• No results found

Methodological considerations

Transient Transfection

Transfection is a method to aid delivery of genetic material (DNA or RNA) through the cell membrane by the use of an agent called transfection reagent. In general, transfection reagents can be divided into two groups; cationic lipids or cationic polycations. Cationic lipids such as Lipofectamine 2000 and RNAiMAX are effective transfection reagents for the transfer of mRNA and siRNA molecules into the cell.

However, possible toxicity must be controlled. Even though the toxicity of these agents (Lipofectamine 2000 and RNAiMAX) is reported to be limited, the transfection protocol may promote unspecific responses that can interfere with cell viability. These effects are cell line dependent and some cell lines may also show increased proliferation as a result of the transfection procedure. Previously, it has been reported that the transfection reagents (cationic lipids) may interfere with the functions of the cell membrane [165, 166], and the cells can become more sensitive to adverse effects such as immune activation, inflammation and cytotoxicity, thus inhibiting cell cycle progression [167]. The differences in the amount of oligonucleotide uptake between experiments as well as differences in uptake between cell lines can affect the results. In this study, siRNA molecules were used to reduce expression of MITF-M and p16INK4A mRNA levels. Gene silencing by the use of siRNA molecules will never be 100% efficient, keeping small amounts of targeted protein present within the cell. In contrast to modulations at the DNA level (CRISPR) where gene expression can be completely abolished. When using siRNA molecules it

69 is important to consider the protein turnover with respect to choosing an appropriate transfection time. In addition siRNA molecules have a restricted duration of efficacy before they are degraded and/or are diluted into daughter cells. Many of these challenges when using siRNA molecules can to some extent be avoided by stably transducing cell lines with vectors containing short hairpin RNA (shRNA), a method that is also reported to have a lower probability of off-target effects [168].

Nevertheless, the use of siRNA molecules and transient transfection was selected in this study, based on previously optimized transfection protocols and the ease of use.

Cell viability

The MTS One Solution Cell Proliferation Assay from Promega is easy to use, and requires only a spectrophometer. The values measured after running a MTS assay are often used to assess cell viability, since the amount of formazan color produced are proportional to the count of viable cells [169]. However, it is important to recognize that the MTS assay measures metabolically active cells, and not cell viability in form of the amount of e.g. apoptotic cells. Thus, the difference measured in absorbance might not reflect a difference in only proliferation, as senescent cells also are metabolically active. Moreover, the MTS assay can be influenced by changes in culture conditions such as pH and nutrients (glucose and fatty acid content) [170].

Therefore, it is beneficial to compare the results from several cell viability assays to strengthen the data. In this study we also investigated cell proliferation rates after various treatments by using time-laps microscopy assay (IncuCyte).

Gene expression measured by qPCR

qPCR was used to study potential differential gene expression. One of the important considerations in this method is the normalization of the results. Reference genes correct the variation between the samples and serve as loading control. Choosing reference genes that are stable among the samples is therefore important.

Housekeeping genes TBP and RPLPO were used in this study, as their gene expression was not altered by experimental conditions. The Ct values obtained from qPCR are the normalized value of the target gene. The normalizer is a mean Ct value from the two reference genes.

70 A comparison between different genes in the same conditions using Ct values is challenging. In this study the same target genes with the same reference genes (TBP and RPLPO) were used on the same samples from different cell lines. These Ct values can then be compared as they have the same reference point. To ensure that the threshold does not influence the Ct values, the threshold was set at the same point (in the gene analysis program, Bio-Rad) where the amplification curves had an exponential increase.

71

5 FUTURE PERSPECTIVES

To further investigate the role of MITF upon the CDKN2A gene regulation in melanoma, it would be mandatory to investigate the potential binding of MITF protein to the INK4A and ARF promoter in various cell backgrounds spanning the disease. This could be achieved by running a ChIP-PCR of the relevant promoter. If MITF binds to the promoter of p16INK4A and/or p14ARF, it would be interesting to mutate the MITF binding site by CRISP. Then, overexpress p16INK4A/p14ARF and compare against wild type (non-mutated promoter) for downstream effects such as senescence (by using a β-galactosidase assay). Since MITF has been reported to activate p16INK4A, and they both have been found to induce senescence (after modulation), it is challenging to separate downstream effects from the two genes. We find it interesting that depletion of MITF alone and induction of p16 both induces senescence, while depletion of MITF has been found to inhibit p16INK4A expression. Is MITF using an alternative pathway to induce senescence (p21/p53) or do our results make sense (imply that MITF could inhibit p16INK4A expression)? Several questions need to be addressed to elucidate the role of MITF and p16INK4A in the senescence program. A better understanding of the senescence program could give rise to new treatment modalities involving e.g. natural killer (NK) cells that can eliminate malignant senescent cells [171].

