全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Delineating Molecular Mechanisms of Squamous Tissue Homeostasis and Neoplasia: Focus on p63

DOI: 10.1155/2013/632028

Full-Text   Cite this paper   Add to My Lib

Abstract:

Mouse models have informed us that p63 is critical for normal epidermal development and homeostasis. The p53/p63/p73 family is expressed as multiple protein isoforms due to a combination of alternative promoter usage and C-terminal alternative splicing. These isoforms can mimic or interfere with one another, and their balance ultimately determines biological outcome in a context-dependent manner. While not frequently mutated, p63, and in particular the ΔNp63 subclass, is commonly overexpressed in human squamous cell cancers. In vitro keratinocytes and murine transgenic and transplantation models have been invaluable in elucidating the contribution of altered p63 levels to cancer development, and studies have identified the roles for ΔNp63 isoforms in keratinocyte survival and malignant progression, likely due in part to their transcriptional regulatory function. These findings can be extended to human cancers; for example, the novel recognition of NFκB/c-Rel as a downstream effector of p63 has identified a role for NFκB/c-Rel in human squamous cell cancers. These models will be critical in enhancing the understanding of the specific molecular mechanisms of cancer development and progression. 1. Introduction p53 is a tumor suppressor that is upregulated and activated across organ systems as a tissue protective stress response mechanism [1]. p63 is a member of the p53 gene family which also includes p73. In contrast to p53, both p63 and p73 exhibit cell-type-specific expression patterns and exert tissue-specific functions [2, 3]. Relevant to this review, p63 plays an essential role in the development and maintenance of normal stratified squamous epithelium. All p53 family members encode multiple protein isoforms that act in overlapping or opposing manners both within and across family members. Given the complexity of the p53 family and the potential for the different family members to mimic or interfere with each other, the balance of p53 family isoforms in a given cellular context can impact the biological outcome. In this review, we highlight how information derived from mouse models has provided insight into molecular mechanisms of normal keratinocyte growth regulation and human cancer pathogenesis. In particular, we focus on the p63 gene, the role of its gene products in normal epidermal development and homeostasis, and how dysregulation of p63 protein expression, which is tightly controlled under normal conditions, contributes to squamous carcinogenesis, not only of the skin, but also in other squamous epithelial cancers such as those of the head and

References

[1]  B. Vogelstein, D. Lane, and A. J. Levine, “Surfing the p53 network,” Nature, vol. 408, no. 6810, pp. 307–310, 2000.
[2]  A. Yang, N. Walker, R. Bronson et al., “p73-Deficient mice have neurological, pheromonal and inflammatory defects but lack spontaneous tumours,” Nature, vol. 404, no. 6773, pp. 99–103, 2000.
[3]  A. Yang, R. Schweitzer, D. Sun et al., “p63 is essential for regenerative proliferation in limb, craniofacial and epithelial development,” Nature, vol. 398, no. 6729, pp. 714–718, 1999.
[4]  A. Yang, M. Kaghad, Y. Wang et al., “p63, a p53 homolog at 3q27-29, encodes multiple products with transactivating, death-inducing, and dominant-negative activities,” Molecular Cell, vol. 2, no. 3, pp. 305–316, 1998.
[5]  M. Kaghad, H. Bonnet, A. Yang et al., “Monoallelically expressed gene related to p53 at 1p36, a region frequently deleted in neuroblastoma and other human cancers,” Cell, vol. 90, no. 4, pp. 809–819, 1997.
[6]  M. Dohn, S. Zhang, and X. Chen, “p63α and ΔNp63α can induce cell cycle arrest and apoptosis and differentially regulate p53 target genes,” Oncogene, vol. 20, no. 25, pp. 3193–3205, 2001.
[7]  E. S. Helton, J. Zhu, and X. Chen, “The unique NH2-terminally deleted (ΔN) residues, the PXXP motif, and the PPXY motif are required for the transcriptional activity of the ΔN variant of p63,” Journal of Biological Chemistry, vol. 281, no. 5, pp. 2533–2542, 2006.
[8]  P. Ghioni, F. Bolognese, P. H. G. Duijf, H. Van Bokhoven, R. Mantovani, and L. Guerrini, “Complex transcriptional effects of p63 isoforms: identification of novel activation and repression domains,” Molecular and Cellular Biology, vol. 22, no. 24, pp. 8659–8668, 2002.
[9]  M. Mangiulli, A. Valletti, M. F. Caratozzolo et al., “Identification and functional characterization of two new transcriptional variants of the human p63 gene,” Nucleic Acids Research, vol. 37, no. 18, pp. 6092–6104, 2009.
[10]  C. D. Thanos and J. U. Bowie, “p53 Family members p63 and p73 are SAM domain-containing proteins,” Protein Science, vol. 8, no. 8, pp. 1708–1710, 1999.
