r a... IV‘IIL‘ Awflfll'wfiflf A”, "I \Ermiv 54h P a; . y k [Rm 2825 . -‘ ‘§)£ I a}? 1“}? 1:4 \mufl. -’ \3 3 '\ ~k «E This is to certify that the dissertation entitled THE ROLE OF SPROUTY-2 IN THE MALIGNANT TRANSFORMATION OF HUMAN FIBROBLASTS BY HRAS ONCOGENE presented by Piro Lito has been accepted towards fulfillment of the requirements for the PhD degree in Biochemistry and Molecular Biology 7 \Xmlé ~~ /Z((/{1\;/{M Major Professor’ 3 Signature § / 8’ /04. Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/07 p:/C|RC/DaleDue.indd-p.1 THE ROLE OF SPROUTY-Z IN THE MALIGNANT TRANSFORMATION OF HUMAN FIBROBLASTS BY HRAS ONCOGENE By Piro Lito A DISSERTATION Submitted to Michigan State Univerisity in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Biochemistry and Molecular Biology 2006 Abstract THE ROLE OF SPROUTY-2 IN THE MALIGNANT TRANSFORMATION OF HUMAN FIBROBLASTS BY HRAS ONCOGENE By Piro Lito Sprouty-2 (Spry2) plays a regulatory role in the signaling pathways induced by a number of growth factor receptors. One aspect of the function of Spry2 is to prevent the c-Cbl-induced degradation of epidermal growth factor receptor (EGFR). I found that a human fibroblast cell strain, malignantly transformed by HRasWZ, exhibited an increase in the expression of Spry2, compared to its parental cell strain. This correlated with an increase in the level of EGFR protein in HRas-transformed cells compared to their parental cells. EGFR activity was required if HRas-transformed cells were to exhibit growth factor independence. To determine whether Spry2 plays a role in HRas-transformation, l down- regulated the expression of Spry2 using Spry2-specific shRNA. The cell strains with down-regulated Spry2 exhibited not only decreased levels of EGFR, but also decreased levels of ERK activation. I also demonstrated that HRas and Spry2 interact in HRas-transformed fibroblasts, and that HRas interacts with c-Cbl and CIN85 in a Spry2-dependent fashion, suggesting that HRas regulates the turnover of EGFR through Spry2. HRas-transformed cells with down-regulated levels of Spry2 failed to form tumors when injected into athymic mice. Expression V12 of Spry2 in immortalized human fibroblasts independently of HRas , was not sufficient to induce the malignant transformation of these cells, suggesting that the role of Spry2 in cancer formation is dependent on HRas oncogene. Ras is reported to reduce the sensitivity of cells to DNA damage-induced apoptosis. In the HRas-transformed cell strain that expresses high levels of Spry2, the ability of HRas to prevent UV-induced apoptosis was dependent on intact Pl3K- and Rac1-activity. By comparing HRas-transformed cells with endogenous levels of Spry2 to HRas-transformed cells with down-regulated Spry2, I found that Spry2 sustained the activation of PI3K and Akt. Furthermore, I demonstrated that Spry2 sustained the activation of Rac1 by HRas, in part by modulating the interaction between HRas and Tiam1, a GDP-releasing factor for R301. Consistent with these findings, the down-regulation of Spry2 in HRas-transformed cells resulted in increased levels of UV-induced apoptosis. The down-regulation of Spry2, also, resulted in an increase in the level of p53, paralleled by a decrease in the level of MDM2 phosphorylated at Ser166, an Akt specific site. Expression of Spry2 in immortalized human fibroblasts resulted in a decrease in the level of UV-induced apoptosis, and in a decrease in the level of p53 protein. Taken together these results indicate that Spry2 facilitates the regulation of EGFR by HRas, and is necessary for the malignant transformation of human fibroblasts by HRas oncogene. The data also suggest that Spry2 is an important mediator of survival signals in HRas-transformed cells, because it inhibits DNA-damage induced- apoptosis through the regulation of the p53/MDM2 pathway. Dedicated to Alex, Zana and Melina “Science and everyday life cannot and should not be separated. Science, for me, gives a partial explanation to life. In so far as it goes, it is based on fact, experience and experiment”. Rosalind Franklin, in a letter to her father, 1940 Acknowledgements I wish to thank Dr. Justin McCormick, my thesis advisor, and Dr. Veronica Maher, the co-director of the Carcinogenesis Laboratory, for their constant support and guidance. I will never forget their passion for research, their endurance and perseverance in addressing scientific questions, as well as their readiness to help others overcome their problems. In addition, I have benefited greatly by the experience and dedication that they bring to the Medical Scientist Training Program. I wish to thank the members of my guidance committee, Dr. Kathleen Gallo, Dr. William Henry and Dr. Donald Jump, for their advice through the years. I would like to particularly thank Dr. Henry, who has been a source of support and advice for me since the days of undergraduate biochemistry classes. I sincerely thank past and current members of the Carcinogenesis Laboratory for their assistance: Dr. Michele Battle, Dr. Zhenjun Lou, Dr. ZiQiang Li, Dr. Jackie Dao, Terry McManus, Yun Wang, Dr. Igor Zlatkin, Clarissa Dallas, Jessica Apostol, Jie Zhang, Rick Tobby, Suzanne Kohler, Bethany Heinlen and Katherine Bergdol. I thank Dr. Susanne Kleff for her help in getting this project started, Dr. Sandra O’Reilly for her help with the tumor studies, and Dr. Katheryn Meek for her advice on my work and manuscript preparation. I thank Dr. Kristin McNally and Dan Appledom for their friendship and helpful scientific discussions. I will not forget our escapes form the lab and our adventures in East Lansing. I would also like to thank the various undergraduate students that have helped me with my dissertation research. In particular, I would like to acknowledge Bryan Mets for his dedication, determination and hard work. He has been very important to my research and I sincerely hope that Bryan pursues a career in science. I thank my parents from the bottom of my heart. Their experiences in life so far have taught me that nothing is out of reach. If anything that I do in life has value, it is because of them. I thank my sister for being the happy person that she is, and for making me smile in times of stress. Finally, I thank my grandparents for inspiring in me the love of research and medicine. vi TABLE OF CONTENTS LIST OF FIGURES .............................................................................................. x LIST OF TABLES .............................................................................................. xii ABBREVIATIONS ............................................................................................ xiii INTRODUCTION .................................................................................................. 1 REFERENCES ...................................................................................................... 7 CHAPTER I: REVIEW OF LITERATURE .......................................................... 10 CANCER AS A GENETIC DISORDER ....................................................................... 1O Cancer-related genes .................................................................................. 10 Characteristics of cancer cells ..................................................................... 14 Uncontrolled proliferation ......................................................................... 15 Aberrant cell cycle regulation ................................................................... 19 Limitless replication potential .................................................................... 23 Evasion of apoptosis ................................................................................ 24 Inefficient DNA repair mechanisms .......................................................... 27 Genetic instability ..................................................................................... 29 Angiogenesis ............................................................................................ 3O Invasion and metastasis ........................................................................... 32 MSU-1 LINEAGE AS MODEL SYSTEM TO STUDY MALIGNANT TRANSFORMATION ........ 34 RAS GTPASE .................................................................................................... 44 Catalytic function ......................................................................................... 45 Regulation of Ras activity ............................................................................ 46 GEFS and activation of Ras ...................................................................... 49 GAPS and inactivation of Ras ................................................................... 51 Posttranslational regulation ...................................................................... 51 Res Effectors ............................................................................................... 52 Ref ............................................................................................................ 53 PI3K ......................................................................................................... 57 Tiam1 ....................................................................................................... 59 RaIGEFs ................................................................................................... 60 MEKK1 ..................................................................................................... 61 Rin1 .......................................................................................................... 61 AF-6 ......................................................................................................... 62 RASSF ..................................................................................................... 62 Cellular functions mediated by Ras ............................................................. 62 Regulation of cell cycle and proliferation .................................................. 62 Regulation of protein synthesis ................................................................ 64 Regulation of apoptosis ............................................................................ 65 Regulation of the actin cytoskeleton ......................................................... 66 Regulation of cellular migration ................................................................ 67 Regulation of invasion and metastasis ..................................................... 68 SPROUTY .......................................................................................................... 69 Sprouty structure ......................................................................................... 70 vii Sprouty expression ...................................................................................... 74 Sprouty localization ...................................................................................... 76 Regulation of Sprouty .................................................................................. 78 Phosphorylation of the N-terminus of Sprouty .......................................... 78 Sprouty-2 kinase ...................................................................................... 81 Sprouty-2 phosphatase ............................................................................ 81 Sprouty-2 ubiquitination ............................................................................ 82 Phosphorylation of the C-terminus of Sprouty-2 ....................................... 83 Cellular functions of Sprouty ........................................................................ 84 Inhibition of receptor tyrosine kinase signaling ......................................... 84 Activation of epidermal growth factor receptor signaling .......................... 90 Regulation of integrin-mediated cell spreading by Sprouty-4 ................... 94 Regulation of cell migration by Sprouty .................................................... 94 Sprouty deficient mice ............................................................................. 95 Sprouty-1" ............................................................................................ 95 Sprouty-2" ............................................................................................ 96 Sprouty proteins in cancer ........................................................................ 98 REFERENCES .................................................................................................. 101 CHAPTER II. SPROUTY 2 IS NECESSARY FOR TUMOR FORMATION BY HRAS ONCOGENE-TRANSFORMED HUMAN FIBROBLASTS ................... 134 ABSTRACT ...................................................................................................... 1 35 INTRODUCTION ................................................................................................ 1 36 RESULTS ........................................................................................................ 1 38 Determination of the Expression of Spry2 in HRas-transformed Cells. ..... 138 Effect of Spry2 on Tumor Formation by HRas-transformed Cells. ............. 141 Effect of HRas-transformation on the Level of E GFR protein. ................... 145 Effect of Depletion of Spry2 Protein on the Level of EGFR. ...................... 149 Interaction of Spry2 with HRas. ................................................................. 152 Interaction of HRas with c-Cbl and CIN85 in a Spry2-dependent Fashion. 155 Interaction of HRas with c-Cbl and CIN85 in a Spry2-dependent Fashion. 155 Effect of Spry2 Expression in Immortalized Human Fibroblasts. ............... 155 DISCUSSION .................................................................................................... 161 MATERIALS AND METHODS ............................................................................... 164 Cells and Cell Culture. ............................................................................... 164 Northern Blot Analysis. .............................................................................. 164 Western Blot Analysis. ............................................................................... 164 Preparation of Spry2-shRNA Constructs. .................................................. 165 Stable Infection. ......................................................................................... 166 AG1478 Inhibitor Study .............................................................................. 167 lmmunoprecipitation Reactions. ................................................................ 167 Anchorage independence assay. .............................................................. 167 Tumorigenicity Assay ................................................................................. 168 Ras Activation Assay. ................................................................................ 168 ACKNOWLEDGEMENTS ..................................................................................... 1 69 REFERENCES .................................................................................................. 1 70 viii CHAPTER III: SPROUTY-2 PREVENTS APOPTOSIS IN HRAS- TRANSFORMED HUMAN FIBROBLASTS ..................................................... 174 ABSTRACT ...................................................................................................... 175 INTRODUCTION ................................................................................................ 1 76 RESULTS ........................................................................................................ 1 79 Effect of HRas oncogene-transformation on DNA-damage induced apoptosis .................................................................................................................. 1 79 Effect of Spry2 on the activation of the PI3K pathway in HRas-transformed cells ........................................................................................................... 183 Effect of Spry2 on the activation of Rac1 in HRas-transformed cells ......... 186 Effect of Spry2 on the induction of apoptosis in response to DNA-damage190 Effect of Spry2 on the MDM2/p53 pathway ............................................... 193 DISCUSSION .................................................................................................... 197 MATERIAL AND METHODS ................................................................................. 200 Cells and Cell Culture. ............................................................................... 200 Apoptosis assay ........................................................................................ 200 Western blotting ......................................................................................... 201 Rac1 activation .......................................................................................... 201 Staining for stress fibers ............................................................................ 202 lmmunoprecipitation Reactions. ................................................................ 202 REFERENCES .................................................................................................. 203 APPENDIX A: ANALYSIS OF EXPRESSIONAL CHANGES BETWEEN MSU- 1.0, MSU-1.1 AND PH3MT CELLS ................................................................. 208 INTRODUCTION ................................................................................................ 209 RESULTS ........................................................................................................ 212 Comparison of the RNA expression profiles of MSU 1.0, MSU 1.1 and PH3MT cells .............................................................................................. 212 Northern blot analysis and confirmation of the gene chip data .................. 219 DISCUSSION .................................................................................................... 226 MATERIALS AND METHODS ............................................................................... 229 Total RNA extraction .................................................................................. 229 Gene Chip Analysis ................................................................................... 229 Data analysis ............................................................................................. 230 REFERENCES .................................................................................................. 232 APPENDIX B: THE ROLE OF SPROUTY-2 IN THE MALIGNANT PHENOTYPE OF PATIENT DERIVED FIBROSARCOMA CELL LINES ............................... 234 INTRODUCTION ................................................................................................ 235 RESULTS ........................................................................................................ 