72

6 CONCLUSIONS

- Positive association between p16INK4A and p14ARF mRNA expression in an extended melanoma cell line panel.

- MITF-M depletion up-regulates mRNA levels of p16INK4A and p14ARF the majority of melanoma cell lines tested. The exception found in the WM983B cell line indicates that regulation of p16INK4A and p14ARF expression by MITF-M is cell line specific.

- Knockdown of p16INK4A have no specific effect on proliferation rate in melanoma cell line WM1382.

- Overexpression of p16 wild type mRNA have no significant effect on proliferation rate, compared to overexpression of p16 P81T mutated mRNA in cell lines WM9 and WM983B.

73

7 APPENDIX

Table A1. Cell culturing Producer

RPMI-1640 medium Sigma@ Life Sciences

L-alanyl-L-glutamine (Glutamax) Sigma@ Life Sciences Fetal Bovine Serum (FBS) Sigma@ Life Sciences

Culture EasY-flasks Nunc

6-well plates Nunclon Thermo-fisher

Trypsin-EDTA Sigma@ Life Sciences

CountessTM automated cell counter Invitrogen CountessTM cell counter chamber slides Invitrogen

Dimethyl-sulphoxide-hybri-max (DMSO) Sigma@ Life Sciences 1420 multilable counter victor2 Wallac

PBS Sigma@ Life Sciences

96-well plates Corning Costar

254 medium Gibco Invitrogen

Phorbol-12-myristate-13-acetate Sigma-Altdich

Cholera toxin Sigma-Altdich

Human stem cell factor (hSCF) Thermo-Fisher

Endothilin (EDN1) Milipore

VenorGEM Mycoplasma detection kit Biolabs

Cell lines Source

WM35 Coriell Institute (Wistar collection)

WM1366 Coriell Institute (Wistar collection)

WM983 Coriell Institute (Wistar collection)

WM115 Coriell Institute (Wistar collection)

WM1341 Coriell Institute (Wistar collection)

FMX-I Norwegian Radium Hospital

FMX-V Norwegian Radium Hospital

LOXIMVI Norwegian Radium Hospital

MEWO American Type Culture Collection

(ATCC)

WM9 Coriell Institute (Wistar collection)

WM1382 Coriell Institute (Wistar collection)

WM239A Coriell Institute (Wistar collection)

WM45.1 Coriell Institute (Wistar collection)

SKMEL-28 American Type Culture Collection

(ATCC)

WM266-4 Coriell Institute (Wistar collection)

A375 American Type Culture Collection

(ATCC)

74

WM852 Coriell Institute (Wistar collection)

Hermes 3C The welcome trust genomic cell bank

Hermes 4C The welcome trust genomic cell bank

Table A2. Transfection Producer

siRNA MITF Eurogenetic

siRNA p16 Medprobe

EGFP mRNA Medprobe

Lipofectamine RNAiMAX Invitrogen

Lipofectamine 2000 Invitrogen

Negative control #1 random sequence Eurogenetic

Table A3.Real time PCR Producer

NANODROP 2000 Thermo Scientific

qScriptTM DNA synthesis kit Quanta Biosciences Total RNA Miniprep kit Sigma-Aldrich

Thermo-cycler Gene-amp@ PE Applied Biosystems Nucleotide-Perfecta@ super mix Quanta Biosciences Real time PCR CFX-connectTM Bio-RAD

Perfecta TM SYBR@ Green supermix Quanta Biosciences CFX manager 3.1 softweare Bio-Rad

Table A4. Western immunoblotting

Antibodies Dilution Producer

Anti-β-actin (mouse) 1:10 000 (primary) Sigma-Aldrich Anti-α-tubulin (mouse) 1:10 000 (primary) Sigma-Aldrich Anti-mouse (rabbit) 1:10 000 (secundary) Daco

Anti-MITF (mouse) 1:10 00 (primary) Cell signaling

Anti-p16 (mouse) 1:10 00 (primary) Millipore

Phosphatase inhibitors 1:100 Sigma-Aldrich

Sulotion buffers

Lysis buffer:

NaCl 150mM

Tris-HCL pH 7,5 50mM

NP-40 0,1% (v/v)

EDTA 1 med mer

ddH2O Adjust

TBS-T:

NaCl (150mM) 5M - 30ml Tris-HCL pH 7,5

(20mM) 1M - 20ml

Tween-20 (0,1%) 20% - 5ml

ddH2O Adjust to 1

75 Liter

Table A5. PCR primers (producer IDT)

Primer target Sequence Direction

MITF-M

5'-CATTGT-TAT-GCT-GGA-AAT-GCT-AGA-3' Forward

5'-GC-TAA-AGT-GGT-AGA-AAG-GTA-CTG-C-3' Reverse

p16v3

5'-GTCTG-TGA-TTA-CAA-ACC-CCT-TCT-G-3' Forward

5’-GTCCC-CTT-GCC-TGG-AAA-GAT-AC-3’ Reverse

p14 5’-TCCCA-GTC-TGC-AGT-TAA-GGG-3’ Forward

5’-TGAAC-CAC-GAA-AAC-CCT-CAG-3’ Reverse

TBP 5’-GCCCGA-AAC-GCC-GAA-TAT-3’ Forward

5’-CGTGGC-TCT-CTT-ATC-CTC-ATG-A-3’ Reverse

RPLPO 5’-CGC-TGC-TGA-ACA-TGC-TCA-AC-3’ Forward

5’-TCG-AAC-ACC-TGC-TGG-ATG-AC-3’ Reverse

Primer efficiency tests was conducted with the following results: p16v3 95,1%, p14 97,6% and MITF-M 93,8%.

76

8 SUPPLEMENTARY FIGURES

Supplementary figure S1: images (20x magnification) captured by phase contrast microscopy (X-Cite®, Life Science) after siRNA silencing of MITF in the WM1382 cell line (120 hours). Image A represents untreated control, B: negative control (NC siRNA), and C: MITF target siRNA to the right.

Supplementary figure S2: the WM1382 cell line were transfected with p16 siRNA and monitored for 120 hours by time-lapse microscopy (incuCyteZoom). A: untreated control. B: negative control (NC siRNA) and C: p16 siRNA.

A B C

A B C

77

Supplementary figure S3: show representative growth curves of WM9 and WM1382. Cells were monitored for 120 hours by time-lapse microscopy (incuCyteZoom) in two independent experiments.

The program was set to take contrast images every 3 hours. The incorporated software analyzed the phase contrast images, and calculated confluence over time, by measuring the area covered with cells.

A much higher growth rate was observed for the WM9 cells (grey curve) compared to WM1382 cells (black curve).

0 10 20 30 40 50 60 70 80 90 100

0 24 48 72 96 120

Cell density (%)

Time (h) WM1382

WM9

78

9 REFERENCES

1. Hanahan, D. and R.A. Weinberg, The hallmarks of cancer. Cell, 2000.

100(1): p. 57-70.

2. Ferlay, J., et al., Cancer incidence and mortality worldwide: sources, methods and major patterns in GLOBOCAN 2012. Int J Cancer, 2015.

136(5): p. E359-86.

3. Kristina Kjærheim, J.I.M., Hilde Langseth, Tone Eggen og Tom K. Grimsrud, Cancer in Norway 2015- Cancer incidence, mortality, survival and

prevalence in Norway. Cancer Registry of Norway, 2016(Special issue): p.

179.

4. Negrini, S., V.G. Gorgoulis, and T.D. Halazonetis, Genomic instability--an evolving hallmark of cancer. Nat Rev Mol Cell Biol, 2010. 11(3): p. 220-8.

5. Knudson, A.G., Jr., Mutation and cancer: statistical study of retinoblastoma.

Proc Natl Acad Sci U S A, 1971. 68(4): p. 820-3.

6. Hengartner, M.O., The biochemistry of apoptosis. Nature, 2000. 407(6805):

p. 770-6.

7. Fulda, S., Evasion of apoptosis as a cellular stress response in cancer. Int J Cell Biol, 2010. 2010: p. 370835.

8. Shay, J.W. and W.E. Wright, Role of telomeres and telomerase in cancer.

Semin Cancer Biol, 2011. 21(6): p. 349-53.