[11]  Z. Serber, H. C. Lai, A. Yang et al., “A C-terminal inhibitory domain controls the activity of p63 by an intramolecular mechanism,” Molecular and Cellular Biology, vol. 22, no. 24, pp. 8601–8611, 2002.
[12]  W. E. Straub, T. A. Weber, B. Sch?fer et al., “The C-terminus of p63 contains multiple regulatory elements with different functions,” Cell Death and Disease, vol. 1, no. 1, article e5, 2010.
[13]  P. Ghioni, Y. D'Alessandra, G. Mansueto et al., “The protein stability and transcriptional activity of p63α are regulated by SUMO-1 conjugation,” Cell Cycle, vol. 4, no. 1, pp. 183–190, 2005.
[14]  M. Rossi, R. I. Aqeilan, M. Neale et al., “The E3 ubiquitin ligase Itch controls the protein stability of p63,” Proceedings of the National Academy of Sciences of the United States of America, vol. 103, no. 34, pp. 12753–12758, 2006.
[15]  F. Galli, M. Rossi, Y. D'Alessandra et al., “MDM2 and Fbw7 cooperate to induce p63 protein degradation following DNA damage and cell differentiation,” Journal of Cell Science, vol. 123, no. 14, pp. 2423–2433, 2010.
[16]  D. Coutandin, F. L?hr, F. H. Niesen et al., “Conformational stability and activity of p73 require a second helix in the tetramerization domain,” Cell Death and Differentiation, vol. 16, no. 12, pp. 1582–1589, 2009.
[17]  E. A. Ratovitski, M. Patturajan, K. Hibi, B. Trink, K. Yamaguchi, and D. Sidransky, “p53 associates with and targets ΔNp63 into a protein degradation pathway,” Proceedings of the National Academy of Sciences of the United States of America, vol. 98, no. 4, pp. 1817–1822, 2001.
[18]  S. Strano, G. Fontemaggi, A. Costanzo et al., “Physical interaction with human tumor-derived p53 mutants inhibits p63 activities,” Journal of Biological Chemistry, vol. 277, no. 21, pp. 18817–18826, 2002.
[19]  M. Osada, H. L. Park, Y. Nagakawa et al., “Differential recognition of response elements determines target gene specificity for p53 and p63,” Molecular and Cellular Biology, vol. 25, no. 14, pp. 6077–6089, 2005.
[20]  C. A. Perez, J. Ott, D. J. Mays, and J. A. Pietenpol, “p63 consensus DNA-binding site: identification, analysis and application into a p63MH algorithm,” Oncogene, vol. 26, no. 52, pp. 7363–7370, 2007.
[21]  K. Ortt and S. Sinha, “Derivation of the consensus DNA-binding sequence for p63 reveals unique requirements that are distinct from p53,” FEBS Letters, vol. 580, no. 18, pp. 4544–4550, 2006.
[22]  A. Yang, Z. Zhu, A. Kettenbach et al., “Genome-wide mapping indicates that p73 and p63 Co-occupy target sites and have similar DNA-binding profiles in vivo,” PLoS ONE, vol. 5, no. 7, Article ID e11572, 2010.
[23]  B. Ory, M. R. Ramsey, C. Wilson et al., “A microRNA-dependent program controls p53-independent survival and chemosensitivity in human and murine squamous cell carcinoma,” Journal of Clinical Investigation, vol. 121, no. 2, pp. 809–820, 2011.
[24]  M. Lanza, B. Marinari, M. Papoutsaki et al., “Cross-talks in the p53 family: ΔNp63 is an anti-apoptotic target for ΔNp73α and p53 gain-of-function mutants,” Cell Cycle, vol. 5, no. 17, pp. 1996–2004, 2006.
[25]  A. A. Mills, B. Zheng, X. J. Wang, H. Vogel, D. R. Roop, and A. Bradley, “p63 is a p53 homologue required for limb and epidermal morphogenesis,” Nature, vol. 398, no. 6729, pp. 708–713, 1999.
[26]  R. Parsa, A. Yang, F. McKeon, and H. Green, “Association of p63 with proliferative potential in normal and neoplastic human keratinocytes,” Journal of Investigative Dermatology, vol. 113, no. 6, pp. 1099–1105, 1999.
[27]  M. I. Koster, B. Marinari, A. S. Payne, P. N. Kantaputra, A. Costanzo, and D. R. Roop, “ΔNp63 knockdown mice: a mouse model for AEC syndrome,” American Journal of Medical Genetics A, vol. 149, no. 9, pp. 1942–1947, 2009.
[28]  K. Hibi, B. Trink, M. Patturajan et al., “AIS is an oncogene amplified in squamous cell carcinoma,” Proceedings of the National Academy of Sciences of the United States of America, vol. 97, no. 10, pp. 5462–5467, 2000.
