: a window into the nervous system. J Invest Dermatol. 2012;132:835–45. doi: 10.1038/jid.2011.386. [DOI] [PubMed] [Google Scholar]
  • 149.Spinella F, Rosanò L, Di Castro V, Decandia S, Nicotra MR, Natali PG, et al. Endothelin-1 and endothelin-3 promote invasive behavior via hypoxia-inducible factor-1alpha in human melanoma cells. Cancer Res. 2007;67:1725–34. doi: 10.1158/0008-5472.CAN-06-2606. [DOI] [PubMed] [Google Scholar]
  • 150.Kudchadkar R, Paraiso KH, Smalley KS. Targeting mutant BRAF in melanoma: current status and future development of combination therapy strategies. Cancer J. 2012;18:124–31. doi: 10.1097/PPO.0b013e31824b436e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Baudy AR, Dogan T, Flores-Mercado JE, Hoeflich KP, Su F, van Bruggen N, et al. FDG-PET is a good biomarker of both early response and acquired resistance in BRAFV600 mutant melanomas treated with vemurafenib and the MEK inhibitor GDC-0973. EJNMMI Res. 2012;2:22. doi: 10.1186/2191-219X-2-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Yan N, Liao X, Cai SP, Lan C, Wang Y, Zhou X, et al. A novel nonsense mutation of the GPR143 gene identified in a Chinese pedigree with ocular albinism. PLoS ONE. 2012;7:e43177. doi: 10.1371/journal.pone.0043177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Sodhi A, Montaner S, Patel V, Zohar M, Bais C, Mesri EA, et al. The Kaposi’s sarcoma-associated herpes virus G protein-coupled receptor up-regulates vascular endothelial growth factor expression and secretion through mitogen-activated protein kinase and p38 pathways acting on hypoxia-inducible factor 1alpha. Cancer Res. 2000;60:4873–80. [PubMed] [Google Scholar]
  • 154.Hussussian CJ, Struewing JP, Goldstein AM, Higgins PA, Ally DS, Sheahan MD, et al. Germline p16 mutations in familial melanoma. Nat Genet. 1994;8:15–21. doi: 10.1038/ng0994-15. [DOI] [PubMed] [Google Scholar]
  • 155.Zhang J, Lu A, Li L, Yue J, Lu Y. p16 Modulates VEGF expression via its interaction with HIF-1alpha in breast cancer cells. Cancer Invest. 2010;28:588–97. doi: 10.3109/07357900903286941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Bryne JC, Valen E, Tang MH, Marstrand T, Winther O, da Piedade I, et al. JASPAR, the open access database of transcription factor-binding profiles: new content and tools in the 2008 update. Nucleic Acids Res. 2008;36(Database issue):D102–6. doi: 10.1093/nar/gkm955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Matys V, Kel-Margoulis OV, Fricke E, Liebich I, Land S, Barre-Dirrie A, et al. TRANSFAC and its module TRANSCompel: transcriptional gene regulation in eukaryotes. Nucleic Acids Res. 2006;34(Database issue):D108–10. doi: 10.1093/nar/gkj143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Tonon L, Touzet H, Varré JS. TFM-Explorer: mining cis-regulatory regions in genomes. Nucleic Acids Res. 2010;38(Web Server issue):W286-92. doi: 10.1093/nar/gkq473. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Dermato-endocrinology are provided here courtesy of Taylor & Francis

RESOURCES