236 The role of Spry2 in the malignant phenotype of patient derived fibrosarcoma cell lines ..................................................................................................... 236 Effect of Spry2 on E GF-induced cell cycle progression ............................. 241 DISCUSSION .................................................................................................... 243 MATERIAL AND METHODS ................................................................................. 246 Cell cycle analysis ..................................................................................... 246 List of Figures Chapter I Figure 1. MSU 1 lineage of human fibroblasts ................................................... 35 Figure 2. MSU 1 lineage as a tool to study malignant transformation ................ 39 Figure 3. Res GTPase as molecular switch ....................................................... 47 Figure 4. Res effector pathways ......................................................................... 54 Figure 5. Structure of Spry2 ............................................................................... 71 Figure 6. Regulation of RTK signaling by Spry ................................................... 85 Chapter II Figure 1. Expression profile of Spry2 in RaS-transtormed cells and in patient derived cancer cells. ......................................................................................... 139 Figure 2. Effect of Spry2 on the anchorage independent growth of HRas- transforrned fibroblasts ...................................................................................... 142 Figure 3. Effect of HRas-transformation on the level of EGFR .......................... 147 Figure 4. Effect of Spry2 depletion on the level of EGFR in HRas-transformed cells ................................................................................................................... 150 Figure 5. Interaction of HRas with Spry2 and Spry2 binding-partners c-Cbl and CIN85 ................................................................................................................ 153 Figure 6. Effect of Spry2 expression in immortalized human fibroblasts ........... 156 Chapter III Figure 1. Effect of HRas-transformation on UV-induced apoptosis .................. 180 Figure 2. Effect of Spry2 on PI3K signaling in HRas-transformed cells ............ 184 Figure 3. Effect of Spry2 on the activation Rac1 in HRas-transformed cells ....187 Figure 4. Effect of Spr2 on UV-induced apoptosis ........................................... 191 Figure 5. Effect of Spry2 on the MDM2/p53 pathway ....................................... 194 Appendix A Figure 1. Gene Chip comparison of MSU1.1 and PH3MT cells to MSU-1.0 cells 213 Figure 2. K-means clustering ............................................................................ 217 Figure 3. Northern analysis that validates the gene chip data for Spry2, on, fib5 and s.jag1 ........................................................................................................ 224 Appendix B Figure 1. Down regulation of Spry2 in fibrosarcomas with wild type Ras or NRasQ59 expression ......................................................................................... 237 Figure 2. Down regulation of Spry2 in VleFT cells delays progression through the cell cycle ..................................................................................................... 244 xi List of Tables Chapter II Table l Tumorigenicity of the cell strains with down-regulated Spry2 ............... 146 Table II Tumorigenicity of the cell strains expressing Spry2 ............................ 160 Appendix A Table I Grouping of differentially expressed genes according to their function .220 Appendix B Table | Tumorigenicity of HT1080 cell lines with down-regulated Spry2 .......... 240 Table II Tumorigenicity of VleFT cell lines with down-regulated Spry2 ........... 242 xii Abbreviations Rb ................................................................................................. Retinoblastoma APC ............................................................................ Adanomatus Polyposis Coli EGF ................................................................................ Epidermal Growth Factor PDGF .................................................................... Platelet-derived Growth Factor EGFR ............................................................................................... EGF-receptor PDGFR .......................................................................................... PDGF-receptor RTK ............................................................................. Receptor Tyrosine Kinases SOS ........................................................................................... Son of Sevenless GTP ........................................ , ........................................... Guanine Triphosphate GDP .................................................................................... Guanine Diphosphate ERK ............................................................ Extracellular signal-Regulated Kinase TGFoc ...................................................................... Transforming Growth Factor a NFKB ........................................................................... Nuclear Factor kappa Beta STAT ................................................ Signal Transducer Activator of Transcription elF .............................................................................................. Elongation Factor CDK ............................................................................ Cyclin Ddependent Kinase ATM ........................................................................... Ataxia telangectasla mutant MDM2 ............................................................................... Murine double mutant-2 ALT ............................................................... Alternative lengthening of telomeres TNF ..................................................................................... Tumor necrosis facotr IAP ......................................................................... Inhibitor of Apoptosis Proteins XP ................................................................................. Xeroderma Pigmentosum xiii VEGF ............................................................... Vascular endothelial growth factor HIF1 ............................................................................. Hypoxia inducible factor-1 vHL ........................................................................................... Von-Hippel Lindau FAK .................................................................................... Focal adhesion kinase ECM ........................................................................................ Extracellular Matrix MMP .................................................................................. Matrix Metalloprotease BPDE ...................................................................... Benzo-A-pyrene—diol-epoxide ENU ......................................................................................... Ethyl-nitrosourea a HGF ............................................................................... Hepatocyte growth factor GEF ............................................................. Guanine nucleotide exchange factors GDI ......................................................... Guanine nucleotide dissociation inhibitor GAP ............................................................................ GTPase activating proteins DH .................................................................................................... Dbl homology PH .......................................................................................... Pleckstrin homology IGFR ....................................................................... Insulin growth factor receptor PIP ...................................................................... Phosphatidyl inositol phosphate RBD ...................................................................................... Ras binding domain RA ................................................................................................ Ras association CR .............................................................................................. conserved region MAPK ................................................................. mitogen-activated protein kinase MAPKK ................................................... mitogen-activated protein kinase Kinase MAPKKK .................................... mitogen-activated protein kinase Kinase Kinase PI3K ...................................................................... Phosphatidyl inositol-3-kinase xiv mTOR ............................................................ Mammalian inhibitior of rapamyocin Spry ........................................................................................................... Sprouty SH2 ............................................................................................... Src homology-2 SH3 ............................................................................................... Src homology-3 WT1 ............................................................................................... Wilm’s tumor-1 FGF ................................................................................ Fibroblasts growth factor NGF ....................................................................................... Nerve growth factor PTP1 B ................................................................ Protien tyrosine phosphatase-1 B GDNF ...................................................................................... Glial derived growth factor XV Introduction Cancer is a genetic disorder that results from the accumulation of genetic and/or epigenetic changes. Such changes uncouple cellular functions form their regulatory mechanisms, and lead to the emergence of a population of cells that has acquired the necessary characteristics to form a cancer. Although distinct types of cancers exhibit specific characteristics, there are several traits that are commonly observed in various types of cancers. These traits include uncontrolled proliferation, deregulated cell cycle, limitless replicative potential, evasion of apoptosis, inefficient DNA repair mechanisms, genetic instability, angiogenesis and invasion and metastasis [1]. As cancer results from Changes at the genomic level, a great effort has been placed to determine which genes play a role in cancer formation. In a broad sense, cancer-related genes are classified as oncogenes and tumor suppressor genes, which are altered through gain-of-function or loss-of-function genetic events, respectively [2-4]. These events include mutations, epigenetic regulation, chromosomal translocations and expressional changes [5]. The process of carcinogenesis progresses through distinct intermediate clonal populations of cels, which may have accumulated some of the characteristics of cancer cells, even though such populations of cells are not malignant. The isolation of such intermediate populations from human tumors in vivo has been successful for colorectal cancer [6], but has proven difficult other types of cancer. To overcome this problem, a number of systems have been generated in order to mimic the process of cancer formation. One such system is the MSU-1 lineage of human fibroblasts [7, 8]. This system consists of isogenic cell strains, which have acquired specific genetic changes in a sequential order, and display a progressive accumulation of traits related to cancer. This lineage originates form a normal fibroblast cell line, and culminates in a cell strain capable of forming tumors in athymic mice. Within these extremes, there are a series of isogenic cell strains with an intermediate status. These characteristics make this lineage an efficient model system to the study genetic elements that play a role at the various stages of carcinogenesis. The Ras GTPase binds to and hydrolyzes GTP to GDP. Ras functions as a critical molecular switch that regulates a number of cellular signaling pathways important for cellular proliferation, survival, and organization of the actin cytoskeleton [9, 10]. Binding to GTP induces a conformational change within Ras that results in Ras activation [11, 12]. Active Ras binds to, and activates a number of effector proteins, including Raf, PI3K and Ral. When GTP is dissociated to GDP, Ras assumes an inactive conformation. This results in the dissociation of effector proteins from Ras and the activation of effector-mediated pathways by Ras is attenuated. Ras, a proto-oncogene, is activated to its oncogenic form in approximately 30% of human tumors [13]. The oncogenic activation of Ras is the result of mutations in several codons, including codons 12 and 59. These P mutations abrogate the ability of Ras to hydrolyze GTP, stabilizing the active conformation of Res [14]. The role of Ras oncogenes in cancer formation is mediated by the same effectors that mediate the effect of Ras proto-oncogenes under normal cellular conditions (i.e. Raf, Pl3K, Ral etc.) [15-17]. The difference with oncogenic Ras, however, is that the activation of effector pathways by Ras is constitutive, resulting in unrestrained proliferation and in the inactivation of apoptotic programs. Sprouty (Spry) was identified in Drosophila melanogaster, as an inhibitor of receptor tyrosine kinase (RTK) signaling [18-20]. This function of Spry is important for normal development of several organs including the tracheal system and the eye. Mammalian cells express four Spry proteins, which, like the Drosophila homolog, retain the ability to suppress RTK signaling [18]. The expression of Spry is prominent in locations where fibroblast growth factor (FGF) and epidermal growth factor (EGF) are prominent. In mammals, Spry acts in a negative feedback fashion to repress signaling form these growth factors. This function of Spry is important in several developmental processes including the development of the kidney, the vestibular apparatus and the bone [21-23]. In Drosophila Spry is a general inhibitor of RTK signaling [19]. In mammalian cells, however, Spry proteins, particularly Spry2, sustain RTK signaling induced by EGF [24—26]. The cellular functions that are mediated by this ability of Spry remain uncharacterized. Spry proteins may play a role in cancer formation, as several recent studies found that the expression of Spry proteins is altered in some types of cancer. Spry1 and Spry2 are expressed at lower levels in breast and pancreatic tumors [27, 28]. Moreover, Spry2 suppresses tumor formation upon expression in breast cancer cells [29]. This ability is consistent with the function of Spry as an inhibitor of RTK signaling. Spry2 is also expressed at higher levels in melanomas [30, 31]. Although this finding correlates with the ability of Spry2 to sustain epidermal growth factor receptor Signaling, the role of Spry2 in the formation of these tumors remains unknown. The interest of our laboratory in the study of Spry was sparked by a gene expression analysis comparing the expression profiles of cells in the MSU 1 lineage. In particular, this study compared the expression profiles of MSU-1.0, an immortalized diploid human fibroblast cell strain, MSU-1.1, a cell strain derived from MSU-1.0 cells, and PH3MT, a tumor derived cell strain originating from the malignant transformation of MSU-1.1 cell with the HRasV’z-oncogene. Spry2 was identified as one of the genes that were increased in expression in MSU—1.1 and PH3MT cells, when compared to MSU-1.0 cells. This finding led to the hypothesis that Spry2 promotes tumor formation in HRas-transformed cells, the proof of which is the scope of this dissertation. Chapter I will provide a broad review of the literature in the field of cancer research, then proceed with more depth to the review of the function of Res, and finally address in detail the functions of Spry proteins, with an emphasis placed on Spry2. Chapter II will describe the role of Spry2 in the transformation of immortalized human fibroblasts by oncogenic HRas. This study found that Spry2 is necessary for the ability of HRas transformed fibroblasts to form tumors in athymic mice. Furthermore, this study demonstrated that HRas interacts with Spry2 and two Spry2-binding partners c-Cbl and CIN85. Chapter III will describe the role of Spry2 in the ability of HRas to induce survival pathways that desensitize human fibroblasts to DNA damage-induced apoptosis. This study found that Spry2 is necessary to protect Res-transformed cells from UV-induced apoptosis. In this context, Spry2 sustained the activation of enzymes involved in survival pathways, including phosphatidyl inositol-3 kinase (Pl3K), Akt, and Rac1, while maintaining a low level of the pro-apoptotic tumor suppressor p53. Appendix A will describe the gene expression study that compared the expression profiles of MSU-1.0, MSU-1.1 and PH3MT cells. In addition to the identification of Spry2 with a possible role in cancer formation, this study also found a number of other genes differentially expressed between malignant and pre-malignant cells. Appendix B will describe research in progress to determine the role of Spry2 in the ability of human patient-derived fibrosarcoma cell lines to form tumors in athymic mice. Two cell lines, HT1080 and VIP:FT, which contain oncogenic NRastg and wild type Ras respectively, express high levels of Spry2. This study found that Spry2 contributes to the malignant phenotype of VleFT cells, and, at a lesser extent, to the malignant phenotype of HT1080 cells, suggesting that Spry2 contributes to tumor formation in a context-specific fashion. References 10. 11. 12. 13. Hanahan, D. and Weinberg, RA, The hallmarks of cancer. Cell, 2000. 100(1): p. 57-70. Ponder, B.A., Cancer genetics. Nature, 2001. 411(6835): p. 336-341. Todd, R. and Wong, D.T., Oncogenes. Anticancer Res, 1999. 19(6A): p. 4729-4746. Macleod, K., Tumor suppressor genes. Curr Opin Genet Dev, 2000. 10(1): p. 81-93. Macaluso, M., Paggi, MG. and Giordano, A., Genetic and epigenetic alterations as hallmarks of the intricate road to cancer. Oncogene, 2003. 22(42): p. 6472-6478. Kinzler, K.W. and Vogelstein, 3., Lessons from hereditary colorectal cancer. Cell, 1996. 87(2): p. 159-170. McCormick, J.J. and Maher, V.M., Analysis of the multistep process of carcinogenesis using human fibroblasts. Risk Anal, 1994. 14(3): p. 257- 263. McCormick, J.J., Fry, D.G., Hurlin, P.J., Morgan, TL, Wilson, D.M. and Maher, V.M., Malignant transformation of human fibroblasts by oncogene transfection or carcinogen treatment. Prog Clin Biol Res, 1990. 3400: p. 195-205. Coleman, M.L., Marshall, OJ. and Olson, M.F., RAS and RHO GTPases in G1-phase cell-cycle regulation. Nat Rev Mol Cell Biol, 2004. 