[29]  S. Wolff, F. Talos, G. Palacios, U. Beyer, M. Dobbelstein, and U. M. Moll, “The α/β carboxy-terminal domains of p63 are required for skin and limb development. New insights from the Brdm2 mouse which is not a complete p63 knockout but expresses p63 γ-like proteins,” Cell Death and Differentiation, vol. 16, no. 8, pp. 1108–1117, 2009.
[30]  M. L. Mikkola, A. Costanzo, I. Thesleff, D. R. Roop, and M. I. Koster, “Treasure or artifact: a decade of p63 research speaks for itself,” Cell Death and Differentiation, vol. 17, no. 1, pp. 180–183, 2010.
[31]  F. Talos, S. Wolff, U. Beyer, M. Dobbelstein, and U. M. Moll, “Brdm2-an aberrant hypomorphic p63 allele,” Cell Death and Differentiation, vol. 17, no. 1, pp. 184–186, 2010.
[32]  R. Shalom-Feuerstein, A. M. Lena, H. Zhou et al., “Δnp63 is an ectodermal gatekeeper of epidermal morphogenesis,” Cell Death and Differentiation, vol. 18, no. 5, pp. 887–896, 2011.
[33]  D. Aberdam and R. Mantovani, “A new p63-deficient mouse model or a fresh look at an old one?” Cell Death and Differentiation, vol. 16, no. 8, pp. 1073–1074, 2009.
[34]  M. I. Koster, S. Kim, A. A. Mills, F. J. DeMayo, and D. R. Roop, “p63 is the molecular switch for initiation of an epithelial stratification program,” Genes and Development, vol. 18, no. 2, pp. 126–131, 2004.
[35]  R. A. Romano, K. Ortt, B. Birkaya, K. Smalley, and S. Sinha, “An active role of the ΔN isoform of p63 in regulating basal keratin genes K5 and K14 and directing epidermal cell fate,” PLoS ONE, vol. 4, no. 5, Article ID e5623, 2009.
[36]  E. Candi, A. Rufini, A. Terrinoni et al., “Differential roles of p63 isoforms in epidermal development: selective genetic complementation in p63 null mice,” Cell Death and Differentiation, vol. 13, no. 6, pp. 1037–1047, 2006.
[37]  X. Su, M. Paris, Y. J. Gi et al., “TAp63 prevents premature aging by promoting adult stem cell maintenance,” Cell Stem Cell, vol. 5, no. 1, pp. 64–75, 2009.
[38]  R. A. Romano, K. Smalley, C. Magraw, et al., “DeltaNp63 knockout mice reveal its indispensable role as a master regulator of epithelial development and differentiation,” Development, vol. 139, pp. 772–782, 2012.
[39]  T. Rinne, H. G. Brunner, and H. Van Bokhoven, “p63-associated disorders,” Cell Cycle, vol. 6, no. 3, pp. 262–268, 2007.
[40]  J. A. McGrath, P. H. G. Duijf, V. Doetsch et al., “Hay-Wells syndrome is caused by heterozygous missense mutations in the SAM domain of p63,” Human Molecular Genetics, vol. 10, no. 3, pp. 221–229, 2001.
[41]  J. Celli, P. Duijf, B. C. J. Hamel et al., “Heterozygous germline mutations in the p53 homolog p63 are the cause of EEC syndrome,” Cell, vol. 99, no. 2, pp. 143–153, 1999.
[42]  H. Van Bokhoven, B. C. J. Hamel, M. Bamshad et al., “p63 gene mutations in EEC syndrome, limb-mammary syndrome, and isolated split hand-split foot malformation suggest a genotype-phenotype correlation,” American Journal of Human Genetics, vol. 69, no. 3, pp. 481–492, 2001.
[43]  H. G. Brunner, B. C. J. Hamel, and H. Van Bokhoven, “The p63 gene in EEC and other syndromes,” Journal of Medical Genetics, vol. 39, no. 6, pp. 377–381, 2002.
[44]  A. Fomenkov, Y. P. Huang, O. Topaloglu et al., “p63α mutations lead to aberrant splicing of keratinocyte growth factor receptor in the Hay-Wells syndrome,” Journal of Biological Chemistry, vol. 278, no. 26, pp. 23906–23914, 2003.
[45]  Y. P. Huang, Y. Kim, Z. Li, T. Fomenkov, A. Fomenkov, and E. A. Ratovitski, “AEC-associated p63 mutations lead to alternative splicing/protein stabilization of p63 and modulation of notch signaling,” Cell Cycle, vol. 4, no. 10, pp. 1440–1447, 2005.
[46]  G. Ferone, H. A. Thomason, D. Antonini, et al., “Mutant p63 causes defective expansion of ectodermal progenitor cells and impaired FGF signalling in AEC syndrome,” EMBO Molecular Medicine, vol. 4, pp. 192–205, 2012.