9. DeBerardinis, R.J., et al., The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab, 2008. 7(1): p. 11-20.

10. Colotta, F., et al., Cancer-related inflammation, the seventh hallmark of cancer: links to genetic instability. Carcinogenesis, 2009. 30(7): p. 1073-11. 81. Folkman, J., Angiogenesis. Annu Rev Med, 2006. 57: p. 1-18.

12. Kim, R., M. Emi, and K. Tanabe, Cancer immunoediting from immune surveillance to immune escape. Immunology, 2007. 121(1): p. 1-14.

13. Lazebnik, Y., What are the hallmarks of cancer? Nat Rev Cancer, 2010.

10(4): p. 232-3.

14. Malumbres, M. and M. Barbacid, Cell cycle, CDKs and cancer: a changing paradigm. Nat Rev Cancer, 2009. 9(3): p. 153-166.

15. Nasmyth, K., Viewpoint: putting the cell cycle in order. Science, 1996.

274(5293): p. 1643-5.

16. Nigg, E.A., Cyclin-dependent protein kinases: Key regulators of the eukaryotic cell cycle. BioEssays, 1995. 17(6): p. 471-480.

17. Lakin, N.D. and S.P. Jackson, Regulation of p53 in response to DNA damage.

Oncogene, 1999. 18(53): p. 7644-55.

18. Shaw, P.H., The role of p53 in cell cycle regulation. Pathol Res Pract, 1996.

192(7): p. 669-75.

19. Ortega, S., M. Malumbres, and M. Barbacid, Cyclin D-dependent kinases, INK4 inhibitors and cancer. Biochim Biophys Acta, 2002. 1602(1): p. 73-87.

79 20. Sang, L., H.A. Coller, and J.M. Roberts, Control of the reversibility of cellular

quiescence by the transcriptional repressor HES1. Science, 2008.

321(5892): p. 1095-100.

21. Campisi, J., Cellular senescence as a tumor-suppressor mechanism. Trends Cell Biol, 2001. 11(11): p. S27-31.

22. Taylor, W.R. and G.R. Stark, Regulation of the G2/M transition by p53.

Oncogene, 2001. 20(15): p. 1803-15.

23. Kastan, M.B. and J. Bartek, Cell-cycle checkpoints and cancer. Nature, 2004.

432(7015): p. 316-23.

24. Rivlin, N., et al., Mutations in the p53 Tumor Suppressor Gene: Important Milestones at the Various Steps of Tumorigenesis. Genes Cancer, 2011.

2(4): p. 466-74.

25. Bartkova, J., et al., Cyclin D1 oncoprotein aberrantly accumulates in malignancies of diverse histogenesis. Oncogene, 1995. 10(4): p. 775-8.

26. Muthusamy, V., et al., Amplification of CDK4 and MDM2 in malignant melanoma. Genes Chromosomes Cancer, 2006. 45(5): p. 447-54.

27. Monahan, K.B., et al., Somatic p16(INK4a) loss accelerates melanomagenesis. Oncogene, 2010. 29(43): p. 5809-17.

28. Soussi, T. and K.G. Wiman, Shaping genetic alterations in human cancer:

the p53 mutation paradigm. Cancer Cell, 2007. 12(4): p. 303-12.

29. Hodis, E., et al., A Landscape of Driver Mutations in Melanoma. Cell, 2012.

150(2): p. 251-63.

30. Lo, J.A. and D.E. Fisher, The melanoma revolution: from UV carcinogenesis to a new era in therapeutics. Science, 2014. 346(6212): p. 945-9.

31. Ali, Z., N. Yousaf, and J. Larkin, Melanoma epidemiology, biology and prognosis. EJC Suppl, 2013. 11(2): p. 81-91.

32. Jovanovic, P., et al., Ocular melanoma: an overview of the current status. Int J Clin Exp Pathol, 2013. 6(7): p. 1230-44.

33. Peterson, M., M.R. Albertini, and P. Remington, Incidence, Survival, and Mortality of Malignant Cutaneous Melanoma in Wisconsin, 1995-2011.

WMJ, 2015. 114(5): p. 196-201.

34. Geiser, J., et al., Malignant melanoma-diagnosis, treatment and follow-up in Norway. Tidsskr Nor Leageforen, 2013. 20: p. 133.

35. Gandini, S., et al., Meta-analysis of risk factors for cutaneous melanoma: I.

Common and atypical naevi. Eur J Cancer, 2005. 41(1): p. 28-44.

36. Green, A.C., S.C. Wallingford, and P. McBride, Childhood exposure to ultraviolet radiation and harmful skin effects: Epidemiological evidence.