5(5): p. 355- 366. Downward, J., Targeting RAS signalling pathways in cancer therapy. Nat Rev Cancer, 2003. 3(1): p. 11-22. Colicelli, J., Human RAS superfamily proteins and related GTPases. Sci STKE, 2004. 2004(250): p. RE13. Milbum, M.V., Tong, L., deVos, A.M., Brunger, A., Yamaizumi, Z., Nishimura, S. and Kim, S.H., Molecular switch for signal transduction: structural differences between active and inactive forms of protooncogenic ras proteins. Science, 1990. 247(4945): p. 939-945. 803, J.L., ras oncogenes in human cancer: a review. Cancer Res, 1989. 49(17): p. 4682-4689. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. Bollag, G. and McCormick, F., Regulators and effectors of res proteins. Annu Rev Cell Biol, 1991. 7: p. 601-632. Repasky, G.A., Chenette, E.J. and Der, C.J., Renewing the conspiracy theory debate: does Raf function alone to mediate Ras oncogenesis? Trends Cell Biol, 2004. 14(11): p. 639-647. Shields, J.M., Pruitt, K., McFall, A., Shaub, A. and Der, C.J., Understanding Ras: 'it ain't over til it's over'. Trends Cell Biol, 2000. 10(4): p. 147-154. Campbell, S.L., Khosravi-Far, R., Rossman, K.L., Clark, G.J. and Der, C.J., Increasing complexity of Ras signaling. Oncogene, 1998. 17(11 Reviews): p. 1395-1413. Casci, T., Vinos, J. and Freeman, M., Sprouty, an intracellular inhibitor of Ras signaling. Cell, 1999. 96(5): p. 655-665. Reich, A., Sapir, A. and Shilo, B., Sprouty is a general inhibitor of receptor tyrosine kinase signaling. Development, 1999. 126(18): p. 4139-4147. Hacohen, N., Kramer, S., Sutherland, D., Hiromi, Y. and Krasnow, M.A., Sprouty encodes a novel antagonist of FGF signaling that patterns apical branching of the Drosophila airways. Cell, 1998. 92(2): p. 253-263. Kim, H.J. and Bar-Sagi, D., Modulation of signalling by Sprouty: a developing story. Nat Rev Mol Cell Biol, 2004. 5(6): p. 441-450. Guy, G.R., Wong, E.S., Yusoff, P., Chandramouli, 8., Lo, T.L., Lim, J. and Fong, C.W., Sprouty: how does the branch manager work? J Cell Sci, 2003. 116(Pt 15): p. 3061-3068. Christofori, G., Split personalities: the agonistic antagonist Sprouty. Nat Cell Biol, 2003. 5(5): p. 377-379. Egan, J.E., Hall, A.B., Yatsula, BA. and Bar-Sagi, D., The bimodal regulation of epidermal growth factor signaling by human Sprouty proteins. Proc Natl Acad Sci U S A, 2002. 99(9): p. 6041-6046. Rubin, C., Litvak, V., Medvedovsky, H., Zwang, Y., Lev, S. and Yarden, Y., Sprouty fine-tunes E GF signaling through interlinked positive and negative feedback loops. Curr Biol, 2003. 13(4): p. 297-307. Wong, E.S., Fong, C.W., Lim, J., Yusoff, P., Low, 8.0., Langdon, W.Y. and Guy, G.R., Sprouty2 attenuates epidermal growth factor receptor ubiquity/ation and endocytosis, and consequently enhances Ras/ERK signalling. Embo J, 2002. 21(18): p. 4796-4808. 27. 28. 29. 30. 31. Kwabi-Addo, B., Ozen, M. and Ittmann, M., The role of fibroblast growth factors and their receptors in prostate cancer. Endocr Relat Cancer, 2004. 11(4): p. 709-724. McKie, A.B., Douglas, D.A., Olijslagers, 8., Graham, J., Omar, M.M., Heer, R., Gnanapragasam, V.J., Robson, ON. and Leung, H.Y., Epigenetic inactivation of the human Sprouty2 (hSPRY2) homologue in prostate cancer. Oncogene, 2005. 24(13): p. 2166-2174. Lo, T.L., Yusoff, P., Fong, C.W., Guo, K., McCaw, B.J., Phillips, W.A., Yang, H., Wong, E.S., Leong, H.F., Zeng, Q., Putti, TC. and Guy, G.R., The res/mitogen-activated protein kinase pathway inhibitor and likely tumor suppressor proteins, Sprouty 1 and Sprouty 2 are deregulated in breast cancer. Cancer Res, 2004. 64(17): p. 6127-6136. Abe, M. and Naski, M.C., Regulation of Sprouty expression by PLCgamma and calcium-dependent signals. Biochem Biophys Res Commun, 2004. 323(3): p. 1040-1047. Tsavachidou, 0., Coleman, M.L., Athanasiadis, G., Li, S., Licht, J.D., Olson, M.F. and Weber, B.L., SPRY2 is an inhibitor of the ras/extracellular signal-regulated kinase pathway in melanocytes and melanoma cells with wild-type BRAF but not with the V599E mutant. Cancer Res, 2004. 64(16): p. 5556-5559. Chapter I: Review of Literature Cancer as a genetic disorder The American cancer society estimates that cancer caused approximately 556,000 deaths in 2003, accounting for nearly 23% of deaths in the United States [1]. Worldwide, 10 million new cases were reported in 2000, and 6 million people died from cancer [2]. Cancer is a complex pathologic disorder that is a result of multiple changes in normal physiology. Initially, change(s) within a cell result in unregulated cellular growth, leading to an abnormal accumulation of cells and the formation of a tumor at a particular location, which represents the primary tumor site or tumor origin. In some cases, tumors infiltrate the tissue surrounding the primary site and invade lymph and/or blood vessels, a process known as invasion. In addition, tumors that invade can implant to a secondary location, a process known as metastasis. Tumors that invade and metastasize are defined as malignant tumors, and constitute a cancer. In contrast tumors that are restricted to the site of origin and exhibit no evidence of invasion are defined as benign tumors. These are often precursors of malignant tumors, but benign tumors themselves do not constitute a cancer [3]. Cancer-related genes Cancers arise by the sequential acquisition of genetic and/or epigenetic changes, which ultimately modify the activity of proteins encoded by the affected genes [4]. IO Many of these changes confer upon a cell some selective advantage that allows the expansion of a clonal population. By reiteration of this pattern over a period of years, a single cell emerges that has acquired all the necessary changes to form a cancer [5, 6]. Based on epidemiological evidence on the frequency of cancer incidence, Renan predicted that a normal cell requires between four to seven Changes to form a cancer [7]. A great deal of effort has been made to determine which cellular genes are involved in cancer formation. Such genes can be classified into two groups: oncogenes, which are activated by gain-of—function genetic events and tumor suppressor genes, which inactivated by loss-of—function genetic events [8-10]. In an attempt to elucidate the origins of cancer, early studies found that injection of several RNA viruses in some animals resulted in cancer formation [8, 11]. These viruses were also found to induce cellular transformation in culture. Cellular transformation refers to the acquisition of some of the characteristics of tumor cells (e.g. morphological change or limitless replicative potential), without including malignant tumor formation. The transforming ability sUch viruses is mediated by single genetic elements (eg. v-Src and v-Myc), which are homologous to genes found in normal cells [8, 12, 13]. The cellular counterparts of these genes (e.g. c-Src and c-Myc) are important for maintaining normal cell growth and differentiation, and play a causal role in cancer when they are activated [14, 15]. ll Genes that lead to cancer formation upon their activation are defined as oncogenes, while the cellular genes form which they are derived are defined as proto-oncogenes. Oncogenes act in a dominant fashion, implying that the activation of a single genetic allele suffices for their activation. Oncogene activation, results from point mutations (e.g. Rasv’z in pancreatic carcinomas [16], [17]), amplified expression (e.g. Her2/Neu in breast cancers [18, 19]), and chromosomal translocations (e.g. c-Myc in Burkitts lymphoma and c—Abl in chronic myelogenous leukemia [20, 21]). Tumor suppressor genes also play an important role in cancer formation. The cancer in which the role of tumor suppressors became evident is Retinoblastoma. This condition is characterized by unilateral or bilateral retinal tumors that afflict young children [22, 23]. Knudson [24] hypothesized that Retinoblastoma resulted from inactivating mutations in a gene encoding a growth inhibitory protein. In light of the fact that some children develop unilateral lesions, whereas others develop bilateral lesions, Knudson postulated that when tumors were found in both eyes, this was a consequence of a single inactivating mutation, which was inherited, and a second mutation acquired independently as retinal cells divided to form the retina. The gene responsible for this condition, designated retinoblastoma (Rb), was the first tumor suppressor gene to be identified [25-27]. Normally, Rb serves to repress cellular proliferation by inhibiting cell cycle progression (discussed below), thus suppressing uncontrolled proliferation of cells. In Retinoblastoma, 12 this gene is inactivated by loss-of-function mutations, and therefore the ability of Rb to inhibit cell cycle progression is lost [28]. Since the discovery of Rb, a number of other tumor suppressor genes have been identified, which regulate diverse cellular functions and protect cells form malignant transformation [29]. Inactivation of a single allele of a tumor suppressor gene is not sufficient for tumor formation, as the other allele of the gene remains intact and can provide enough wild type protein to maintain a normal phenotype. Therefore, both copies of the gene must be inactivated in order for the function of the tumor suppressor gene to be lost. This is commonly known as the “two hit hypothesis” [30]. Tumor suppressor genes have been further classified as gatekeepers or as caretakers [31]. Vogelstein and colleagues discovered that APC, a tumor suppressor gene, is inactivated in a type of colon cancer known as Familial Adenomatus Coli [32, 33]. Biallelic inactivation of APC is the rate limiting step for the formation of these tumors. Genes with this property in cancer formation are designated gatekeepers. Another subset of colon cancers, known as Hereditary Nonpolyposis Colorectal Cancer, arises form mutations in DNA mismatch repair genes [34, 35]. Because these genes are important in repairing damaged DNA, their inactivation increases the mutation rate and chromosomal instability, which facilitate cancer formation [36]. Tumor suppressor genes that act in this fashion are designated as caretaker genes, i.e. genes whose inactivation accelerates malignant transformation. 13 Mutations are the most important cause for the alteration of normal cellular genes in the process of tumor formation. Nevertheless, epigenetic events, such as imprinting and hypermethylation have also proven to be important in inducing the necessary changes for cancer formation. Epigenetlc regulation refers to the control of gene expression through modifications of chromatin structure in the gene promoter region [37]. These modifications facilitate (e.g. acetylation), or impede (e.g. methylation) transcription from a particular gene promoter, resulting in alterations in gene expression. Disruption of epigenetic regulation can also contribute to cancer formation. For example, loss of imprinting facilitates the formation of Wilm's tumor [18] and hypermethylation of the p21 promoter results in the loss of expression of p21 in various tumors [38]. Characteristics of cancer cells The family of cancer consists of a large number of distinct types, each displaying some unique characteristics, particularly in regards to the cell of origin. Nevertheless, there are also characteristics that are commonly found in various cancer subtypes [39]. Such common traits include: (1) enhanced or uncontrolled proliferation, (2) aberrant cell cycle control, (3) limitless replication potential, (4) evasion of apoptosis, (5) inefficient DNA repair mechanisms, (6) genomic instability, (7) angiogenesis and (8) invasion and metastasis. I4 Uncontrolled proliferation Under normal conditions, a cell requires mitogenic growth signals for proliferation. Cancer cells, however, have acquired the ability of autonomous proliferation, i.e. they can replicate in the absence of exogenous growth factors. This autonomy is acquired as a result of changes that enable cancer cells to uncouple normal proliferation pathways from their regulatory mechanisms. With this in mind, it is important to describe the process responsible for the regulation of proliferation in normal cells, before describing how cancer cells acquire their proliferative autonomy. Of the several pathways that regulate cellular proliferation within a cell, the pathways induced by growth factors are the most important. Extracellular protein growth factors, such as epidermal growth factor (EGF) and platelet-derived growth factor (PDGF), bind to and activate transmembrane receptors, such as the EGF-receptor (EGFR) and the PDGF- receptor (PDGFR), respectively. These receptors belong to a wider family, known as receptor tyrosine kinases (RTK). Growth factors bind to these receptors and induced their dimerization. This process leads to autophosphorylation of the receptors on their cytoplasmic domains, and this leads to receptor activation [40]. The phosphorylation sites on RTKs, serve as docking sites for numerous effectors, including the adaptor protein Grb2 [41]. Grb2 contains a Src homology 2 (SH2) domain, which binds to phospho-tyrosine residues and guides the migration of Grb2 to the 15 phosphotyrosine residues on the cytoplasmic tail of the RTK [42]. In this fashion, growth factors stimulate the translocation of Grb2 from its cytosolic localization to the plasma membrane. Grb2 forms a complex with son of sevenless (SOS), leading to the translocation of SOS to the plasma membrane alongside Grb2 [43, 44]. At the plasma membrane SOS activates Ras GTPases, which are embedded in the plasma membrane by farnesylation [45]. In this fashion, the extracellular ligand-induced activation of receptor tyrosine kinases results in the activation of R33 GTPases. These GTPases catalyze the hydrolysis of GTP into GDP, and serve as a molecular switch for the activation of many intracellular pathways [46, 47]. In normal cells, Ras oscillates between an active, GTP-bound state and an inactive, GDP-bound state [48]. The Grb2:SOS-mediated activation of Ras in response to growth factor stimulation is a critical event in the regulation of proliferation. This is mainly because active Ras regulates the Raf/MEK/ERK pathway [49]. Activation of this pathway directly regulates the expression of many genes that are important for cell cycle progression and proliferation (e.g. cyclin D) [50-52]. One mechanism by which cancer cells hijack this regulatory pathway is the production of endogenous growth factors. For example, glioblastomas and sarcomas have been frequently found to produce PDGF and TGFa respectively [39, 53, 54]. Upon secretion, these growth factors act in an autocrine fashion to stimulate RTK signaling and proliferation of cancer cells in the absence of exogenous growth factors. 16 Cancer cells may also exhibit an increased level of RTK activity, because of mutations that result in constitutive activation of the receptor, or amplification of the receptor. For example, c-Kit receptor tyrosine kinase, is activated in gastrointestinal stromal tumors by a point mutation, which results in the constitutive dimerization of this receptor [55]. Growth factors activate RTKS by inducing their dimerization. Upon growth factor withdrawal, the receptors revert to a monomeric state. Thus, constitutively dimerized receptors are active even in the absence of growth factor stimulation, which results in uncontrolled signaling activity. Alternatively, and more commonly, growth factor receptors are overexpressed in cancer cells, as is the case in brain, breast and stomach tumors, which overexpress the tyrosine kinase receptor EGFR [56]. This also results in sustained activity from the receptor, and a sustained drive for cellular proliferation. Proliferative autonomy in cancer is most often acquired via the constitutive activation of proto-oncogenes, such as Ras in colon and pancreatic cancers [57] and Abl in chronic myelogenous leukemia [58]. Activation of the Ras proto- oncogene is a result of point mutations in codons 12, 13, 59 and 61 [59-61]. These mutations diminish the GTPase activity of Ras, which results in higher levels of active, GTP-bound Res, and lead to sustained proliferative signaling from Ras. Abl is a non receptor tyrosine kinase that also functions in signaling pathways that regulate cell growth [62]. In chronic myelogenous leukemia Abl is 17 activated through reciprocal translocation between chromosomes 9 and 22, resulting in the formation of the “Philadelphia chromosome”, which contains a fusion of Abl with BCR [63]. Although the exact mechanism by which BCR-Abl affects cancer formation remains under investigation, one mechanism appears to be an increase in the kinase activity of Abl [64, 65]. This increase results in part because an inhibitory region on the N-terminus of Abl is lost upon fusion of Abl with BCR [66]. The modulation of transcription factor activity also contributes to the higher rate of proliferation observed in most cancer cells. The transcription factors that play a role in cancer can be generalized into three groups [67, 68], including steroid re(lealzbtors, such as estrogen and androgen receptors, nuclear transcription faCto rs, such as c-Myc and c-Jun, and cytoplasmic factors, such as NFKB and STATS. Expressional amplification, or constitutive activation of these transcription fac>‘|Z<)rs not only contributes to the unrestrained growth of cancer cells, but is also important for the acquisition of other characteristics of cancer cells, such as eVasion of apoptosis and deregulation of the cell cycle. The c-Myc transcription factor is of particular importance, given that c-Myc proto-oncogene is involved in the formation of many cancers, including those of the prostate, breast and skin [69‘71]. c-Myc encodes a basic helix-loop-helix-zipper transcription factor, which Specifically binds to E-box sequences in the DNA, upon heterodimerization with 'ts binding partner MAX [12]. As part of this complex, c-Myc activates a diverse grc) Up of genes, including cell cycle regulators (e.g. cyclin D2 and CDK2) and 18 translation initiation factors (e.g. elF2 and eIF4). Functionally, c-Myc drives cell ular proliferation, inhibits differentiation, and induces apoptosis [72-74]. Abe rrant cell cycle regulation In somatic cells, the normal cell cycle consists of a resting stage, Go, in which the cel I is not dividing, and four stages, G1, S, G2 and M, which are present in divi ding cells. S is the stage during which DNA synthesis take places, while M is the stage of mitosis, were cell division occurs. G1 is a gap phase between M and S . whereas G2 is a gap phase between S and M. These gaps allow the cell to pre Dare for the ensuing stage, as well as to repair DNA damage [75]. The cell cycle is regulated by an array of proteins, including cyclins (A, B and D), CYCI in-dependent kinases (CDK1-4) and CDK inhibitors (CKI, e.g. p21 and p27) [76 . 