[47]  G. Ferone, M. R. Mollo, H. A. Thomason, et al., “p63 control of desmosome gene expression and adhesion is compromised in AEC syndrome,” Human Molecular Genetics, vol. 22, pp. 531–543, 2012.
[48]  S. E. Clements, T. Techanukul, J. E. Lai-Cheong, et al., “Mutations in AEC syndrome skin reveal a role for p63 in basement membrane adhesion, skin barrier integrity and hair follicle biology,” British Journal of Dermatology, vol. 167, pp. 134–144, 2012.
[49]  A. B. Truong, M. Kretz, T. W. Ridky, R. Kimmel, and P. A. Khavari, “p63 regulates proliferation and differentiation of developmentally mature keratinocytes,” Genes and Development, vol. 20, no. 22, pp. 3185–3197, 2006.
[50]  K. E. King, R. M. Ponnamperuma, T. Yamashita et al., “ΔNp63α functions as both a positive and a negative transcriptional regulator and blocks in vitro differentiation of murine keratinocytes,” Oncogene, vol. 22, no. 23, pp. 3635–3644, 2003.
[51]  K. E. King, R. M. Ponnamperuma, M. J. Gerdes et al., “Unique domain functions of p63 isotypes that differentially regulate distinct aspects of epidermal homeostasis,” Carcinogenesis, vol. 27, no. 1, pp. 53–63, 2006.
[52]  M. D. Westfall, D. J. Mays, J. C. Sniezek, and J. A. Pietenpol, “The ΔNp63α phosphoprotein binds the p21 and 14-3-3σ promoters in vivo and has transcriptional repressor activity that is reduced by Hay-Wells syndrome-derived mutations,” Molecular and Cellular Biology, vol. 23, no. 7, pp. 2264–2276, 2003.
[53]  M. P. DeYoung, C. M. Johannessen, C. O. Leong, W. Faquin, J. W. Rocco, and L. W. Ellisen, “Tumor-specific p73 up-regulation mediates p63 dependence in squamous cell carcinoma,” Cancer Research, vol. 66, no. 19, pp. 9362–9368, 2006.
[54]  B. C. Nguyen, K. Lefort, A. Mandinova et al., “Cross-regulation between Notch and p63 in keratinocyte commitment to differentiation,” Genes and Development, vol. 20, no. 8, pp. 1028–1042, 2006.
[55]  M. K. Leonard, R. Kommagani, V. Payal, L. D. Mayo, H. N. Shamma, and M. P. Kadakia, “DeltaNp63alpha regulates keratinocyte proliferation by controlling PTEN expression and localization,” Cell Death & Differentiation, vol. 18, pp. 1924–1933, 2011.
[56]  R. J. Richardson, J. Dixon, S. Malhotra et al., “Irf6 is a key determinant of the keratinocyte proliferation-differentiation switch,” Nature Genetics, vol. 38, no. 11, pp. 1329–1334, 2006.
[57]  F. Moretti, B. Marinari, N. L. Iacono et al., “A regulatory feedback loop involving p63 and IRF6 links the pathogenesis of 2 genetically different human ectodermal dysplasias,” Journal of Clinical Investigation, vol. 120, no. 5, pp. 1570–1577, 2010.
[58]  I. Masse, L. Barbollat-Boutrand, M. Molina, et al., “Functional interplay between p63 and p53 controls RUNX1 function in the transition from proliferation to differentiation in human keratinocytes,” Cell Death & Disease, vol. 3, article e318, 2012.
[59]  Y. Xin, Q. Lu, and Q. Li, “IKK1 control of epidermal differentiation is modulated by Notch signaling,” American Journal of Pathology, vol. 178, no. 4, pp. 1568–1577, 2011.
[60]  S. Liu, Z. Chen, F. Zhu, and Y. Hu, “IkappaB kinase alpha and cancer,” Journal of Interferon & Cytokine Research, vol. 32, pp. 152–158, 2012.
[61]  Q. Li, Q. Lu, J. Y. Hwang et al., “IKK1-deficient mice exhibit abnormal development of skin and skeleton,” Genes and Development, vol. 13, no. 10, pp. 1322–1328, 1999.
[62]  E. Candi, A. Terrinoni, A. Rufini et al., “p63 is upstream of IKKα in epidermal development,” Journal of Cell Science, vol. 119, no. 22, pp. 4617–4622, 2006.
[63]  M. I. Koster, D. Dai, B. Marinari et al., “p63 induces key target genes required for epidermal morphogenesis,” Proceedings of the National Academy of Sciences of the United States of America, vol. 104, no. 9, pp. 3255–3260, 2007.