Prog Biophys Mol Biol, 2011. 107(3): p. 349-55.

37. Brash, D.E., UV Signature Mutations (). Photochem Photobiol, 2015. 91(1):

p. 15-26.

38. Krauthammer, M., et al., Exome sequencing identifies recurrent somatic RAC1 mutations in melanoma. Nat Genet, 2012. 44(9): p. 1006-14.

39. Montagut, C., et al., Elevated CRAF as a potential mechanism of acquired resistance to BRAF inhibition in melanoma. Cancer Res, 2008. 68(12): p.

4853-61.

80 40. Villanueva, J., et al., Acquired resistance to BRAF inhibitors mediated by a

RAF kinase switch in melanoma can be overcome by cotargeting MEK and IGF-1R/PI3K. Cancer Cell, 2010. 18(6): p. 683-95.

41. Wagle, N., et al., Dissecting therapeutic resistance to RAF inhibition in melanoma by tumor genomic profiling. J Clin Oncol, 2011. 29(22): p. 3085-42. 96. Paraiso, K.H., et al., PTEN loss confers BRAF inhibitor resistance to

melanoma cells through the suppression of BIM expression. Cancer Res, 2011. 71(7): p. 2750-60.

43. Girotti, M.R., et al., Inhibiting EGF receptor or SRC family kinase signaling overcomes BRAF inhibitor resistance in melanoma. Cancer Discov, 2013.

3(2): p. 158-67.

44. Johannessen, C.M., et al., COT drives resistance to RAF inhibition through MAP kinase pathway reactivation. Nature, 2010. 468(7326): p. 968-72.

45. Nazarian, R., et al., Melanomas acquire resistance to B-RAF(V600E)

inhibition by RTK or N-RAS upregulation. Nature, 2010. 468(7326): p. 973-46. 7. Tsao, H., M.B. Atkins, and A.J. Sober, Management of cutaneous melanoma.

N Engl J Med, 2004. 351(10): p. 998-1012.

47. Bhatia, S., S.S. Tykodi, and J.A. Thompson, Treatment of Metastatic Melanoma: An Overview. Oncology (Williston Park), 2009. 23(6): p. 488-48. 96. Serrone, L., et al., Dacarbazine-based chemotherapy for metastatic

melanoma: thirty-year experience overview. J Exp Clin Cancer Res, 2000.

19(1): p. 21-34.

49. Kilbridge, K.L., et al., Patient preferences for adjuvant interferon alfa-2b treatment. J Clin Oncol, 2001. 19(3): p. 812-23.

50. Gibney, G.T. and V.K. Sondak, Has targeted therapy for melanoma made chemotherapy obsolete? The Lancet Oncology, 2013. 14(8): p. 676-677.

51. Maverakis, E., et al., Metastatic melanoma - a review of current and future treatment options. Acta Derm Venereol, 2015. 95(5): p. 516-24.

52. Hodi, F.S., et al., Improved survival with ipilimumab in patients with metastatic melanoma. N Engl J Med, 2010. 363(8): p. 711-23.

53. Friedlander, P. and F.S. Hodi, Advances in targeted therapy for melanoma.

Clin Adv Hematol Oncol, 2010. 8(9): p. 619-27.

54. Khan, M.K., et al., Future of radiation therapy for malignant melanoma in an era of newer, more effective biological agents. Onco Targets Ther, 2011.

4: p. 137-48.

55. Jazirehi, A.R., A. Lim, and T. Dinh, PD-1 inhibition and treatment of

advanced melanoma-role of pembrolizumab. Am J Cancer Res, 2016. 6(10):

p. 2117-28.

56. Culos, K.A. and S. Cuellar, Novel targets in the treatment of advanced melanoma: new first-line treatment options. Ann Pharmacother, 2013.

47(4): p. 519-26.

57. Falchook, G.S., et al., Dabrafenib in patients with melanoma, untreated brain metastases, and other solid tumours: a phase 1 dose-escalation trial.

Lancet, 2012. 379(9829): p. 1893-901.

81 58. Eisen, T., et al., Sorafenib in advanced melanoma: a Phase II randomised

discontinuation trial analysis. Br J Cancer, 2006. 95(5): p. 581-6.

59. Buchbinder, E.I., et al., Phase 2 study of sunitinib in patients with metastatic mucosal or acral melanoma. Cancer, 2015. 121(22): p. 4007-15.