77]. In broad terms, a Specific cyclin binds to a specific CDK, and leads to CDK activation. This CDK, in turn, phosphorylates and regulates enzymes that are responsible for progression through the different stages of the cells cycle. In this fashion, cyclin/CDK complexes promote cell cycle progression. lmportantly, the progression through a particular phase of the cell cycle is regulated by a Specific cyclin/CDK complex. CDK inhibitors bind to cyclin/CDK complexes and In“ 5 bit their activity. Consequently, these inhibitors arrest progression through the Ce l l cycle [76, 78. 79]- 19 The transition through G1 to S is of particular importance, because it is the part of the cell cycle where many intracellular signaling pathways exert their control on cell cycle progression. This transition is mainly regulated by the E2F family of transcription factors. E2F regulates a number of genes that are involved in me i ntaining S phase, such as DNA polymerases and cyclin E [80]. In quiescent cel ls, E2F is repressed by the Rb tumor suppressor, when the latter is in a hYIDOphosphorylated form. Growth stimulation by mitogenic signals activates the Re s/MAPK cascade, resulting in transcriptional activation of cyclin D. Cyclin D inte racts with CDK4, and the cyclinD/CDK4 complex hyperphosphorylates Rb, promoting the release of E2F from RB. Active E2F transcriptionally activates Qe hes that are necessary for progression from G1 to S [81, 82]. Aberrations in the mechanisms that regulate this transition through the cell cycle OCCUr frequently in human cancers, and include inactivation of Rb (e.g. Retinoblastoma), as well as overexpression of cyclin 01, which is commonly Observed in breast, lung and colon cancers [83. 84]. The S phase is maintained predominantly by the cyclin ElCDK2 complex [85, 86]. The cyclin E/CDK2 complex also activates E2F [87]. Activation of E2F, at this De l‘iod of the cell cycle, induces cyclin A expression. Cyclin A, is responsible for the transition from G2 to M, which leads the cells into the stage of division [88]. c3)’<:Iin A also regulates the first half of mitosis, whereas cyclin B, in complex with CDK‘I, is responsible for regulation of the second half of mitosis [3]. 20 CDK inhibitors include members of the ClP/KIP family (p21, p27 and p57) and the INK 4a locus (p16'NK4a and p14ARF). p21 and p27 inhibit the function of cyclin/CDK complexes, and therefore, they prevent progression through 8 and G2 phases [89] - p16lNK4al inhibits the cyclin D/CDK interaction, thereby inactivating E2F, whereas p14ARF prevents cell cycle progression by inhibiting the ubiquitinylation and degradation of p53 by MDM2 [90]. In addition to the regulatory programs described above, there exists another level Of Control, termed cell cycle checkpoint control, which is responsible for ensuring the integrity of the genome as the cell progresses through the different phases of the cell cycle [91]. Regulation at this level involves sensor, transducer, and effector proteins, and is divided into the G1/S checkpoint, the S-phase chefizzkpoint, and the GZ/M checkpoint. Such checkpoints prevent cells with Ciat‘haged DNA from entering S-phase (G1/S), prevent the start of replication in the case of genotoxic insults (S-phase), or prevent cells with aberrant DNA replication from entering mitosis (G2lM) [92—94]. In the absence, or in the ihaCtivation of these checkpoints, damaged DNA is replicated, leading to a higher freq uency of mutations, thus increasing the risk of cancer formation. The sensors of DNA damage, is. the factors that initiate checkpoint control 'nVO Ive mainly ATM and ATR. The former is instantaneously activated by DNA es i(Dr‘rs, whereas the latter is activated by stalling of the replication fork. ATM and 21 ATR activate signaling transducers, such as CHK proteins, which then recruit and activate effectors including p53 [91]. The tumor suppressor gene p53 plays an important role in cell cycle regulation and cancer formation. p53 is found to be mutated in approximately 50% of human ca hcers. p53 encodes for a transcription factor that acts as a homotetramer to reg ulate cell cycle arrest, apoptosis and DNA repair [95-97]. Under physiological CO hditions, the levels of p53 are maintained at a low level by MDM2, which is an E3 ubiquitin ligase enzyme [98]. MDM2 ubiquitinates p53 and promotes its d e9 radation by the proteasome [99]. The transcriptional activity of p53 is activated in response to cellular stresses that ind Lice DNA damage. Upon DNA damage, the ubiquitination of p53 by MDM2 is abOlished, and p53 translocates to the nucleus, where it induces the transcription 0f genes such as p21, Bax, and Gadd5 [100]. This results in the activation of the 92 ’1 CKl, which, as described above, arrests the cell cycle by inhibiting c:yczlin/CDK complexes [101]. The consequence of this arrest is to allow more time for the repair of DNA damage. If the damage is not repaired, p53 may induce apoptosis via the transcriptional activation of BAX, a pro-apoptotic factor that in h i bits the anti-apoptotic factor Bcl-2 [102]. 22 Limitless replication potential In 1 961, Hayflick suggested that human cells in culture have a limited lifespan [1 O 3]. Fibroblasts, for example, can replicate for 60-80 cell doublings, after which they stop the progression through the cell cycle and enter a metabolically active state termed senescence. Senescent fibroblasts are enlarged and exhibit a flat mo rphology, and can persist for many years in this state. The cell cycle arrest d u ring senescence is dependent on the activities of the p53 and Rb tumor 3” Dpressors [39]. Upon inhibition of these pathways, cells can propagate beyond 11"‘Ieir normal life span in culture. Nevertheless, within a limited number of DO pulation doublings, such cells enter a state termed crisis that is characterized by massive apoptosis. In some instances, cells survive this state and emerge with a limitless replicative potential [104]. Cell replication in culture is limited by the shortening of telomeric ends of Ch r0 mosomes. Telomeres consist of tandem repeats of the sequence TTAGGG [1 0 5]. These repeats are limited in number, and they are consecutively shortened d Uri ng the replication of chromosome [106]. In this way, the length of telomeres Se Wes as a molecular timer that counts down with every cell replication. When the telomeric repeats fall below a critical number, terminal parts of chromosomal DNA are lost during DNA synthesis. This process triggers the cellular programs that bring on senescence, although the exact mechanism is not fully understood [1 07]. If the cell continues to divide the progressive loss of chromosomal ends I eads to genomic instability, which is responsible for the induction of crisis. 23 Cancer cells in culture typically exhibit limitless replicative potential, suggesting that this characteristic is acquired during tumor progression [108]. In some malignant cells, proliferation is paralleled by widespread apoptosis, suggesting that the limited lifespan of somatic cells needs to be surpassed for cancers to fo m [39]. Telomere maintenance is present in the majority of cancer cells [109]. This is mainly attributed to the up regulation of telomerase expression, an enzyme responsible for the synthesis of telomeres [110]. Not surprisingly, expression of telomerase is sufficient to bypass senescence and crisis, conferring “mitless replicative potential to cells [111]. It should be noted that there is a telOmerase-independent mechanism for telomere maintenance. This mechanism ir‘\/c>lves recombination and is referred to as alternative lengthening of telomeres [1 1 2]. E\'i.=ls.ion of apoptosis ADthosis is regulated by two distinct pathways. One is mediated by death re(T-eptors in response to extracellular signals and is termed the ‘extrinsic pathway’. The other is mediated by the mitochondria in response to internal cues, i"(3| uding DNA damage, and is referred to as the ‘intrinsic pathway’ [113, 114]. In the extrinsic pathway, death receptors, such as CD95 and TRAIL-R1, are acti\rated through their interaction with various ligands that belong to the TNF f . . a "h Ily of secreted proteins [115]. This interaction induces receptor clustering, and 24 leads to the recruitment of caspase-8 and caspase-10 to the receptor, via the adaptor protein FADD [116]. Caspases are cysteine proteases that are synthesized as inactive zymogens. Most caspases are activated by proteolytic cleavage, usually induced by another active caspase [117]. After their recruitment to the death receptor, both caspase-8 and caspase-10 are cleaved into their active forms [118]. c-FLIP negatively regulates the extrinsic apoptotic pathway by inhibiting caspase-8 activation by the death receptor-FADD complex [119]. Activation of caspase-8 leads to the step wise activation of a cascade of Ca 8 pases culminating in the activation of caspase-3. In the intrinsic pathway, DNA damage results in the secretion of cytochrome c fro m the mitochondria [120]. Cytochrome c release from mitochondria is regulated by pro-apoptotic (BAX, BID, BAD) and anti-apoptotic (BCL2, BCL-XL) factors, and the net outcome results from an imbalance between the two types of factors. In the cytosol, cytochrome c interacts with Apaf-1 and caspase-9, resulting in the for l'1r1ation of the “apoptosome” [121]. The ‘apoptosome’ activates caspase-3, a prOczess that is antagonized by inhibitor of apoptosis proteins (IAP), which in turns are inactivated by Smac/DIABLO [122]. Therefore the intrinsic and extrinsic p"athways converge at the activation of caspase-3, an important effector caspase that targets critical cellular enzymes resulting in apoptosis. The cellular phenotype associated with apoptosis results form the proteolysis of v - . . e""<>us cellular substrates. Proteolysrs of substrates such as nuclear lamrns 25 results in nuclear condensation, proteolysis of DNase inhibitor ICAD, activates an endonuclease that fragmentates DNA, whereas proteolysis of cytoskeletal proteins results in cell fragmentation [114]. Apo ptosis acts as a potent control to prevent malignant transformation. As noted above, cancers arise as a result of genetic alterations (e.g. oncogene activation), and subsequent clonal expansion of the cells. Activation of apoptotic pathways aCts to prevent clonal expansion, and thus limit the chance for full blown neOplasm formation. Interestingly, the activation of some oncogenes (e.g. OVe rexpression of c-Myc and E2F), in itself sensitizes cells to apoptosis, thus prevanting malignant transformation [73]. Apoptosis can also be initiated in response to extensive DNA damage. This is mediated by the function of p53, which “senses” DNA damage and activates the intl"i nsic apoptotic pathway through the induction of BAX [102]. The treatment of cancers by radiation or chemotherapy also relies on the ihd LlCtiOl'I of apoptosis. Inactivation of pro-apoptotic pathways in cancer cells compromises the efficacy of such treatments. For example, malignant rT‘lelanomas that have lost expression of APAF1 become resistant to ch e motherapy [123]. 26 The importance of apoptosis in slowing cancer progression is also apparent in that almost all of the factors involved in the regulation of apoptosis are affected in human cancers. For example, the anti-apoptotic factors BCL2 and BCL-XL are overexpressed in myeloid leukemia and in acute lymphoblastic leukemia, respectively. Also, the pro-apoptotic factor BAX has been found to be down- regulated in colon cancer. Furthermore, additional apoptotic regulators, such as FLIP, soluble death ligands, IAPs, p53, Pl3K, AKT and PTEN are reported to be deregulated in tumors [114]. Inefficient DNA repair mechanisms DNA damage is especially important for cancer formation, because it causes mutations in replicating cells. As indicated above, mutations may result in the loss of function of tumor suppressor genes, as well as in the activation of oncogenes. These events are sufficient to initiate and maintain the process of malignant transformation [124]. DNA damage is as a result of exogenous or endogenous chemicals that form adducts with DNA bases directly, or indirectly, through their metabolites. Furthermore, DNA damage is a result of physical agents such as UV and ionizing irradiation. In order to maintain the information encoded in the DNA unaltered, cells have developed mechanisms to repair such damage, which results in a low frequency of mutation. The repertoire of the cell’s DNA repair machinery includes base excision repair, nucleotide excision repair and mismatch repair, as well as 27 homologous recombination and nonhomologous and joining [125]. What is more, there exists a “damage tolerance" pathway, in which specialized polymerases (Y- Family polymerases) bypass lesions that stall the major replication polymerase (POI 5) [126]. Furthermore, as cells progress through the cell cycle, the cells must pass through several checkpoints, which are regulated by distinct cellular enzymes, such as p53 and ATM. These enzymes are responsible for arresting cell cycle progression, and inducing programmed cell death, thus preventing cells with damaged DNA from propagating. Deficiencies in the DNA-repair machinery enhance mutation frequency and are detrimental for normal cell function. Paradigmatically, mismatch repair genes are inactivated in human hereditary colorectal cancer, a common malignancy of the colon [31, 34, 35]. Mismatch repair genes (e.g. MSH2) act as caretaker genes and their inactivation enhances the mutation frequency, which can lead to cancer formation. In this setting, oncogenes and tumor suppressor genes are more likely to be activated and suppressed, respectively. Xeroderma Pigmentosum (XP) is a syndrome that renders patients susceptible to skin cancer [127]. The cells of these patients are deficient in nucleotide excision repair, resulting in a higher frequency of sunlight-induced mutations in these cells [128, 129]. A subset of XP patients has a normal nucleotide excision repair 28 mechanism, yet the frequency of UV-induced mutations in cells derived from these patients is also high. This condition, designated Xeroderma Pigmentosum Variant is characterized by defects in polymerase eta, an error-free specialized polymerase involved in translesion synthesis [130]. Genetic instability Genetic instability, a feature of many cancer cells, refers to the consistent failure to transmit an accurate copy of a complete genome from one cell to its two daughter cells. This can be subdivided into microsatellite and chromosomal instability [131]. Microsatellite instability is a result of mutations or inactivation of DNA-mismatch-repair genes, such as MSH2. Chromosomal instability can be further subdivided into instability in chromosome structure and instability in chromosome number. Instability in chromosome structure involves deletions, inversions, translocations and insertions of small sequences of DNA. In tumor cells, this type of instability often results from the inactivation of DNA-damage checkpoint genes, such as ATM and p53, as well as from deficiencies in genes involved in double strand break repair like DNA-PK. Instability in chromosome number is a product of abnormal centrosome duplication with multipolar mitoses, and arises form deficiencies in BRCA1 and spindle checkpoint genes (MAD1). Chromosomal instability leads to an enhanced rate of loss of heterozygosity, which is an important mechanism of inactivating tumor suppressor genes [132]. 29 Angiogenesis Cells within a tissue require the delivery of oxygen and nutrients for their growth and proliferation. To achieve this, cells are generally located within 100-200 pm from blood vessels [3]. With this in mind, organismal growth requires that new blood vessels are formed, so that the new cells are constantly being perfused. The process by which new vessels are formed is defined as angiogenesis. This process is strictly regulated by a plethora of factors, which either stimulate or inhibit the formation of new blood vessels. In normal tissue, the tendency of the pro-angiogenic factor to stimulate angiogenesis is balanced by that of the anti- angiogenic factors [133, 134]. In solid tumors, growth is frequently limited by the hypoxic condition at the center of the tumor, and by the sparsity of blood vessels to deliver oxygen and nutrients to the tumor site. Therefore, a tumor cannot grow beyond a critical size (estimated at 200 um) unless the tumor receives sufficient blood perfusion to support its own metabolic needs. Tumors can remain in this stage for a period of months to years [3]. When the balance between pro- and anti- angiogenic factors is shifted to promote the formation of new vessels, a process referred to as “angiogenic switch”, tumor growth resumes, as oxygen and nutrients are being delivered to the tumor site [135, 136]. Typically, angiogenesis is initiated and carried out by endothelial cells lining up existing blood vessels [137, 138]. Endothelial cells express receptors for pro- 30 angiogenic factors that are secreted in the interstitial fluid, or that are incorporated in the extracellular matrix. Upon binding to their ligands, these receptors become active and they turn on endothelial angiogenic programs [139]. In addition, upon their stimulation to neovascularize, endothelial cells secrete matrix proteases that degrade the extracellular matrix and allow the endothelial cells to proliferate towards the source of their stimulus [140]. Furthermore, angiogenesis also involves the arrangement of endothelial cells into tubular structures, their canalization and their intussusceptions into existing vessels [138, 141, 142]. Vascular endothelium growth factor (VEGF) is the best characterized regulator of angiogenesis. VEGF is secreted by the tumor cells or by the tumor stroma, and is critically important for the ‘de novo’ formation of angiogenesis [143-145]. The secretion of VEGF is regulated by the transcription factor HIF1, as a response to hypoxic stress in the tumor microenvironment [146]. When oxygen is abundant in the microenvironment, non-heme iron dependent oxygeneases hydroxylate HIF1 on specific residues, mediating an interaction between HIF1 and the vHL E3 ubiquitin ligase. This process results in the ubiquitinylation and subsequent degradation of HIF1. Upon hypoxic conditions, the oxygenases responsible for HIF hydroxylation remain in an inactive state, thereby failing to induce the post- translational down regulation of HIF1 [147]. This failure results in a transcriptional activation of HIF1 inducible genes, which include the angiogenic stimulator VEGF. The VEGF receptor, belongs to the class of RTKS, and transduces the 31 VEGF angiogenic Signal through a number of effector proteins including FAK, PI3K and Ras-dependent pathways [148, 149]. Invasion and metastasis Malignant tumors are characterized by their ability to invade surrounding tissues and as a result to metastasize into other locations in the body. The metastatic process involves detachment of the tumor from its primary site, degradation of its extracellular matrix (ECM), invasion into the blood vessel, and transplantation to a secondary site [3]. The initial stages in the metastatic process are regulated by cell adhesion molecules and integrins [150, 151]. E-cadherin, a transmembrane glycoprotein, is an important factor for epithelial cell adhesion. Some carcinomas express reduced levels of E-cadherin, resulting in loose attachment between the epithelial cells, thereby enhancing their potential for metastasis. Alternatively, carcinomas with normal levels of E-cadherin, contain inactivated catenin, which is an intracellular effector of E-cadherin [152]. To infiltrate through the ECM, tumor cells first bind to the components of the ECM via transmembrane proteins of the integrin family and laminin family [151, 153]. These proteins, which are expressed in normal cells as well, are frequently amplified in cancer cells. In particular, cancer cells express integrins that are not specific for the type of tissue were the cancer originated from. This enables 32 invading cells to attach to new locations giving rise to new foci of tumor growth [151]. Once bound to ECM, the tumor cell secretes specific proteases, such as matrix metalloproteases (MMP2 and 9), which degrade ECM to facilitate the infiltration of tumor cells [154, 155]. In the process of invading adjacent tissue, cancer cells may invade into blood vessels. Once in the blood vessel, the tumor cells evade the immune system by homotypic adhesion (i.e. aggregation of tumor cells with each other) or heterotypic adhesions (aggregation between tumor and blood cells) [156]. The site of extravasation is dependent in part on the anatomical location of the primary tumor. Extravasation and transplantation of the tumor cells involves the attachment of tumor cells into a new tissue type, a process that is facilitated by laminin receptors and integrins, which bind to ligands embedded in the ECM of the metastatic site [157]. In addition, chemokines and their receptors also play a role in determining the target site for metastasis, particularly in breast cancer cells [158]. 