[64]  R. A. Romano, B. Birkaya, and S. Sinha, “A functional enhancer of keratin14 is a direct transcriptional target of ΔNp63,” Journal of Investigative Dermatology, vol. 127, no. 5, pp. 1175–1186, 2007.
[65]  A. M. Lena, R. Cipollone, I. Amelio et al., “Skn-1a/Oct-11 and ΔNp63α exert antagonizing effects on human keratin expression,” Biochemical and Biophysical Research Communications, vol. 401, no. 4, pp. 568–573, 2010.
[66]  S. Borrelli, E. Candi, B. Hu et al., “The p63 target HBP1 is required for skin differentiation and stratification,” Cell Death and Differentiation, vol. 17, no. 12, pp. 1896–1907, 2010.
[67]  G. L. Sen, L. D. Boxer, D. E. Webster, et al., “ZNF750 is a p63 target gene that induces KLF4 to drive terminal epidermal differentiation,” Developmental Cell, vol. 22, pp. 669–677, 2012.
[68]  S. Kim, I. F. Choi, J. R. Quante, L. Zhang, D. R. Roop, and M. I. Koster, “P63 directly induces expression of Alox12, a regulator of epidermal barrier formation,” Experimental Dermatology, vol. 18, no. 12, pp. 1016–1021, 2009.
[69]  A. Chikh, R. N. Matin, V. Senatore, M. Hufbauer, D. Lavery, et al., “iASPP/p63 autoregulatory feedback loop is required for the homeostasis of stratified epithelia,” The EMBO Journal, vol. 30, pp. 4261–4273, 2011.
[70]  R. Yi, M. N. Poy, M. Stoffel, and E. Fuchs, “A skin microRNA promotes differentiation by repressing ‘stemness’,” Nature, vol. 452, no. 7184, pp. 225–229, 2008.
[71]  D. Antonini, M. T. Russo, L. De Rosa, M. Gorrese, L. Del Vecchio, and C. Missero, “Transcriptional repression of miR-34 family contributes to p63-mediated cell cycle progression in epidermal cells,” Journal of Investigative Dermatology, vol. 130, no. 5, pp. 1249–1257, 2010.
[72]  N. Wu, E. Sulpice, P. Obeid, et al., “The miR-17 family links p63 protein to MAPK signaling to promote the onset of human keratinocyte differentiation,” PLoS One, vol. 7, Article ID e45761, 2012.
[73]  L. De Rosa, D. Antonini, G. Ferone et al., “P63 suppresses non-epidermal lineage markers in a bone morphogenetic protein-dependent manner via repression of Smad7,” Journal of Biological Chemistry, vol. 284, no. 44, pp. 30574–30582, 2009.
[74]  C. E. Barton, K. N. Johnson, D. M. Mays et al., “Novel p63 target genes involved in paracrine signaling and keratinocyte differentiation,” Cell Death and Disease, vol. 1, no. 9, article e74, 2010.
[75]  R. A. Ihrie, M. R. Marques, B. T. Nguyen et al., “Perp is a p63-regulated gene essential for epithelial integrity,” Cell, vol. 120, no. 6, pp. 843–856, 2005.
[76]  V. G. Beaudry, R. A. Ihrie, S. B. Jacobs, et al., “Loss of the desmosomal component perp impairs wound healing in vivo,” Dermatology Research and Practice, vol. 2010, Article ID 759731, 11 pages, 2010.
[77]  C. E. Barbieri, C. A. Perez, K. N. Johnson, K. A. Ely, D. Billheimer, and J. A. Pietenpol, “IGFBP-3 is a direct target of transcriptional regulation by ΔNp63α in squamous epithelium,” Cancer Research, vol. 65, no. 6, pp. 2314–2320, 2005.
[78]  D. K. Carroll, J. S. Carroll, C. O. Leong et al., “p63 regulates an adhesion programme and cell survival in epithelial cells,” Nature Cell Biology, vol. 8, no. 6, pp. 551–561, 2006.
[79]  L. Hayflick and P. S. Moorhead, “The serial cultivation of human diploid cell strains,” Experimental Cell Research, vol. 25, no. 3, pp. 585–621, 1961.
[80]  L. Hayflick, “The limited in vitro lifetime of human diploid cell strains,” Experimental Cell Research, vol. 37, no. 3, pp. 614–636, 1965.
[81]  G. Pellegrini, E. Dellambra, O. Golisano et al., “p63 identifies keratinocyte stem cells,” Proceedings of the National Academy of Sciences of the United States of America, vol. 98, no. 6, pp. 3156–3161, 2001.
[82]  M. Senoo, F. Pinto, C. P. Crum, and F. McKeon, “p63 is essential for the proliferative potential of stem cells in stratified epithelia,” Cell, vol. 129, no. 3, pp. 523–536, 2007.