60. Kluger, H.M., et al., A phase II trial of dasatinib in advanced melanoma.

Cancer, 2011. 117(10): p. 2202-8.

61. Carvajal, R.D., et al., KIT as a therapeutic target in metastatic melanoma.

JAMA, 2011. 305(22): p. 2327-34.

62. Falchook, G.S., et al., Activity of the MEK Inhibitor Trametinib

(GSK1120212) in Advanced Melanoma in a Phase I, Dose-escalation Trial.

Lancet Oncol, 2012. 13(8): p. 782-9.

63. Kirkwood, J.M., et al., Phase II, open-label, randomized trial of the MEK1/2 inhibitor selumetinib as monotherapy versus temozolomide in patients with advanced melanoma. Clin Cancer Res, 2012. 18(2): p. 555-67.

64. Yang, Z., et al., Rapamycin inhibits lung metastasis of B16 melanoma cells through down-regulating alphav integrin expression and up-regulating apoptosis signaling. Cancer Sci, 2010. 101(2): p. 494-500.

65. Morris, E.J., et al., Discovery of a novel ERK inhibitor with activity in models of acquired resistance to BRAF and MEK inhibitors. Cancer Discov, 2013.

3(7): p. 742-50.

66. Robert , C., et al., Nivolumab in Previously Untreated Melanoma without BRAF Mutation. New England Journal of Medicine, 2015. 372(4): p. 320-330.

67. Larkin, J., F.S. Hodi, and J.D. Wolchok, Combined Nivolumab and

Ipilimumab or Monotherapy in Untreated Melanoma. N Engl J Med, 2015.

373(13): p. 1270-1.

68. Atkins, M.B., et al., High-dose recombinant interleukin 2 therapy for patients with metastatic melanoma: analysis of 270 patients treated between 1985 and 1993. J Clin Oncol, 1999. 17(7): p. 2105-16.

69. Ugurel, S., A. Paschen, and J.C. Becker, Dacarbazine in Melanoma: From a Chemotherapeutic Drug to an Immunomodulating Agent. Journal of Investigative Dermatology, 2013. 133(2): p. 289-292.

70. Yadav, V., et al., The CDK4/6 inhibitor LY2835219 overcomes vemurafenib resistance resulting from MAPK reactivation and cyclin D1 upregulation.

Mol Cancer Ther, 2014. 13(10): p. 2253-63.

71. Lee, B., S. Sandhu, and G. McArthur, Cell cycle control as a promising target in melanoma. Curr Opin Oncol, 2015. 27(2): p. 141-50.

72. Dupin, E., et al., Endothelin 3 induces the reversion of melanocytes to glia through a neural crest-derived glial-melanocytic progenitor. Proc Natl Acad Sci U S A, 2000. 97(14): p. 7882-7.

73. Herlyn, M., Human melanoma: development and progression. Cancer Metastasis Rev, 1990. 9(2): p. 101-12.

74. Bevona, C., et al., Cutaneous melanomas associated with nevi. Arch Dermatol, 2003. 139(12): p. 1620-4; discussion 1624.

75. Shitara, D., et al., Mutational status of nevus associated-melanomas. Br J Dermatol, 2015. 173(3): p. 671-80.

82 76. Pollock, P.M., et al., High frequency of BRAF mutations in nevi. Nat Genet,

2003. 33(1): p. 19-20.

77. Davies, H., et al., Mutations of the BRAF gene in human cancer. Nature, 2002. 417(6892): p. 949-954.

78. Flaherty, K.T. and D.E. Fisher, New strategies in metastatic melanoma:

oncogene-defined taxonomy leads to therapeutic advances. Clin Cancer Res, 2011. 17(15): p. 4922-8.

79. Bennett, D.C., Human melanocyte senescence and melanoma susceptibility genes. Oncogene, 2003. 22(20): p. 3063-9.

80. Michaloglou, C., et al., BRAFE600-associated senescence-like cell cycle arrest of human naevi. Nature, 2005. 436(7051): p. 720-4.

81. Dhillon, A.S., et al., MAP kinase signalling pathways in cancer. Oncogene, 2007. 26(22): p. 3279-90.

82. Serrano, M., et al., Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell, 1997. 88(5): p.

593-602.

83. Zhu, J., et al., Senescence of human fibroblasts induced by oncogenic Raf.

Genes Dev, 1998. 12(19): p. 2997-3007.