33 MSU-1 lineage as model system to study malignant transformation To study the process of malignant transformation, McCormick and Maher developed the MSU-1 lineage of human fibroblasts (Fig. 1), as a model system that mimics the process by which normal cells become malignant [159-163]. This lineage originates from the transfection of normal foreskin-derived human fibroblasts with a vector encoding the v-Myc oncogene and a neomycin resistance marker [164]. As neomycin resistant clones were being propagated in culture, they underwent senescence, and the majority of the clones succumbed to crisis. Nevertheless, it became apparent that several clones had survived this process. Because cells that escape senescence and crisis spontaneously acquire an immortal life span [104, 165], the v-Myc-expressing clones were propagated in culture for many cell doublings to determine if exhibited extended lifespan. It was found that the cells from these clones were indeed immortal, and they were designated MSU-1.0. Experiments conducted later showed that MSU-1.0 cells express telomerase, a gene known to immortalize to cells (McCormick, unpublished data). Myc confers immortality to cells by inducing the expression of telomerase [166-169], yet in our system it is unlikely that Myc alone is responsible for the immortalization of MSU- 1.0 cells. This is because all but one of the Myc expressing clones succumbed to crisis, just like the mock transfected clones. What is more, MSU-1.0 cells also express elevated levels of the transcription factor Sp1, when compared to their 34 Figure 1. MSU 1 lineage of human fibroblasts. The MSU 1 lineage consists of isogenic cell strains which have been derived from the same normal cell line, and have progressively increasing malignant characteristics. Some characteristics of each cell line, as well as the known genetic modification(s) that are responsible for its formation are indicated. 35 F 0.59". 36 oocmuconou:_ mo> Bi :2th oz 02 oz .308 526.0 88863:. nmEcocmcmc. wm> 02 oz 02 635E555; mo> mo> we> mm; 02 oz fitoEE. moEOmoEoEo mo> mo> mm; 02 oz 02 59:22 2me2 2.82 .2802 mm; wo> wo> 2035 Qaaum wm> wm> wo> mo> >=mEomoEoEo 20.: ._5a 5 not: .5306? precursor cells. Sp1 has been shown to cooperate with Myc in the induction of telomerase expression [170, 171], suggesting that Myc and Sp1 are most likely responsible for telomerase expression in MSU-1.0 cells (McCormick, unpublished data). Apart from being immortal, MSU-1.0 cells exhibit the usual characteristics of normal fibroblasts, i.e., they are diploid, chromosomally stable, require normal levels of growth factors for proliferation, do not form foci or colonies in agar, and they fail to form tumors upon injection in athymic mice. Furthermore, expression of oncogenes such as, and v-sis in these cells, as well as treatment with benzo-A- pyrene-diol-epoxide (BPDE) and focus selection alone, or treatment with BPDE, V12 focus selection and subsequent expression of the HRas oncogene, did not malignantly transformed these cells. A cell in the MSU-1.0 population spontaneously undenivent two chromosomal translocations, giving rise to a variant clonal population, designated MSU-1.1. MSU-1.1 cells are chromosomally stable, and nearly diploid, i.e., they consist of 45 chromosomes, including two characteristic chromosome markers, M1 and M2 along with a monosomy of chromosomes 11, 12 and 15 [164]. M1 resulted from the translocation of chromosome 11 (11p15-)qter) to chromosome 1 at p11, whereas M2 resulted from translocation of chromosome 12 (12qter->12q11) to chromosome 15 (15p11-)15qter). Fingerprint analysis, and analysis of the v-myc integration site by southern blotting, showed that MSU-1.0 and MSU-1.1 cells 37 were both derived from LG1 cells, and MSU-1.1 cells must have been derived form MSU-1.0 cells, respectively. MSU-1.1 cells exhibit an immortal life span in culture, but they do not form foci, colonies in agar, or tumors upon injection in athymic mice. Nevertheless, unlike their precursor cells, MSU-1.1 cells exhibit partial growth factor independence, and can be malignantly transformed by the expression of various oncogenes, or by carcinogen treatment followed by focus selection. Expression of oncogenes, such as HRasV’Z, NRastg and v-sis at expression levels similar to the expression of the respective proto-oncogenes, results in transformation of MSU-1.1 cells (Fig. 2A) [172-174]. MSU-1.1 cells expressing these oncogenes form foci and colonies in agarose at higher rates that the parental MSU-1.1 cells. These cells, however, are not malignant, i.e., they do not form tumors when injected in athymic mice. However, overexpression of the HRaswz, NRaswz, v-KRas and KRasV’2 oncogenes, results in malignant transformation of these cells, i.e., MSU-1.1 cells expressing these oncogenes form malignant tumors in athymic mice [173-177]. These results suggest that more than one genetic change is required for the malignant transformation of MSU-1.1 cells. Consistently, sequential expression of two oncogenes, each of which is expressed at levels similar to its endogenous level, followed by clonal selection after each expression, results in the malignant transformation of MSU- 1.1 cells. In this way, the coexpression of either HRasWZ or NRasQ59 with v-fes, or 38 Figure 2. MSU 1 lineage as a tool for the study of malignant transformation. The MSU 1 lineage has been used to study the role of many genes in malignant transformation. (A) Unlike their precursor cells, MSU-1.1 cells are malignantly transformed by the overexpression of a single oncogene, or consecutive expression of two cooperative oncogenes. Expression of Ras oncogenes at a high level is sufficient for malignant transformation of MSU-1.1 cells. Instead, expression of Ras or sis (PDGF) oncogenes at a low level is insufficient to malignantly transform MSU-1.1 cells. Nevertheless, consecutive expression of a cooperative oncogene in cells expressing Ras or sis oncogenes results in malignant transformation of MSU-1.1 cells, suggesting that MSU-1.1 cells are at least two genetic changes short of being malignant. Interestingly, application of the same genetic changes that malignantly transform MSU-1.1 cells in MSU-1.0 cells fails to malignantly transform these cells. (B) Cells derived from tumors resulting form the malignant transformation of MSU-1.1 cells (MSU-1.1 derivative) have also been used to study genes that are involved in cancer. When the expression of the indicated genes is altered, the malignant MSU-1 derivatives exhibit a reduction in their ability to form tumors in athymic mice. 39 N 0.59... 2 05082650 went: QTESQE :o_mmo.axo._o>o o_com_._oE3-coz UmZBULoEJH 2:82ch Toms. co=a_:a$-cioo FQM. . morhms m NVOBU FUNK NSMMKX $3me «EMMKI «Swat: :ofiwEango 2:82.60 Swim—ESE accuse 9:22.000 co_mme.axm Mai. mOQl wmknx. ~F>wmml .< $0312 40 coexpression of v-sis with v-fes, in MSU-1.1 cells, results in their malignant transformation (Lin et al., 1995; Lin et al., 1994). The viral oncogene v-sis encodes for platelet derived growth factor (PDGF), while v-fes encodes an active form of the soluble tyrosine kinase c-fes. The cooperativity between v-sis, v-fes and Ras oncogenes in the malignant transformation of MSU-1.1 cells is evident in the fact that these enzymes are involved in receptor tyrosine signaling. Growth factors (such as PDGF) induce the activation of RTK signaling, which results in the activation of intracellular effectors such as Res and Src [178, 179]. Soluble tyrosine kinases (such as c-fes) partially mediate the functions of activated RTKS. In addition, soluble tyrosine kinases, especially Src and c-fes, can also activate Ras [180-182]. As indicated above, MSU-1.1 can be malignantly transformed by carcinogen treatment followed by focus selection. Treatment of MSU-1.1 cells with a number of carcinogens, including BPDE, ethyl-nitrosourea (ENU) and gamma-irradiation, results in the malignant transformation of these cells [183-185]. In these cases, the resulting malignant cells exhibit inactivation of the p53 tumor suppressor pathway and are chromosomally unstable (McCormick, unpublished data). The cells derived from the tumors formed by the malignantly transformed MSU- 1.1 cells have been studied to determine genes that are important for this process (Fig. 28). Expression of dominant negative forms of Rho GTPases Rac1 and Cdc42 in HRasVIz-transformed MSU-1.1 cells results in a decrease in the ability 41 of HRas-transformed cells to form tumors in athymic mice, indicating that Rac1 and Cdc42 are necessary for Res-transformation in human cells (Appledom and VIA, and in McCormick, unpublished data). Also, down-regulation of Sp1 in HRas gamma irradiation-transformed MSU-1.1 cells, abrogates tumor-forming ability, suggesting an oncogenic function for Sp1 in these contexts[186]. Furthermore, c- Met and HGF, are both necessary for the malignant phenotype of gamma irradiation-transformed MSU-1.1 cells [187]. The activation of the HGF/c—Met pathway results in the induction of Sp1 expression [187], suggesting that Sp1 mediates the effect of HGF and c-Met in malignant transformation. Finally, expression of Fibulin-1D, an extracellular matrix protein, in BPDE-transformed MSU-1.1 cells abrogates the ability of these cells to form tumors in athymic mice, indicating the importance of extracellular matrix in the transformation of immortalized human fibroblasts [188]. The fact that the cells of the MSU-1 lineage are isogenic, i.e. they are derived from the same precursor, has allowed the study of genetic changes that occur early in the transformation process and predispose the cells to malignant transformation. Understanding the reason behind the ability of MSU-1.1 cells to be malignantly transformed by certain oncogenes, while their precursor MSU-1.0 cells fail to do so, is important in fully understanding the mechanism by which malignant transformation occurs. 42 The marker chromosomes that are present in MSU-1.1 cells may play a role in predisposing MSU-1.1 cells to malignant transformation by the oncogenes indicated above. Insertion of chromosome 15 in MSU-1.1 cells prevented their transforrnability by overexpression of HRasV12 , and also the expression of the same chromosome in cells derived from tumors formed by HRasVIz-transformed MSU-1.1 cells, abrogated their tumor forming potential. Chromosomes 1, 11 and 12 did not have an effect, when studied similarly, suggesting that alteration of chromosome 15 is important for predisposing cells to HRas-transformation (Kaplan et al. manuscript in preparation). 43 Ras GTPase Ras genes, including HRas, KRas and NRas, are prototypical examples of oncogenes [189]. Approximately 30% of human tumors, particularly pancreatic carcinomas, adenocarcinomas of the lung, myeloblastomas and colorectal carcinomas contain activated Ras genes [57, 190, 191]. HRas and KRas were discovered thirty years ago, when mouse leukemia viruses were injected in rats giving rise to soft tissue sarcomas [192, 193]. These viruses encode for oncogenic forms of HRas and KRas cellular genes. The same oncogenic forms of Ras genes, encoded by the transforming viruses, were also found to be expressed in human tumors. [194, 195] The expression of Ras oncogenes is a causal factor for the ability of leukemia viruses to transform cells, as well as for the malignant phenotype of human tumors. The other member of the Ras family, NRas, was discovered in a human tumor of neural origin and plays an oncogenic role in tumor formation as well [194-197]. The oncogenic form of Ras differs by a single point mutation, when compared to the respective Ras proto-oncogene [198]. Although only one amino acid is affected, this mutation is critical for the function of Ras. This is because the protein that is encoded by the oncogenic form of Ras has the ability to evade normal regulation and exhibits increased activity [198-200]. This increased activity mediates a number of cellular functions, including enhanced proliferation and survival, which are necessary for cancer formation. 44 Catalytic function The members of the Ras family of genes encode for small G proteins. Unlike the classic heterotrimeric G proteins, small G proteins lack the regulatory B and 7 subunits, and consist only of the catalytic a subunit [201]. Small G proteins, including Ras family members, are GTPases, i.e., they can bind to, and hydrolyze GTP to GDP [46, 47, 202]. This activity is mediated by six structural components of Ras that are conserved among the family members and can be categorized into “G box” sequences[189, 203]. The G1 box, [aaaanxxxGK(S/T); a: L/lN/M, x: any amino acid] mediates purine nucleotide binding, whereas the G3 box, [blbbexGl; b: hydrophobic, I: hydrophilic] binds to Mg”. The G4 box, [bbbb(N/T)(K/Q)xD], forms hydrogen bonds with the guanine ring, thus conferring specificity over binding to adenine. The G5 box, [bbE(A/C/S/T)SA(K/L)], interacts indirectly with guanine nucleotides and is less conserved among the family members. The catalytic activity of Ras is cyclic over time, la, the hydrolysis of a GTP molecule to GDP, is followed by the dissociation of GDP and the association of a new GTP molecule, leading to a new cycle of hydrolysis [204]. Consequently, Ras exists in two states (conformations): in one state, Ras is bound to GTP and in the other, Ras is bound to GDP [205]. When Ras is bound to GTP, it interacts with its effector proteins, a process that results in the activation of these effectors, which then mediate the cellular functions of Ras [203, 206]. Structurally, this is mediated 45 by the 62 box, [Y DPTIEDSY], which adopts an orientation that has a high affinity for Ras effectors, when Ras is in a GTP-bound state. This is due to conformational changes on two loop regions on Ras, referred to as switch 1 and switch 2 [205]. These changes are induced when GTP binds Ras. Upon hydrolysis of GTP to GDP, the G2 box adopts an inactive conformation, and the activation of Ras effectors is attenuated. The fact that Ras oscillates between an active, GTP-bound conformation and an inactive, GDP-bound conformation, enables this protein to function as a molecular switch to regulate cellular activity. Regulation of Ras activity Because Ras functions as a molecular switch, it is important that Ras is strictly regulated in order to maintain normal cellular function (Fig. 3). The regulation of Ras is based on two kinetic parameters that characterize Res-mediated catalysis: (i) the dissociation of GDP from Ras is rate limiting, and (ii) the intrinsic GTPase activity is low [207]. The dissociation of GDP is catalyzed by guanine nucleotide exchange factors (GEFS) [204]. This process enables Ras to bind another GTP molecule, which enhances its ability to activate its effectors. Furthermore, guanine nucleotide dissociation inhibitors (GDls) bind Specifically to GDP-bound Res and inhibit the dissociation of GDP, thus prolonging the inactive state[45, 208-210]. The hydrolysis of GTP to GDP is catalyzed by GTPase activating proteins (GAPS). These enhance the rate of GTP-hydrolysis by Ras, leading to the inactivation of Ras [211-213]. The point mutations present in the oncogenic 46 Figure 3. Res GTPase as a molecular switch. The ability of Ras GTPases to hydrolyze GTP to GDP is associated with two distinct Ras conformations: an active conformation, when Ras is GTP-bound, and an inactive conformation, when R35 is GDP-bound. In normal cells, Ras oscillates between the inactive to the active conformation and back. This is regulated by guanine nucleotide exchange factors, including 808 and RasGEF, which catalyze the release of GDP, and enhance Ras activation. Instead, GTPase activating proteins, such as p120GAP enhance the intrinsic ability of Ras to hydrolyze GTP to GDP, 3 process that leads to Ras inactivation. 