[83]  V. Rivetti d, A. M. Lena, M. Nicoloso, et al., “p63-microRNA feedback in keratinocyte senescence,” Proceedings of the National Academy of Sciences of the United States of America, vol. 109, pp. 1133–1138, 2012.
[84]  W. M. Keyes, Y. Wu, H. Vogel, X. Guo, S. W. Lowe, and A. A. Mills, “p63 deficiency activates a program of cellular senescence and leads to accelerated aging,” Genes and Development, vol. 19, no. 17, pp. 1986–1999, 2005.
[85]  E. R. Flores, S. Sengupta, J. B. Miller et al., “Tumor predisposition in mice mutant for p63 and p73: evidence for broader tumor suppressor functions for the p53 family,” Cancer Cell, vol. 7, no. 4, pp. 363–373, 2005.
[86]  L. Ha, R. M. Ponnamperuma, S. Jay, M. S. Ricci, and W. C. Weinberg, “Dysregulated δNp63α inhibits expression of ink4a/arf, blocks senescence, and promotes malignant conversion of keratinocytes,” PLoS ONE, vol. 6, no. 7, Article ID e21877, 2011.
[87]  X. Su, M. S. Cho, Y. J. Gi, B. A. Ayanga, C. J. Sherr, and E. R. Flores, “Rescue of key features of the p63-null epithelial phenotype by inactivation of Ink4a and Arf,” The EMBO Journal, vol. 28, no. 13, pp. 1904–1915, 2009.
[88]  M. Sommer, N. Poliak, S. Upadhyay et al., “ΔNp63α overexpression induces downregulation of Sirt1 and an accelerated aging phenotype in the mouse,” Cell Cycle, vol. 5, no. 17, pp. 2005–2011, 2006.
[89]  J. G. Toma, I. A. McKenzie, D. Bagli, and F. D. Miller, “Isolation and characterization of multipotent skin-derived precursors from human skin,” Stem Cells, vol. 23, no. 6, pp. 727–737, 2005.
[90]  K. Hagiwara, M. G. McMenamin, K. Miura, and C. C. Harris, “Mutational analysis of the p63/p73L/p51/p40/CUSP/KET gene in human cancer cell lines using intronic primers,” Cancer Research, vol. 59, no. 17, pp. 4165–4169, 1999.
[91]  J. N. Glickman, A. Yang, A. Shahsafaei, F. McKeon, and R. D. Odze, “Expression of p53-related protein p63 in the gastrointestinal tract and in esophageal metaplastic and neoplastic disorders,” Human Pathology, vol. 32, no. 11, pp. 1157–1165, 2001.
[92]  H. Hu, S. H. Xia, A. D. Li et al., “Elevated expression of p63 protein in human esophageal squamous cell carcinomas,” International Journal of Cancer, vol. 102, no. 6, pp. 580–583, 2002.
[93]  T. Crook, J. M. Nicholls, L. Brooks, J. O'Nions, and M. J. Allday, “High level expression of ΔN-p63: a mechanism for the inactivation of p53 in undifferentiated nasopharyngeal carcinoma (NPC)?” Oncogene, vol. 19, no. 30, pp. 3439–3444, 2000.
[94]  J. S. Reis-Filho, B. Torio, A. Albergaria, and F. C. Schmitt, “p63 expression in normal skin and usual cutaneous carcinomas,” Journal of Cutaneous Pathology, vol. 29, no. 9, pp. 517–523, 2002.
[95]  D. A. Wrone, S. Yoo, L. K. Chipps, and R. L. Moy, “The expression of p63 in actinic keratosis, seborrheic keratosis, and cutaneous squamous cell carcinomas,” Dermatologic Surgery, vol. 30, no. 10, pp. 1299–1302, 2004.
[96]  P. P. Massion, P. M. Taflan, S. M. J. Rahman et al., “Significance of p63 amplification and overexpression in lung cancer development and prognosis,” Cancer Research, vol. 63, no. 21, pp. 7113–7121, 2003.
[97]  G. Pelosi, F. Pasini, C. O. Stenholm et al., “p63 immunoreactivity in lung cancer: yet another player in the development of squamous cell carcinomas?” Journal of Pathology, vol. 198, no. 1, pp. 100–109, 2002.
[98]  K. Nylander, P. J. Coates, and P. A. Hall, “Characterization of the expression pattern of p63 alpha and delta Np63 alpha in benign and malignant oral epithelial lesions,” International Journal of Cancer, vol. 87, pp. 368–372, 2000.
[99]  M. Senoo, I. Tsuchiya, Y. Matsumura et al., “Transcriptional dysregulation of the p73L/p63/p51/p40/KET gene in human squamous cell carcinomas: expression of δNp73L, a novel dominant-negative isoform, and loss of expression of the potential tumour suppressor p51,” British Journal of Cancer, vol. 84, no. 9, pp. 1235–1241, 2001.