84. Palmieri, G., et al., Main roads to melanoma. J Transl Med, 2009. 7: p. 86.

85. Yang, H., et al., Activated MEK cooperates with Cdkn2a and Pten loss to promote the development and maintenance of melanoma. Oncogene, 2017.

86. Wu, H., V. Goel, and F.G. Haluska, PTEN signaling pathways in melanoma.

Oncogene, 0000. 22(20): p. 3113-3122.

87. Delmas, V., et al., Beta-catenin induces immortalization of melanocytes by suppressing p16INK4a expression and cooperates with N-Ras in melanoma development. Genes Dev, 2007. 21(22): p. 2923-35.

88. Webster, M.R. and A.T. Weeraratna, A Wnt-er Migration: The Confusing Role of β-Catenin in Melanoma Metastasis. Science Signaling, 2013. 6(268):

p. pe11-pe11.

89. Chien, A.J., et al., Activated Wnt/ß-catenin signaling in melanoma is associated with decreased proliferation in patient tumors and a murine melanoma model. Proc Natl Acad Sci U S A, 2009. 106(4): p. 1193-8.

90. Jouneau, A., et al., Plasticity of cadherin-catenin expression in the melanocyte lineage. Pigment Cell Res, 2000. 13(4): p. 260-72.

91. Biechele, T.L., et al., Wnt/beta-catenin signaling and AXIN1 regulate apoptosis triggered by inhibition of the mutant kinase BRAFV600E in human melanoma. Sci Signal, 2012. 5(206): p. ra3.

92. Widlund, H.R., et al., β-Catenin–induced melanoma growth requires the downstream target Microphthalmia-associated transcription factor. J Cell Biol, 2002. 158(6): p. 1079-87.

93. Curtin, J.A., et al., Distinct sets of genetic alterations in melanoma. N Engl J Med, 2005. 353(20): p. 2135-47.

94. Levy, C., M. Khaled, and D.E. Fisher, MITF: master regulator of melanocyte development and melanoma oncogene. Trends Mol Med, 2006. 12(9): p.

406-14.

95. Hartman, M.L. and M. Czyz, MITF in melanoma: mechanisms behind its expression and activity. Cell Mol Life Sci, 2015. 72(7): p. 1249-60.

83 96. Fuse, N., et al., Identification of a melanocyte-type promoter of the

microphthalmia-associated transcription factor gene. Biochem Biophys Res Commun, 1996. 219(3): p. 702-7.

97. Hemesath, T.J., et al., microphthalmia, a critical factor in melanocyte development, defines a discrete transcription factor family. Genes Dev, 1994. 8(22): p. 2770-80.

98. Hornyak, T.J., et al., Transcription factors in melanocyte development:

distinct roles for Pax-3 and Mitf. Mech Dev, 2001. 101(1-2): p. 47-59.

99. Cheli, Y., et al., Fifteen-year quest for microphthalmia-associated transcription factor target genes. Pigment Cell Melanoma Res, 2010.

23(1): p. 27-40.

100. Levy, C., M. Khaled, and D.E. Fisher, MITF: master regulator of melanocyte development and melanoma oncogene. Trends in Molecular Medicine.

12(9): p. 406-414.

101. Garraway, L.A., et al., "Lineage addiction" in human cancer: lessons from integrated genomics. Cold Spring Harb Symp Quant Biol, 2005. 70: p. 25-102. Garraway, L.A., et al., 34. Integrative genomic analyses identify MITF as a

lineage survival oncogene amplified in malignant melanoma. Nature, 2005.

436(7047): p. 117-22.

103. Goding, C.R., A picture of Mitf in melanoma immortality. Oncogene, 2011.

30(20): p. 2304-2306.

104. Murakami, H. and H. Arnheiter, Sumoylation modulates transcriptional activity of MITF in a promoter-specific manner. Pigment Cell Res, 2005.

18(4): p. 265-77.

105. de la Serna, I.L., et al., The microphthalmia-associated transcription factor requires SWI/SNF enzymes to activate melanocyte-specific genes. J Biol Chem, 2006. 281(29): p. 20233-41.

106. Du, J., et al., Critical role of CDK2 for melanoma growth linked to its

melanocyte-specific transcriptional regulation by MITF. Cancer Cell, 2004.

6(6): p. 565-76.

107. Carreira, S., et al., Mitf cooperates with Rb1 and activates p21Cip1

expression to regulate cell cycle progression. Nature, 2005. 433(7027): p.