47 93* Integrins p120GAP Figure 3 forms of Ras, e.g., at codons 12 or 61, render Ras resistant to GAP activity, thereby prolonging the active state of Res, and leading to tumor formation [191, 214]. GEFS and activation of Ras GEFS, including SOS, RasGEF and RasGRP, mediate the activation of small GTPases. These GEFS are activated in a signal-specific manner, and each displays specificity for the type of Rae-family GTPase that it activates. SOS is the best studied Ras specific GEF. Structurally, SOS comprises of a Cdc25 domain, a Dbl homology (DH), and a pleckstrin homology (PH) domain [215]. The Cdc25 domain catalyzes the dissociation of GDP from Res, and is conserved in Ras-specific GEFS. In addition, the C-terminus of SOS contains a proline-rich region, which binds to the SH3 domain of the Grb2 adaptor protein. This adaptor also contains an SH2 domain that is recruited to phospho-tyrosine residues of proteins localized at the plasma membrane (e.g. transmembrane receptors, or other adaptors). As Grb2 is bound to SOS, the recruitment of Grb2 to the plasma membrane facilitates the translocation of SOS to the plasma membrane, where it activates Res [216]. This mechanism of SOS activation is responsible for mediating Ras activation by several upstream receptors including RTKS, Integrins and G-protein coupled receptors (GPCR). Upon binding to their ligands, tyrosine kinase receptors, like 49 the epidermal growth factor receptor (EGFR), dimerize and autophosphorylate their own intracellular domains on tyrosine residues. These residues act as docking sites for the SH2 domain of Grb2 and recruit the Grb2:SOS complex to the plasma membrane [40, 217]. Other RTKS, notably the insulin growth factor receptor (IGFR), recruit and phosphorylate adaptors, such as Shc, which then become docking sites for Grb2:SOS [218]. In a similar fashion, Shc acts as a docking platform for Grb2:SOS during the activation of Ras by integrins or by heterotrimeric G-proteins. Upon binding to the extracellular matrix, integrins dimerize and activate focal adhesion kinase (FAK), which in turn phosphorylates Shc, thus potentiating the recruitment of Grb2:SOS to Shc [151]. Heterotrimeric G proteins can activate Ras through their By subunit, in a process involving the activation of Src, and Shp-mediated recruitment of Grb2:SOS to the plasma membrane [219, 220]. In addition to the interaction with Grb2, the activity of SOS is also regulated by phosphorylation. Extracellular-signal related kinase (ERK) and its effector p90RSK, both phosphorylate and inactivate SOS in a negative feedback fashion [221]. Furthermore, phosphatidyl inositols (i.e. PIP2) interact with the PH domain of SOS to inhibit its activity [222]. The RasGEF family consists of the same structural domains (i.e., DH, PH and Cd025) as the SOS family [223]. Their signal specificity, however, differs from that of the SOS family. RasGEF are mediators of calcium induced-Res activation 50 [204]. These GEFS also contain an IQ motif, which binds to the calcium- calmodulin complex leading to the activation of RasGEF [224]. The RasGRP family is not only regulated by calcium, but its function also depends on diaglycerol [223]. GAPS and inactivation of Ras The ability of GAPS to inactivate Ras lies on their ability to interact with Ras and enhance the rate of GTP hydrolysis by Ras [213]This activity is crucial for the cycling of Ras between “on” and “off” states. The study of GTPase activating 0GAP and Neurofibromatosis-1 (NF1)) has been focused proteins (including p12 mainly on p12OGAP, which consists of a C-teminal catalytic domain, two N-terminal SH2 domains flanking and SH3 domain, a PH domain, as well as a lipid binding domain[212]. The exact mechanism by which the activity of p120GAP is regulated is still under investigation, but it seem that maximal activity is dependent not only on the catalytic domain, but also on the SH2 and SH3 domains of p1206AP [225]. Posttranslational regulation Although synthesized as a cytosolic protein, Ras is subjected to posttranslational processing. This results in the dynamic interaction of Ras with cellular membranes, including the plasma membrane, endosomes, and other intracellular membranes [226-228]. This process is directed by the CAAX motif at the carboxy- terminal of Ras. Initially, the cysteine residue of this motif is famesylated by a SI famesyl transferase [229]. Subsequently, an endopeptidase cleaves off the AAX tripeptide [230], followed by methylation of the or-carboxy group of the famesylated cysteine [231]. At this point, the processing of Ras proteins can follow two distinct pathways, in an isoform specific fashion. HRas and NRaS become palmitoylated on a cysteine residue on the amino terminal Side of the famesylated residue, and are trafficked through the Golgi system to reach the plasma membrane. In contrast, KRas lacks this cysteine residue and reaches the plasma membrane in Golgi-independent mechanism[232]. Ras Effectors Ras plays a central role in intracellular signal transduction, not only because signaling pathways induced by various transmembrane receptors converge at the level of Ras GTPase, but also, because, Ras regulates a number of diverging signaling pathways that regulate many cellular functions. In this fashion, Ras acts as a gearbox that guides cellular function in a particular environment. This function of Ras relies on its ability to activate a number of effectors, including Raf, Pl3K, Tiam1, RalGDS, Rin1, AF-6, RASSF1 and others. Based on their catalytic functions such effectors can be classified into distinct groups consisting of: 1) serine/threonine kinases (Raf), 2) phosphoinositide-3-kinases (Pl3K), 3) GEFS (RalGDS, Tiam1 and Rin1), 4) lipases (PLCy), 5) adaptors (AF-6 and RASSF1) and 6) RasGAPs (p120GAP and NF1) [233]. A Ras effector is defined as a protein that interacts specifically with the effector domain of Ras, when Ras is in its GTP-bound form [233]. Effector proteins are 52 typically characterized by a domain that interacts with Ras. Three such domains have been identified to date; the RBD domain of Raf and Tiam1, the RBD domain found in the isoforms of the p110 subunit of Pl(3)K family and the Ras association domain, such as the one present in RalGDS, Rin1, AF6 and others [233, 234]. Even though these Res-binding regions are different in their amino acid sequences, their three dimensional conformations are similar. This is not surprising, since all these regions interact with the effector-binding domain of Ras (G2 box), and therefore, they need to have a structural conformation that fits that of the G2 box of Ras. This review will focus on effectors of Ras, which have critical and established functions in RaS signaling, as well as on a few additional effectors that have been studied to a lesser extent, but may have a significant role in Ras function (Fig. 4). Raf The Raf serine/threonine kinase is the best characterized effector of Ras. From a structural perspective, Raf consists of three conserved regions CR1, CR2 and CR3, which contains the catalytic domain [235]. Although the exact mechanism by which Raf is activated is still under investigation, Ras interacts with two domains, i.e., RBD and CR0 in the CR1 region of Rat [236, 237]. This association results in the recruitment of Raf to the plasma membrane, and represents the initial step in the activation of Rat [238]. The activation of Raf is also potentiated 53 Figure 4. Res effector pathways. The conformational change that is associated with Ras activation enables Ras to interact with a number of effector proteins, leading to the activation of many intracellular pathways and the regulation of diverse cellular functions. 54 by additional factors such as 14-3-3, phosphatidyl serine, Hsp90 and serine/threonine kinases [239]. Through the activation of Rat, Ras can activate the mitogen-activated protein kinase (MAPK)-cascade. This cascade functions as a phosphorelay system, in which kinases phosphorylate other kinases resulting in their activation in a sequential fashion. This system is organized in MAPK kinase kinases (MAPKKK), which activate MAPK kinases (MAPKK), and these in turn activate MAPKS [240]. Active Raf acts as a MAPKKK to activate MEK proteins by phosphorylation on two serine residues (8217 and 8221 in MEK1). MEK proteins are dual-specificity MAPKKS that phosphorylate ERK1 and ERK2 on threonine and tyrosine residues (T202 and Y204 in ERK1), resulting in ERK activation [235, 241]. ERK proteins, which are serine/threonine MAPKS, translocate to the nucleus when activated. In the nucleus, ERKS phosphorylate and activate transcription factors (e.g. Elk1) [242]. ERKS also activate cytoplasmic substrates, including kinases MSK1 and p90RSk. Upon ERK phsohorylation, MSK1 and p90RSk translocate to the nucleus where they activate other transcription factors (e.g. Fos, SRF), histones (e.g. H3) and transcriptional regulators (e.g. CREB) [242]. In this fashion, Ras, through the Raf/MEK/ERK pathway regulates gene expression at the transcription factor level. 56 PI3K Pl3Ks are lipid kinases that phosphorylate phosphoinositides on the 3’ position of the inositol ring [243, 244]. P|3K consists of two subunits; a regulatory p85 subunit and a catalytic p110 subunit. The regulatory subunit consists of two SH2 domains which interact with phosphorylated tyrosine residues on activated growth factor receptors [245, 246]. Since the region between the two SH2 domains is tightly bound to the p110 catalytic subunit, the association of p85 with activated growth factor receptors recruits the catalytic subunit to the plasma membrane, were it can phosphorylate phosphoinositides. Furthermore, the p110 catalytic subunit of PI3K interacts with Ras, when Ras is GTP-bound [247]. This interaction is synergistic with the function of the p85 subunit to yield optimal activation of PI3K in response to growth factors. Activated PI3K converts Pl(4,5)P2 into Pl(3,4,5)P3, an important secondary messenger that mediates its effects through its binding to plextrin homology (PH) and FYVE domains [248, 249]. Therefore, proteins containing the PH domain, including Akt, PDK, and Tiam1 are critical mediators of PI3K signaling. Akt, a serine/threonine kinase, is the best studied effector of PI3K signaling. Through its PH domain, Akt translocates to the plasma membrane and binds to PI(3,4,5)P3. This results in a conformational change within Akt, which exposes two main phosphorylation sites (T308 and S473) [250]. Subsequently, PDK phosphorylates Akt, leading to the stabilization of its active conformation, resulting in enhanced signaling activity by Akt. 57 The ability of Ras to promote cellular survival is mediated in part by the activation of the Pl3K/Akt pathway. Akt, in particular, is important because it exerts a direct effect on many pathways that regulate survival [251]. One of the substrates of Akt, BAD, is a pro-apoptotic protein that binds to Bcl-2 and Bcl-X and inhibits their anti-apoptotic potential. Phosphorylation of BAD on Ser136 by Akt results in loss of inhibition of Bcl-2 and Bcl-X [252]. IKK, a serine/threonine kinase, is activated upon phosphorylation by Akt. Subsequently, IKK phosphorylates the NFKB inhibitor IKB, which is then ubiquitinated and targeted for degradation by the proteasome pathway. This results in the activation of the NFKB transcription factor, which translocates to the nucleus and transactivates a number of pro-survival genes[253]. Another important Akt substrate is the Forkhead family of transcription factors (FoxO), which regulate the expression of apoptosis inducing factors, such as Fas, TRAIL and TRADD [254]. Akt phosphorylation inhibits the function of FoxO transcription factors, thus enhancing the survival potential of the cell. MDM2 is an ubiquitin ligase that targets p53 for proteosomal degradation. Akt phosphorylates MDM2 on two serine residues, which sustains its ubiquitin ligase activity. This decreases the ability of p53 to induce cell cycle arrest and apoptosis [255, 256]. 58 Finally, GSK3 is primarily involved in regulating the conversion of glucose to glycogen, although it has additional functions, including a role in B-catenin signaling, as well as in cell cycle regulation[257]. GSK3 phosphorylation by Akt results in the inhibition of GSK3 function[258]. Tiam1 Like Ras, Rho GTPases are molecular switches that regulate a number of cellular functions [259]. The Rho family, which includes Rac1, Cdc42 and Rho, plays an important role in mediating the functions of active Ras, particularly in the regulation of the actin cytoskeleton and cellular migration. Nevertheless, the mechanism by which Ras recruits this family has not been fully elucidated. One possible mechanism involves PI3K and the production of the secondary messenger PIP3. As noted above, PI3K interacts with proteins containing PH domains and recruits them to the plasma membrane. Rho GRES have of PH domains, and thus can be recruited to the plasma membrane by PIP3, a fact that enhances the activation of Rho GTPases. Alternatively, RhoGAPs can be regulated by Ral and p120GAP (see below). An important step in determining the mechanism of Rho activation by Ras was made by the discovery that Tiam1, a GEF for R301, is a direct effector of Res [260]. Tiam1 contains a Raf-like RBD domain through which it interacts with Ras, when the latter is in a GTP-bound state. In addition, Tiam1 consists of a catalytic 59 DH/PH domain, characteristic of RhoGEFs, as well as a N-terminal PH domain, which can bind to PIP3. The presence of a PH domain on Tiam1 indicates that, in addition to Ras, Tiam1 can be activated by PI3K [261]. Tiam1 requires membrane localization and threonine phosphorylation for maximal function [262, 263], suggesting that these are necessary regulatory steps in Tiam1 activation. In addition, Tiam1 is regulated by an autoinhibitory region and a PEST domain, both of which are located on its N-terminus. The PEST domain targets Tiam1 for degradation, a process that results in the cleavage of the autoinhibitory region and increased activity of Tiam1 [264]. Although the interaction of Ras with Tiam1 enhances the translocation of Tiam1 to the plasma membrane, whether Ras has an effect on the other aspects of Tiam1 activation remains unknown. RalGEFs The RalGEF family members, including RalGEF, le and Rgl, catalyze guanine nucleotide exchange for Ral GTPases. RalGEFs interact directly with Ras, in a GTP-dependent manner [265]. This interaction is dependent on the RA region of RalGEFs, which is very Similar in structure to the RBD of Raf1 [266]. The interaction with Ras causes the redistribution of RalGEFs to the plasma membrane, where they can activate Ral GTPase [267]. Ral has a number of targets that appear to mediate its activity. Ral interacts with PLD1 and Arf, suggesting that Ral plays a role in vesicular tansport [268]. RaIBP1 is a putative effector of Ral that interacts with tyrosine phosphorylated proteins 60 (Pob1 and Reps1), which complex with EGFR and are homologous to proteins that regulate receptor endocytosis [269]. Furthermore, RalBP1 has GAP activity, which is Specific for Rac1 and Cdc42 GTPases, an ability that implicates Ral in the regulation of Rho GTPases [270]. MEKK1 MEKK1 interacts with GTP-bound Ras through its C-teminal region, [271]. MEKK1 is a serine/threonine kinase MAPKKK, which regulates a Similar MAPK- cascade to that regulated by Raf. MEKK1 activates MKKs, which then activate p38 or JNK [272]. Active p38 and JNK translocate to the nucleus where they activate transcription factors and regulate gene expression. It is important to note that although these MAPKS function primarily in the nucleus, they also regulate cytoplasmic enzymes. Rin1 Ras interacts with Rin1 in a GTP-dependent manner, through its effector-binding domain. Rin1 is a GEF for Rab5 GTPase, which is involved in receptor-mediated endocytosis [273]. This process may be facilitated by the ability of Rin1 to interact with tyrosine phosphorylated receptors through its SH2 domain [274]. In addition to Rab5, Rin1 interacts with Abl tyrosine kinase to enhance its catalytic activity [275]. The interaction with Ras enhances the ability of Rin1 to activate the Rab5 GTPase, as well as the ability of Rin1 to activate Abl [273]. 61 AF-6 AF-6 is another protein that interacts with Ras in its GTP-bound state [276]. Ras interacts with the RA region of AF-6, an interaction that interferes with the binding of Raf to Ras. Although the function of AF-6 in the cell remains unknown, this protein interacts with the Rap1 GTPase, profilin, and with Eph receptors [277- 279] and may be involved in signal transduction at cell-cell junctions [279]. RASSF RASSF proteins (including NORE1 and RASSF1), interact with GTP-bound Ras through their RA region [280]. Interestingly, the expression of these proteins is lost in several tumors [281, 282]. In addition, in vitro and in vivo studies propose that these protein have anti-proliferative functions, thereby acting to suppress Ras-transforrnation. lmportantly, RASSF1 and NORE heterodimerize, and form a complex with Mst1, a serine/threonine kinase that enhances caspase-3 activation. Ras associates with this complex, and induces apoptosis by means of caspase-3 activation [283]. Cellular functions mediated by Ras Regulation of cell cycle and proliferation One of the most important cellular effects induced by active Ras is the progression through cell cycle, leading to cellular replication. Expression of a 62 dominant negative form of Ras inhibits G1-)S cell cycle progression, whereas progression is enhanced upon expression of oncogenic Res [284]. G1->S progression is regulated mainly by the activity of the Rb protein, which is normally in a hypophosphorylated form. In this form Rb sequesters and inhibits the function of E2F transcription factor. Rb phosphorylation is regulated by cyclinzCDK complexes. The cyclin D:CDK4 complex, in particular, hyperphosphorylates Rb, a process that attenuates the inhibition of E2F by Rb. This results in the transcriptional activation a number of E2F-dependent genes that are important for cellular proliferation. The primary regulatory function of Ras at this stage of the cell cycle is to inactivate the suppression of E2F by Rb [285]. The activation of Raf/MEK/ERK pathway by Ras, induces the expression of cyclin D1, which as described leads to the activation of E2F [286]. The activation of additional Res-effector mediated pathways, including those mediated by PI3K and Rac1/Cdc42 are also important in mediating the inactivation of Rb by Res [287, 288]. PI3K regulates the expression of cyclin D1 through GSK3. GSK3 phosphorylates cyclin D1 and this initiates the degradation of cyclin D1 through the proteasomal pathway. Pl3K, by means of an Akt-dependent function, inhibits the activity of GSK3, thus sustaining the levels and activity of cyclin D [289]. Rac1 is also important in inducing cyclin D accumulation, in a process that is dependent on the activation of NFKB [290]. 63 In addition to the regulation of Rb, Ras can also regulate cell cycle inhibitors, such as p21 and p27, to promote G198 transition. Through the Raf/MEK/ERK pathway, Ras can decrease p27 levels both transcriptionally and through proteosomal degradation [50, 291]. What is more, activation of PI3K by Ras results in the inactivation of FoxO-dependent transcription, including the transcription of p27 [292]. Also, the activation of Rho induces the formation of cyclin E:CDK2 complex, which phosphorylates p27, and provides a signal for the ubiquitination and degradation of p27 [293]. Ras may also inhibit cell cycle progression depending on the context of its activation. Expression of Ras oncogene in primary fibroblasts leads to INK4a and p53-dependent cellular senescence, whereas inactivation of these proteins results in evasion of the senescent phenotype [294]. Downstream of Ras, even Raf can have different outcomes on cell fate. Whereas moderate levels of Rat activation result in enhanced proliferation, through the cyclin D pathway, high levels of Raf activation result in p21-mediated cell cycle arrest [295]. Regulation of protein synthesis Ras can regulate protein synthesis through the Pl3K/Akt pathway. In addition to its other functions, Akt also exerts control over the Rheb/mTOR pathway by phosphorylating and inactivating TSC1/2 proteins [296]. TSC1/2 proteins act as GAPS for Rheb GTPase, which activates the mammalian target of rapamycin protein (mTOR) [297], and this in turn activates p7086K and 4EBP1 [298, 299]. 