[100]  R. J. Richardson, J. Dixon, S. Malhotra et al., “Irf6 is a key determinant of the keratinocyte proliferation-differentiation switch,” Nature Genetics, vol. 38, no. 11, pp. 1329–1334, 2006.
[101]  E. Botti, G. Spallone, F. Moretti, et al., “Developmental factor IRF6 exhibits tumor suppressor activity in squamous cell carcinomas,” Proceedings of the National Academy of Sciences of the United States of America, vol. 108, pp. 13710–13715, 2011.
[102]  G. Wu, M. Osada, Z. Guo et al., “ΔNp63α up-regulates the Hsp70 gene in human cancer,” Cancer Research, vol. 65, no. 3, pp. 758–766, 2005.
[103]  M. Patturajan, S. Nomoto, M. Sommer et al., “ΔNp63 induces β-catenin nuclear accumulation and signaling,” Cancer Cell, vol. 1, no. 4, pp. 369–379, 2002.
[104]  M. Herfs, P. Hubert, M. Suarez-Carmona et al., “Regulation of p63 isoforms by snail and slug transcription factors in human squamous cell carcinoma,” American Journal of Pathology, vol. 176, no. 4, pp. 1941–1949, 2010.
[105]  C. A. Di, A. Troiano, M. O. Di, et al., “The p63 protein isoform DeltaNp63alpha modulates Y-box binding protein 1 in its subcellular distribution and regulation of cell survival and motility genes,” The Journal of Biological Chemistry, vol. 287, pp. 30170–30180, 2012.
[106]  R. Kommagani, T. M. Caserta, and M. P. Kadakia, “Identification of vitamin D receptor as a target of p63,” Oncogene, vol. 25, no. 26, pp. 3745–3751, 2006.
[107]  R. Kommagani, M. K. Leonard, S. Lewis, R. A. Romano, S. Sinha, and M. P. Kadakia, “Regulation of VDR by ΔNp63α is associated with inhibition of cell invasion,” Journal of Cell Science, vol. 122, no. 16, pp. 2828–2835, 2009.
[108]  M. Adorno, M. Cordenonsi, M. Montagner et al., “A mutant-p53/Smad complex opposes p63 to empower TGFβ-induced metastasis,” Cell, vol. 137, no. 1, pp. 87–98, 2009.
[109]  C. E. Barbieri, L. J. Tang, K. A. Brown, and J. A. Pietenpol, “Loss of p63 leads to increased cell migration and up-regulation of genes involved in invasion and metastasis,” Cancer Research, vol. 66, no. 15, pp. 7589–7597, 2006.
[110]  R. Kalluri and R. A. Weinberg, “The basics of epithelial-mesenchymal transition,” Journal of Clinical Investigation, vol. 119, no. 6, pp. 1420–1428, 2009.
[111]  K. M. Liefer, M. I. Koster, X. J. Wang, A. Yang, F. McKeon, and D. R. Roop, “Down-regulation of p63 is required for epidermal UV-B-induced apoptosis,” Cancer Research, vol. 60, no. 15, pp. 4016–4020, 2000.
[112]  M. Papoutsaki, F. Moretti, M. Lanza et al., “A p38-dependent pathway regulates ΔNp63 DNA binding to p53-dependent promoters in UV-induced apoptosis of keratinocytes,” Oncogene, vol. 24, no. 46, pp. 6970–6975, 2005.
[113]  K. L. Schavolt and J. A. Pietenpol, “p53 and Δnp63α differentially bind and regulate target genes involved in cell cycle arrest, DNA repair and apoptosis,” Oncogene, vol. 26, no. 42, pp. 6125–6132, 2007.
[114]  R. Zangen, E. Ratovitski, and D. Sidransky, “ΔNp63α levels correlate with clinical tumor response to cisplatin,” Cell Cycle, vol. 4, no. 10, pp. 1313–1315, 2005.
[115]  A. Fomenkov, R. Zangen, Y. P. Huang et al., “RACK1 and stratifin target ΔNp63α for a proteasome degradation in head and neck squamous cell carcinoma cells upon DNA damage,” Cell Cycle, vol. 3, no. 10, pp. 1285–1295, 2004.
[116]  A. Chatterjee, X. Chang, T. Sen, R. Ravi, A. Bedi, and D. Sidransky, “Regulation of p53 family member isoform ΔNp63α by the nuclear factor-κB targeting kinase IκB kinase β,” Cancer Research, vol. 70, no. 4, pp. 1419–1429, 2010.
[117]  T. Sen, X. Chang, D. Sidransky, and A. Chatterjee, “Regulation of ΔNp63α by NFκB,” Cell Cycle, vol. 9, no. 24, pp. 4841–4847, 2010.