764-9.

108. Loercher, A.E., et al., MITF links differentiation with cell cycle arrest in melanocytes by transcriptional activation of INK4A. J Cell Biol, 2005.

168(1): p. 35-40.

109. Wellbrock, C., et al., Oncogenic BRAF regulates melanoma proliferation through the lineage specific factor MITF. PLoS One, 2008. 3(7): p. e2734.

110. Kuilman, T., et al., The essence of senescence. Genes Dev, 2010. 24(22): p.

2463-79.

111. Zhang, Y., Y. Xiong, and W.G. Yarbrough, ARF promotes MDM2 degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and p53 tumor suppression pathways. Cell, 1998. 92(6): p. 725-34.

112. Pomerantz, J., et al., The Ink4a tumor suppressor gene product, p19Arf, interacts with MDM2 and neutralizes MDM2's inhibition of p53. Cell, 1998.

92(6): p. 713-23.

84 113. Kefford, R.F., et al., Counseling and DNA testing for individuals perceived to

be genetically predisposed to melanoma: A consensus statement of the Melanoma Genetics Consortium. J Clin Oncol, 1999. 17(10): p. 3245-51.

114. Young, R.J., et al., Loss of CDKN2A expression is a frequent event in primary invasive melanoma and correlates with sensitivity to the CDK4/6 inhibitor PD0332991 in melanoma cell lines. Pigment Cell Melanoma Res, 2014.

27(4): p. 590-600.

115. Hayward, N.K., Genetics of melanoma predisposition. Oncogene, 2003.

22(20): p. 3053-62.

116. Puntervoll, H.E., et al., Melanoma prone families with CDK4 germline mutation: phenotypic profile and associations with MC1R variants. J Med Genet, 2013. 50(4): p. 264-70.

117. Loo, J.C., et al., Germline splicing mutations of CDKN2A predispose to melanoma. Oncogene, 2003. 22(41): p. 6387-94.

118. Sharpless, N. and L. Chin, The INK4a//ARF locus and melanoma. Oncogene, 0000. 22(20): p. 3092-3098.

119. Sanki, A., et al., Reduced expression of p16 and p27 is correlated with tumour progression in cutaneous melanoma. Pathology, 2007. 39(6): p.

551-7.

120. Jenkins, N.C., et al., Familial melanoma-associated mutations in p16 uncouple its tumor suppressor functions. J Invest Dermatol, 2013. 133(4):

p. 1043-51.

121. Rizos, H., et al., Mutations in the INK4a/ARF melanoma susceptibility locus functionally impair p14ARF. J Biol Chem, 2001. 276(44): p. 41424-34.

122. Dobrowolski, R., et al., Loss of p14ARF expression in melanoma. Arch Dermatol Res, 2002. 293(11): p. 545-51.

123. Freedberg, D.E., et al., Frequent p16-Independent Inactivation of p14(ARF) in Human Melanoma. J Natl Cancer Inst, 2008. 100(11): p. 784-95.

124. Ko, A., S.Y. Han, and J. Song, Dynamics of ARF regulation that control senescence and cancer. BMB Rep, 2016. 49(11): p. 598-606.

125. Ackermann, J., et al., Metastasizing melanoma formation caused by expression of activated N-RasQ61K on an INK4a-deficient background.

Cancer Res, 2005. 65(10): p. 4005-11.

126. Castellano, M., et al., Restoration of CDKN2A into melanoma cells induces morphologic changes and reduction in growth rate but not anchorage-independent growth reversal. J Invest Dermatol, 1997. 109(1): p. 61-8.

127. George, E., N.L. Polissar, and M. Wick, Immunohistochemical evaluation of p16INK4A, E-cadherin, and cyclin D1 expression in melanoma and Spitz tumors. Am J Clin Pathol, 2010. 133(3): p. 370-9.

128. Mihic-Probst, D., et al., p16 expression in primary malignant melanoma is associated with prognosis and lymph node status. Int J Cancer, 2006.

118(9): p. 2262-8.

129. Gil, J. and G. Peters, Regulation of the INK4b-ARF-INK4a tumour suppressor locus: all for one or one for all. Nat Rev Mol Cell Biol, 2006. 7(9): p. 667-677.

130. Gonzalez, S. and M. Serrano, A new mechanism of inactivation of the INK4/ARF locus. Cell Cycle, 2006. 5(13): p. 1382-4.