64 p7086K phosphorylates the S6 ribosomal protein. This results in the translation of enzymes that make up the protein synthesis apparatus. Normally, 4EBP1 inhibits translation by sequestering the translation elongation factor elF4E. mTOR phosphorylates 4EBP1 and induces the release of elF4E, facilitating the initiation of translation [300]. Regulation of apoptosis In addition to its role in cell proliferation and cell cycle progression, Ras can also regulate cellular programs that control apoptosis. Several of the effectors that are activated by Ras are important in mediating this function of Res, and their outcomes, with respect to apoptosis, depend on the particular effector that is activated. For example, activation of PI3K by Ras activates survival pathways, whereas activation of RASSF1 leads to activation of cell death pathways. In some cases the activation of the same effector may lead to distinct apoptotic fates within the same cells, as is the case with the activation of Raf. The activation of survival pathways by the Ras effector PI3K is mediated by the activation of Akt. As indicated above, the ability of Akt to prevent programmed cell death is in turn mediated by a number of Akt effector substrates. Akt can inhibit BAD, thus preventing Bcl-2 and Bcl-X induced cytochrome c release from mitochondria [301]. In addition, Akt can regulate apoptosis by regulating transcription factors as NFKB and FoxO, which regulate cellular fates via the transactivation of anti- and pro- apoptotic genes, respectively. Akt can intercept 65 the p53 apoptotic pathway at the level of MDM2, thus protecting the cell from DNA-damage induced apoptosis [302]. What is more, Ras can promote survival via Rac1-dependent activation of NFI 15:1155 Translocation Domain' Raf-Bindin Re ion Spry1 GECTAPR‘TLFSCLACIIRQCLCSAESI‘I‘vv‘EYGTCi-ICIVR GIT” CSN Spry2 KECTYPRPLPSDUICDPZQCLCSRQNVIDYGTCVCC FPGLE' HCSN 591‘94 ifECASE‘ RTLPSCU'JCNQECLCSRC ITL‘J tI'I’GTCtICL'v'QCI r ESHCTN Spry1; V PCT RE LPDCULCNQI‘ CLCSHESLLDYGTCLCCV “iiE‘GL' CST I; ’ll‘l I1; lllll: .::lIII:N 16:16 Spry1 DDECDSYSDIIPCSCSQSHCCflLCIIGAIvISLFLPCLLCYPPRiCGCLI-(I. Spry2 DDE-DNCADIIPCSCSQSHCC [1’.9314IIG’JItSLFLPCLIJCYLPAi-(CCLKI. Spry4 EDDEGSCADHi-"(I‘EICSR'C IICCAT ISFIICA LS‘.’VLPCLLCYLPATGCVKL Spry3 DDE— DIICADEPCSCGPSSCF'sz. ’AllSLISLFLPCLCCYLE‘THGCLHL :l III-HI :- lr Ir ;~;-n~1 ‘1 II; nr3;1278 'héll' FMum5 72 Additional regulatory elements, which have been studied to a lesser extent than the N—terminal phosphorylation site of Spry, reside within the conserved cys-rich region. The region encompassed by amino acids 178-315 on Spry2 has been defined as the Spry translocation domain[341] and regulates the localization of Spry to plasma membrane upon growth factor stimulation. Spry proteins form homo- and hetero-dimmers with other isoforms. The structural element important for this function is located in a region spanning position 209- 238 on mSpry1 [340]. Spry proteins interact with Raf kinase through a Raf binding domain spanning residues 209-240 on Spry2. Surprisingly, the region important in Raf binding spans the same residues as does the region important for Spry dimerization. Recently a novel growth factor-specific tyrosine phosphorylation site was discovered on the C-terminus (Y227) of Spry2 [342]. Phosphorylation at this site appears to regulate if Spry inhibits or potentates RTK signaling. Finally, Spry2 contains three PxxPR motifs on positions 64, 72 and 309, which act as biding sites for the SH3 domains of CIN85 [343]. 73 Sprouty expression In Drosophila, Spry is expressed during tracheal [324] and eye development [327], as well as in midline glia [344], in wing veins and in ovarian follicle cells [344, 345]. These developmental processes are under the control of FGF or EGF [328], and loss-of-function mutations in Spry mimic the loss of expression of these growth factors in these tissues. In vertebrate embryos, such as the mouse and chick embryo, Spry protein expression is induced by FGF, and not surprisingly, it is more prominent in the locations where FGF signaling predominates [346]. These include the primitive streak, the forebrain and the hindbrain. It is important to note, however, that the expression of Spry genes in these tissues occurs in an isoform-specific fashion. For example, Spry2 and Spry4 are expressed in the primitive streak, whereas Spry1 is not, and Spry1 and Spry2 are expressed in the midbrain region, whereas Spry4 is not [333]. In the mouse, Spry genes are also expressed during organogenesis of the cochlea and semicircular canals, the teeth, the lungs, the digestive track, and the kidneys [347]. In the semicircular canal and the teeth, Spry1 and Spry2 are expressed in the epithelium, whereas Spry4 is expressed in mesenchymal or neuronal tissue. In the lung, Spry1, Spry2 and Spry4 are all expressed in epithelial tissue. In the kidney, Spry1 is expressed in the ureteric bud, whereas 74 Spry2 and Spry4 are expressed in the ureteric bud, mesenchyme and glomerulus. In the adult tissue, Spry1 is expressed in the heart, lung and kidney [348], Spry2 is expressed in the brain, heart, lung and kidney [349] and Spry4 is expressed in liver, skeletal muscle, heart, lung, kidney, spleen, plaCenta and small intestine [350]. Consistent with the observations described above, the expression of Spry in cultured fibroblasts or endothelial cells is induced by growth factor stimulation [351]. At a molecular level, this is in part dependent on ERK activation, as selective inhibitors of MEK abrogates FGF-induced Spry2 and Spry4 expression [351]. Alternatively, FGF may induce the expression of Spry1 and Spry2, in an ERK-independent fashion, through the activation of PLCy and calcium-dependent signaling, as demonstrated in a study were calcium chelation and PLCy inhibition abrogated the induction of Spry by FGF in a mouse chondogenic cell line [352]. At the transcriptional regulation level, the expression of Spry genes, particularly the expression of Spry1, is regulated by the WT1 transcription factor. WT1 directly associates with, and activates the Spry1 promoter. Also, expression of wild type, but not of catalytically inactive mutant WT1, induces the expression of Spry1 in osteosarcoma cells [348]. Furthermore, analysis of the promoter region of Spry2 revealed the presence of several cis-activng elements for AP2, CREB, 75 ETS and Sp1, leading to the hypothesis that the expression of Spry2 may be regulated by these transcription factors/regulators [353]. Sprouty localization Initial studies on the localization of Spry in Drosophila m. showed that Spry is associated with the inner leaf of the plasma membrane [327]. In mammalian cells, however, the attempt to determine the exact localization of Spry proteins has lead to several findings. In human embryonic kidney cells, Spry2 localizes to membrane ruffles upon stimulation with EGF [341, 354]. In the absence of stimulation, Spry proteins attain a diffusely cytoplasmic localization, while Spry2 co-localizes with microtubules [341]. Other studies have found that in Chinese hamster ovary cells, Spry2 is associated with vesicular structures resembling endosomes upon EGF-stimulation. These same structures become loaded with EGFR, under the same stimulation [338, 339]. It may be that Spry2 localizes in both of these structures, in a cell-type specific matter, given that in mouse fibroblasts, Spry2 translocates both to vesicular structures, as well as to the plasma membrane, upon EGF stimulation [355]. In endothelial cells, Spry1 and Spry2 are located in the perinuclear region and in vesicular structures, under serum starvation conditions. Upon growth factor stimulation, they translocate to the lamelipodia at the leading edge of the plasma membrane [356]. In this system, Spry proteins directly interact and co-localize with calveolin-1, but do not localize in lipid rafts [356]. 76 The translocation of Spry protein to the plasma membrane is necessary for their activity, because it facilitates the phosphorylation and subsequent activation of the proteins [340]. One possible mechanism employed by Spry proteins in order to attain their membrane localization is the use of a unique translocation domain (SpryTD) located in the C-terminal region [341]. Deletion of the region spanning residues 178-221 in Spry2 abolishes the EGF-induced translocation of Spry2 to the plasma membrane [341]. The translocation of SpryTD to the plasma membrane is downstream of active Rac1. Rac1N17 (a dominant negative mutant) inhibits the translocation of SpryTD [341, 354]. A later study found that the SpryTD specifically interacts with Ple [354]. This was demonstrated by the co- localization of SpryTD with the PH domain of PLCS, which interacts with Ple at the plasma membrane, as well as by the direct association of SpryTD with Ple- bound lipid vesicles. The importance of the association between SpryTD and PIP2 was evaluated with a Spry2 mutant (Spry2R2520), which fails to interact with Ple. This mutant Spry2 is incapable of repressing MAPK activation in response to FGF stimulation, unlike wild type Spry2, which localizes to PIP2 and inhibits MAPK activation [354]. Alternatively, Spry proteins are targeted to the plasma membrane through palmitoylation. Evidence for this modification comes from a study were radiolabeled palmitate was incorporated in Spry, which was ectopically expressed in endothelial cells [356]. Nevertheless, the palmitoylating enzyme, the position 77 on Spry, and whether it takes place in additional cell types, are questions that remain to be answered. Regulation of Sprouty Spry proteins function in a negative feedback fashion to repress RTK-dependent MAPK activation induced by FGF, PDGF, VEGF and HGF/Met. Surprisingly, in EGF-induced signaling, Spry is involved in a positive feedback loop, sustaining EGFR and MAPK activity. The involvement of Spry proteins in both negative and feedback loops, in a signal signal-specific fashion, suggest an important role for Spry in the regulation of RTK signaling. With this in mind, it is important to consider several mechanisms that regulate the function of Spry. Phosphorylation of the N-terminus of Sprouty The activity of Spry proteins is mainly regulated by the phosphorylation of a conserved tyrosine residue on the N-terminus of Spry (Y53 and Y55 in the case of Spry1 and Spry2, respectively) [338-340]. This phosphorylation has been observed in various cellular contexts, including fibroblasts and endothelial cells of both human and murine origin. The phosphorylation of Spry is induced by stimulation of the cells with growth factors such as EGF, FGF and PDGF [334, 336, 337, 357]. Spry isoforms exhibit different propensities for phosphorylation in response to different growth factors [355]. Spry1 is more likely to become phosphorylated upon stimulation of fibroblasts by FGF and PDGF whereas Spry2 78 is more likely to be phosphorylated in response to EGF and FGF. Furthermore, the kinetics of Spry phosphorylation may differ in a growth factor-specific manner. For example, Spry2 exhibits a relatively transient tyrosine phosphorylation in response to EGF stimulation, whereas in response to FGF signaling, the tyrosine phosphorylation of Spry2 exhibits a sustained profile [355]. The phosphorylation of Spry is necessary for the ability of Spry to regulate MAPK activation in response to growth factor stimulation. Spry mutants that are incapable of becoming phosphorylated on this conserved tyrosine residue (e.g. 2”“), fall to repress ERK activation in response to FGF, contrary to the Spry ability of wild type Spry proteins [340]. Given that these mutants can still translocate to the plasma membrane, it has been proposed that they have a dominant negative function [340, 358]. Unlike the wild type protein, Spry2Y55A fails to sustain EGF signaling as well, suggesting that phosphorylation of Y55 is also important for the ability of Spry2 to sustain EGFR signaling [338]. These findings suggest that the ability of Spry2 to repress RTK signaling, as well as the ability of Spry2 to sustain EGFR signaling, rely, in part, on the same molecular mechanism, i.e. the phosphorylation of Y55. The phosphorylation of Spry on Y53/55 (Spry1/2) has several implications, which attest dependence of Spry2 function on phosphorylation. One of the first studies to identify the phosphorylation of Spry1 and Spry2, demonstrated that in mouse 79 myoblast cells this phsophotyrosine residue serves as a docking site for the SH2 domain of Grb2. Grb2 mutants with a defective SH2-domain fail to interact with Spry1 or Spry2 [340]. In addition, this interaction is abrogated by mutations in the amino acids adjacent to the phosphotyrosine residue (T56l and E570 in Spry2), indicating that these residues are also necessary for the interaction between Spry2 and Grb2. The tyrosine phosphorylation of Spry2 on residue 55 is important for the interaction between Spry2 and the E3 ubiquitin ligase c-Cbl [359, 360]. Normally, c-Cbl ubiquitinates EGFR and targets this receptor for degradation. c-Cbl contains an atypical SH2 domain through which it interacts with phosphotyrosine residues on its substrates, such as EGFR and Zap70 [361]. Upon growth factor stimulation, Spry2 becomes phosphorylated on Y55, a site that then acts as a docking site for c-Cbl’s SH2 domain [338, 339]. In this fashion, Spry2 competes with EGFR for binding to c-Cbl, a process that prevents the degradation of EGFR. The region flanking Y55 on Spry2 is highly conserved among Spry isoforms, and several amino acids in this region (i.e. N53, P59 and G58) are also important for the Spry2-c-Cbl interaction. Consistently, mutations in these residues abrogate the Spry2-c-Cbl complex, regardless of the fact that Y55 phosphorylation is unaffected [362]. It should be noted that Spry2 interacts with c-Cbl even in the absence of growth factor stimulation, albeit this occurs at low levels [338, 339, 362]. This interaction may be attributed to the ability of Spry2 to interact with the RING finger of c-Cbl [363]. 80 Sprouty-2 kinase In view of the importance of the phosphorylation of Spry2 on Y55, efforts have been made to identify the kinase that phosphorylates Spry2. A recent study proposes that Src, activated by FGFR in a FRSZ-dependent fashion, phosphorylates Spry2 on Y55. The same study demonstrated that Spry2 is a direct substrate for Src in vivo and in vitro [364]. Another study suggests that EGFR phosphorylates Spry2 on Y55, in response to EGF stimulation [338]. Although a direct complex formation was not shown, EGFR was found to coimmunoprecipitate with Spry2, and immunoprecipitated EGFR was found to phosphorylate recombinant Spry2. Sprouty-2 phosphatase Phosphorylation frequently serves as a molecular switch to regulate protein activity. In many cases, including Spry2, phosphorylation leads to the activation of the protein function. In order to maintain normal cellular function, proteins that are activated by phosphorylation must be inactivated. This process is accomplished, in part, via the de-phosphorylation of proteins by phosphatases. In the case of Spry2, recent evidence implicates Shp2 as the phosphatase responsible for the inactivation of Spry2 [365]. Shp2 is a widely expressed protein-tyrosine phosphatase that mediates MAPK activation in response to growth factor 81 stimulation [366]. Shp2, through its SH2 domain is recruited to phosphotyrosine residues on RTK, where it dephosphorylates the receptor. Alternatively, Shp2 can function as a molecular adaptor for cellular proteins such as Grb2 [367]. Expression of Shp2 in murine myogenic cells decreases the phosphorylation of Spry2, and recombinant Shp2 reduces the level of phosphorylation on immunoprecipitated Spry2. Dephosphorylation of Spry2 by Shp2 results in the dissociation of Spry2 from Grb2, and decreases the inhibition of FGF signaling by Spry2. Sprouty-2 ubiquitination Another regulatory mechanism for Spry2 involves ubiquitination and proteosomal degradation. Stimulation with either EGF or FGF stimulation results in the ubiquitination of Spry2, followed by a decrease in the levels of Spry2 protein. The ubiquitination of Spry2 is attributed to its interaction with c-Cbl. Co-expression of Spry2 with c-Cbl results in ubiquitination of Spry2, while co—expression of Spry2 with c-Cbl mutants that are defective in their ability to ubiquitinate does not lead to ubiquitination of Spry2 [338, 339]. Also, expression of a Spry2 mutant that cannot interact with c-Cbl (Spry2Y55A) in cells with high levels of endogenous c-Cbl, does not result in ubiquitination and degradation of Spry2 following EGF-stimulation. 82 Phosphorylation of the C-terminus of Sprouty-2 An important characteristic of Spry2 is that it can sustain EGF-induced signaling, while also inhibiting FGF- induced signaling. A recent study by Rubin et al. [342] has made an important effort in explaining the molecular mechanism underlying this function. As noted above, Spry2 is phosphorylated on its N-terminus (Y55) in response to EGF or FGF stimulation. Nevertheless, Spry2 also becomes phosphorylated on several tyrosine residues located on its C—terminus. These sites are specifically phosphorylated in response to FGF stimulation, and they are necessary for maximal inhibitory function of Spry2 in FGF-induced signaling. Among the C-terminal tyrosine phosphorylation sites, residue Y227 is the most important for the efficient inhibition of FGFR signaling by Spry2 [342]. Thus, the inhibitory function Spry2 in RTK signaling appears to be a response to signals that stimulate phosphorylation of both the N-terminal Y55 and the C-terminal Y227 residues, whereas the function of Spry2 as an activator of RTK signaling may be a response to phosphorylation of Y55 alone. 