[118]  M. Yuan, P. Luong, C. Hudson, K. Gudmundsdottir, and S. Basu, “C-Abl phosphorylation of Δnp63α is critical for cell viability,” Cell Death and Disease, vol. 1, no. 1, article e16, 2010.
[119]  J. W. Rocco, C. O. Leong, N. Kuperwasser, M. P. DeYoung, and L. W. Ellisen, “p63 mediates survival in squamous cell carcinoma by suppression of p73-dependent apoptosis,” Cancer Cell, vol. 9, no. 1, pp. 45–56, 2006.
[120]  M. R. Ramsey, L. He, N. Forster, B. Ory, and L. W. Ellisen, “Physical association of HDAC1 and HDAC2 with p63 mediates transcriptional repression and tumor maintenance in squamous cell carcinoma,” Cancer Research, vol. 71, no. 13, pp. 4373–4379, 2011.
[121]  S. Frame, R. Crombie, J. Liddell et al., “Epithelial carcinogenesis in the mouse: correlating the genetics and the biology,” Philosophical Transactions of the Royal Society B, vol. 353, no. 1370, pp. 839–845, 1998.
[122]  W. M. Keyes, H. Vogel, M. I. Koster et al., “p63 heterozygous mutant mice are not prone to spontaneous or chemically induced tumors,” Proceedings of the National Academy of Sciences of the United States of America, vol. 103, no. 22, pp. 8435–8440, 2006.
[123]  X. Su, D. Chakravarti, M. S. Cho et al., “TAp63 suppresses metastasis through coordinate regulation of Dicer and miRNAs,” Nature, vol. 467, no. 7318, pp. 986–990, 2010.
[124]  W. M. Keyes, M. Pecoraro, V. Aranda et al., “Δnp63α is an oncogene that targets chromatin remodeler Lsh to drive skin stem cell proliferation and tumorigenesis,” Cell Stem Cell, vol. 8, no. 2, pp. 164–176, 2011.
[125]  Q. Li, S. A. Sambandam, H. J. Lu, A. Thomson, S. H. Kim, et al., “14-3-3sigma and p63 play opposing roles in epidermal tumorigenesis,” Carcinogenesis, vol. 32, pp. 1782–1788, 2011.
[126]  T. D. Gilmore, “Introduction to NF-κB: players, pathways, perspectives,” Oncogene, vol. 25, no. 51, pp. 6680–6684, 2006.
[127]  G. Courtois and T. D. Gilmore, “Mutations in the NF-κB signaling pathway: implications for human disease,” Oncogene, vol. 25, no. 51, pp. 6831–6843, 2006.
[128]  C. Van Waes, “Nuclear factor-κB in development, prevention, and therapy of cancer,” Clinical Cancer Research, vol. 13, no. 4, pp. 1076–1082, 2007.
[129]  C. S. Seitz, Q. Lin, H. Deng, and P. A. Khavari, “Alterations in NF-κB function in transgenic epithelial tissue demonstrate a growth inhibitory role for NF-κB,” Proceedings of the National Academy of Sciences of the United States of America, vol. 95, no. 5, pp. 2307–2312, 1998.
[130]  R. Gugasyan, A. Voss, G. Varigos et al., “The transcription factors c-rel and RelA control epidermal development and homeostasis in embryonic and adult skin via distinct mechanisms,” Molecular and Cellular Biology, vol. 24, no. 13, pp. 5733–5745, 2004.
[131]  A. Loercher, T. L. Lee, J. L. Ricker, et al., “Nuclear factor-kappaB is an important modulator of the altered gene expression profile and malignant phenotype in squamous cell carcinoma,” Cancer Research, vol. 64, pp. 6511–6523, 2004.
[132]  M. Dajee, M. Lazarov, J. Y. Zhang et al., “NF-κB blockade and oncogenic Ras trigger invasive human epidermal neoplasia,” Nature, vol. 421, no. 6923, pp. 639–643, 2003.
[133]  M. A. Sovak, R. E. Bellas, D. W. Kim et al., “Aberrant nuclear factor-κb/Rel expression and the pathogenesis of breast cancer,” Journal of Clinical Investigation, vol. 100, no. 12, pp. 2952–2960, 1997.
[134]  K. E. King, R. M. Ponnamperuma, C. Allen et al., “The p53 homologue ΔNp63α interacts with the nuclear factor-κB pathway to modulate epithelial cell growth,” Cancer Research, vol. 68, no. 13, pp. 5122–5131, 2008.
[135]  H. Lu, X. Yang, P. Duggal, et al., “TNF-alpha promotes c-REL/DeltaNp63alpha interaction and TAp73 dissociation from key genes that mediate growth arrest and apoptosis in head and neck cancer,” Cancer Research, vol. 71, pp. 6867–6877, 2011.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133

WeChat 1538708413