83 Cellular functions of Sprouty Inhibition of receptor tyrosine kinase signaling The cellular function of Spry proteins that is conserved among different species, is the ability of Spry proteins to inhibit receptor tyrosine kinase signaling (Fig. 6A) [334]. To date, Spry proteins repress RTK signals induced by growth factors, including FGF, PDGF, VEGF and HGF/Met, and therefore play an important role in regulating the biological processes mediated by these signals [329, 340, 356, 368]. RTK signaling is initiated by an extracellular ligand that activates a transmembrane receptor, which undergoes dimerization and autophosphorylation, leading to the recruitment of adaptor and effector proteins, which engage a number of diverse intracellular pathways [40]. Nonetheless, Spry proteins appear to inhibit specific RTK-induced intracellular pathways, which results mainly in the activation of the MAPK-cascade leading to ERK. It does not, however, appear to regulate JNK and p38 MAPKS, at least in response to FGFR activation [369]. In some cases, Spry2 inhibits Akt activation [368, 370], although the mechanism and the consequence of this function have not been fully elucidated. The exact location along the RTK-Ras-Raf/MEK/ERK cascade where Spry exerts its inhibitory role remains under investigation. It appears that Spry acts in a signal 84 Figure 6. Regulation of RTK signaling by Spry. (A) Spry inhibits FGFR-mediated signaling. In the absence of Spry, binding of FGF to its receptor leads to tyrosine phosphorylation of the receptor and the scaffold protein FRSZ, producing docking sites for the Grb2:SOS complex. Recruitment of this complex to the localization of the receptor leads to the activation of Ras, which in turn activates the Raf/MEK/ERK cascade through the direct activation of Raf. When Spry is present, it interacts with Grb2 and prevents Grb2:SOS from being recruited to FGFR, leading to diminished activation of ERK in response to FGF. (B) Spry inhibits VEGFR-mediated signaling, through a different mechanism than that involved in the regulation of FGFR. In this context, Spry is involved in the regulation of Ras-lndependent Raf activation. In the absence of Spry, the activation of VEGFR results in the activation of PKC, in a PLCy-mediated pathway. PK05 activates the Raf/MEK/ERK pathway through the phosphorylation and activation of Raf. Spry interacts with Raf and prevents the activation of Raf by PKCé, inhibiting ERK activation in response to VEGF stimulation. (C) The activation of EGFR induces a signaling pathway that is similar to that induced by FGFR. EGFR is regulated by a number of mechanisms, one of which involves c- Cbl and CIN85. c-Cbl, an E3 ubiquitin ligase, ubiquitinates EGFR and targets it for degradation. CIN85 is involved in receptor endocytosis and facilitates the function of c-Cbl. These functions result in the silencing of receptor mediated signaling. Spry2 interacts with c-Cbl and CIN85, and prevents the degradation of EGFR. This results in sustained signaling activity from this receptor, culminating in enhanced activation of ERK. 85 86 specific and/or in a cellular context-specific manner. The conclusion that emerges from studies in different cellular systems is that Spry’s inhibition on RTK signaling occurs somewhere between the level of the receptor and Ras/Raf activation. The evidence leading to this conclusion is presented in the following paragraphs. One mechanism to explain the repression of FGF signaling by Spry focuses on the interaction of Spry with Grb2. In mouse myoblast cells (CZC12) this interaction is induced by growth factor stimulation, and results in the juxtaposition of the tyrosine phosphorylated Spry with the SH2 domain of Grb2. The consequence of this interaction in FGF signaling is to diminish the association of Grb2:SOS with FGFR adaptors like FRSZ and Shp2, which in turn diminishes the activation of Ras and Raf/MEK/ERK pathway [340]. It should be noted, however, that inducible expression of Spry in mouse fibroblasts (NIH3T3) showed that Spry had no effect on the formation of the FRSZ-Grb2:SOS complex, yet Spry still diminished the activation of Ras. Taken together, these results suggest that Spry may have cellular context-specific mechanism for preventing Ras activation in FGFR signaling [329]. Coexpression of Spry with constitutively active forms of Ras, Raf and MEK, in human embryonic kidney (HEK293) cells, found that Spry inhibits FGF signaling at the level of Raf activation [369]. 87 The ability of Spry to intercept RTK signaling at the level of Raf activation is also observed in human endothelial cells, where VEGF activates ERK in a Ras- independent fashion (Fig. 6B) [371]. In this system VEGFR, through its effector PLCy, activates PKCS and this in turn activates the Raf/MAPK cascade. Expression of Spry4 in this system prevents the Res-independent activation of ERK induced by VEGF, without affecting Ras-dependent ERK activation induced by EGF. Spry4 interacts with Raf, though its carboxy—teminal region. This interaction is necessary for its inhibitory function. Spry2 can also interact with Raf [371], suggesting that Spry2 has a similar effect in this pathway. The interaction between Spry and Raf has is also evident in melanoma cell lines [372]. In melanomas expressing wild type BRaf, both Spry2 and Spry4 interact with BRaf. In melanoma cells with BRaf mutations (V599E, V5990, L596V and K6OOE), however, neither of the two Sprys was able to complex with BRaf. This finding suggest that these residues on BRaf are important for the interaction between Raf and Spry2. Furthermore, transient down regulation of Spry2 by using Spry2-specific siRNA resulted in an increase in the activation of ERK in melanomas with wild type BRaf, whereas no effect was observed in melanomas with mutated BRaf. This suggests that mutant BRaf can bypass the inhibitory effect of Spry2 [372]. The latter may be an important mechanism by which BRaf escapes normal regulatory mechanism and induces malignant transformation. 88 Another proposed mechanism by which Spry inhibits RTK signaling involves an interaction between D. m. Spry and Drk, the 0m. homolog of RasGAP1 [327]. Through this interaction Spry facilitates the recruitment of GAP1 at Ras signaling complexes, thereby promoting the inactivation of Ras. A final line of evidence to support the contention that Spry inhibits RTK signaling upstream of Ras/Raf originates from a study with murine lung epithelial cells, where ectopic Spry2 complexes with Grb2, FRSZ, Raf, Shp2, GAP1, FGFR2b and SOS [331]. FGF stimulation enhanced the interactions of Spry2 with Grb2, Raf, and FRSZ while diminishing the interaction between Spry2 and Shp2 or GAP1. Ectopic Spry2 decreased MAPK activation in response to FGF stimulation, suggesting that the inhibitory role of Spry2 is a result of its ability to differentially interact with key components upstream of MAPK [331]. Although Spry inhibits RTK signaling through distinct mechanisms, the biological phenotype resulting from this activity includes a decrease in growth factor- induced proliferation, differentiation and angiogenesis [329, 356, 373]. The inhibitory effect of Spry on cellular proliferation is mediated not only by the ability of Spry to inhibit ERK activation, but also, by its ability to represses SRE- and Elk- mediated transcription [329, 340]. Moreover, ectopic expression of Spry proteins ectopic expression of Spry proteins represses the ability of fibroblasts and leiomyoma cells to form colonies in soft agar, an observation that is consistent 89 with the ability to repress MAPK activation and cellular proliferation [329, 368]. In addition, Spry attenuates the differentiation of rat phaeochromocytoma (P012) cells induced by FGF [329], consistent with the fact that Spry proteins inhibit FGF signaling. Spry also inhibits the branching and sprouting of small vessels during angiogenesis [335]. Activation of epidermal growth factor receptor signaling EGFR is a key mediator of signal transmission in response to extracellular cues. EGFR regulates many cellular functions including proliferation, differentiation and survival [374]. As such, EGFR must be strictly regulated in order to ensure normal cellular function. Uncontrolled function of EGFR can lead to enhanced proliferation, which may lead to cancer formation. Although there are a number of mechanisms to turn off the signaling activity of EGFR, the one that is affected by the function of Spry involves receptor ubiquitination and proteosomal/lysosomal degradation. The necessity to turn off EGFR signaling following ligand activation is met in part by c-Cbl, an E3 ubiquitin ligase enzyme that catalyzes the poly-ubiquitination of EGFR. The consequence of EGFR ubiquitination is apparent upon the internalization of the ligand-bound receptor. When the receptor is not ubiquitinated it is endocytosed and recycled back to the plasma membrane following the disruption of the ligand-receptor complex (i.e. the inactivation of the receptor). Instead, when the receptor is poly-ubiquitinated it is targeted for 90 degradation by the proteasomal/lysosomal pathway. The function of c-Cbl is mediated through distinct domains including an atypical SH2 domain that interacts with phosphotyrosine residues located on c-Cbl’s substrates. c-Cbl, also contains a RING-finger domain, which is critical for ubiquitination, because it interacts with E2 ubiquitin conjugating enzymes, thus recruiting such factors to the location of the substrate [361]. There, E2 enzymes catalyze the ubiquitination of the substrate. In certain situation more than one ubiquitin moiety is added to the substrate, a modification that targets the substrate for degradation by the proteasome. Upon EGF stimulation, Spry2 can sustain RTK signaling, leading to sustained activation of ERK. Mechanistically this is attributed to the ability of Spry2 to interact with c-Cbl and inhibit the c-Cbl-induced degradation of EGFR [360, 363]. The interaction between Spry and c-Cbl reduces the level of EGFR ubiquitination by c-Cbl, and thus Spry can sustain EGFR levels and signaling activity. The later has been demonstrated in PC12 cells, which have been paradigmatically used to study the effects of RTK signaling on cell fate, given the fact that they can be induced to proliferate or differentiate, depending on the duration of RTK signaling [375]. In PC12 cells, some growth factors, such as EGF, induce transient activation of MAPK and promote cellular proliferation, whereas other growth factors, such as nerve growth factor (NGF) or FGF, induce sustained ERK activation leading to differentiation and neurite outgrowth. However, when PC12 cells expressing Spry2 are stimulated with EGF, they display sustained ERK 91 activation and undergo differentiation, consistent with the ability of Spry2 to prevent EGFR degradation induced by c-Cbl [363]. The ability of Spry proteins to interact with c-Cbl is conserved in D. m. Spry and mammalian Spry1 and Spry2. Nevertheless, this function has almost exclusively been studied in the context of Spry2. Spry4 fails to interact with c-Cbl and cannot inhibit EGFR degradation induced by c-Cbl [360]. As described above, the interaction between Spry2 and c-Cbl, involves Spry2 phosphotyrosine residue on position 55 and c-Cbl’s SH2 domain [334]. Furthermore, amino acids flanking Y55 on Spry2 are also important for this interaction [355, 362]. Structural differences in this region between Spry2 and Spry4 may account for their different ability to regulate EGFR degradation. In fact, mutation of the Spry2 residues surrounding Y55 to the amino acids present in the same region of Spry4, abrogates the ability of Spry2 to bind c-Cbl [355]. In addition to preventing EGFR degradation, Spry2 can also diminish EGFR endocytosis at an early stage of the lntemalization/trafficking process [363, 376]. In the absence of Spry2, EGFR localizes in endocytotic vesicles upon EGF- stimulation, whereas upon expression of Spry2, EGFR remains at the plasma membrane. The role of Spry2 in the regulation of EGFR endocytosis is also supported by recent evidence that Spry2 can interact in vivo and in vitro with CIN85 [343]. 92 CIN85 is part of an endocytotic complex that assists endocytosis by c-Cbl. It has been proposed that CIN85 leads to clustering of c-Cbl, facilitating c-CbI-induced EGFR endocytosis and degradation. CIN85 contains three SH3 domains (A, B and C).through which it interacts with proline rich regions (PxxxPR). Spry2 contains two such motif on positions 64 and 309, which interact with the SH3A domain of CIN85. In addition, Spry2 contains another similar motif (xxPxPR) on position 72, which interacts with the SH3C domain of CIN85. These proline-rich motifs are also present in Spry1, thus enabling this Spry to interact with CIN85. D. m. Spry and Spry4, however, lack such motifs and cannot interact with CIN85 [343]. The interaction of Spry2 with CIN85 is important for the inhibition of c-Cbl induced degradation of EGFR, because a Spry2 mutant that cannot interact with CIN85 2R64'72'309A) fails to inhibit EGFR degradation, regardless of the presence of ($er wild type Y55 residue (i.e., this mutant Spry2 retains an intact ability for c-Cbl interaction). Finally, Spry2 associates with both c-Cbl and CIN85 to form a ternary complex that is necessary for inhibition of EGFR degradation and promotion of neurite outgrowth in PC12 cells [343]. The ability of Spry2 to inhibit c-Cbl induced degradation of EGFR implies that Spry may prevent the degradation of additional c-Cbl substrates. However, this function of Spry2 may be restricted only to EGFR, because although c-Cbl targets 93 FGFR and FRSZ for degradation, Spry2 doesn’t seem to have an inhibitory effect under these circumstances [362]. Regulation of integrin-mediated cell spreading by Sprouty-4 The ability of Spry to regulate cell spreading has been studied only in the context of Spry4. Spry4 interacts with TESK1 through its carboxy-terminal region [377]. TESK1 is a serine/threonine kinase that phosphorylates cofilin, an actin binding protein that is involved in actin depolymerization and severance of actin stress fibers [378, 379]. TESK1 phosphorylation inhibits cofilin, and cofilin-induced actin disassembly. Also, TESK1 plays an important role in integrin-mediated actin remodeling and cell spreading. The interaction between Spry4 and TESK results in the inhibition of the kinase activity of TESK1 in vitro, as well as in the decrease in the levels of phosphorylated cofilin [380]. Thus, Spry4 negatively regulates cell spreading through the inhibition of TESK1, which results in sustained cofilin activity. Phosphorylation of the conserve tyrosine residue on the N-terrninal of Spry4 (Y75) is not necessary for its interaction with TESK1, or for its ability to suppress cell spreading [380]. Regulation of cell migration by Sprouty Spry proteins inhibit cell migration during Xenopus embryogenesis [332], as well as in wound healing assays with mammalian cells [373]. In the latter case, Spry1, Spry2 and Spry4 antagonize growth factor stimulated cellular migration. This 94 ability of Spry is dependent on its cysteine rich carboxy-terminal [373], and appears to be mediated in part by PTP1B and p130Cas [381, 382]. These studies, in which Spry2 protein was transduced in HeLa cells, demonstrated an increase in the levels and activity of PTP1B phosphatase in the soluble fraction. p1300‘as, a substrate for PTP1B, exhibits reduced levels of phosphorylation following transduction of Spry2, a finding that is consistent with the increase in the activity of PTP1B by Spry2. What’s more expression of p130Gas attenuates Spry2’s inhibition of growth factor induced cellular migration [381]. Furthermore, Spry2 represses the activation of R301 in wound healing assays, suggesting that the regulation of cellular migration by Spry2 is dependent of its effect of Rac1 activity. Expression of constitutively active Rac1 attenuates the inhibition of migration by Spry2 [382]. Instead, constitutive active Rac1 had no effect on the inhibitory function of Spry2 in cellular proliferation [381], suggesting that Spry can engage specific pathways to carry through its effect on cellular function. Sprouty deficient mice Recently, Basson et al [383], Shim et al [384],. and Taketomi et al. [370] successfully generated knockout mice models for Spry1 and Spry2. These models have elucidated important roles of Spry genes in mammalian development Sprouty-14' 95 Homozygous Spry1“ mice, which were born at expected Mendelian ratios, display reduced post-natal viability [383]. A small subset (21%) of Spry1-l— mice die within 48 hrs, whereas another population (71%) die within five months. The surviving animals have significant kidney malformations. Normal kidney development involves the outgrowth of a single ureteric bud from the Wolfian duct during early development [385]. Glial derived growth factor (GDNF), and its tyrosine kinase receptor c-Ret are part of a signaling pathway that is important for ureteric bud formation[386-388]. Spry1-/- mice develop supemumerary ureteric buds from the Wolfian duct, which results in multiple ureters and a multiplex kidney [383]. An important characteristic of Spry1-/- mice is their sensitivity to GDNF signaling. This is indicated by the levels of active ERK in ectopic sites, as well as by the expression of Wnt11. Normally, Wnt11 is expressed at the tips of ureteric buds [389], where Ret signaling is active, but in Spry1'/' mice Wnt11 expression is extended to more anterior sites and in discrete ectopic sites suggesting an overactive c-Ret signaling pathway [383, 390]. The phenotype of Spry1'/' mice is mediated by the GDNF sensitization of kidney tissue, because reduction of GDNF gene dosage, as a result of crossing Spry1"‘ with Gdnfi” (Spry1"';Gan/') mice, reverts the phenotype of Spry1'/' mice. Together, these findings suggest that Spry1 regulates normal ureteric bud and kidney development, through suppression of GDNF/Ret signaling. Sprouty-2" 96 Half of the Spry2" mice die within six weeks after birth, while the rest can survive for at least six months. These animals are smaller in size compared to litterrnates [370]. Notably, the surviving Spry2" mice are characterized by hearing impairment [384], as well as neuronal hyperplasia and esophageal achalasia (constriction of the lower part of the esophagus) [370]. The sense of hearing involves the conversion of the sound waves into vibrational energy at the tympanic membrane and the transmission through the middle ear to the organ of Corti in the inner ear, where vibrational energy is converted into electrical impulses that are transmitted into the brain. The hearing impairment in Spry2"' mice reflects a disruption in the architecture of the organ of Corti [384]. While this organ normally consists of three rows of longitudinal outer hair cells, the organ of Corti in Spry2" mice consists of four rows. In addition, Spry2" contained an ectopic pillar cell, leading to the formation of an ectopic Corti-like tunneL The normal development of the organ of Corti is dependent in part on FGF8, which is secreted by inner hair cells and acts on the FGFR3 receptors, which are located on the outer hair cells and the pillar cells [384, 391, 392]. It is possible that in Spry2" mice, FGFR signaling is overactive, resulting in the noted malformations in the organ of Corti. Indeed, reduction of FGF8 gene dosage in Spry2 null mice, achieved by crossing these mice with FGF8“ mice (Spry2‘/' ;Fgf8+/') prevents the extra pillar formation and partially reverts the hearing loss 97 phenotype of Spry2”‘ mice. This suggests that Spry2 suppresses FGF8 signaling in the organ of Corti and promotes normal development of this organ [384]. Phenotypically, Spry2" mice are also characterized by abnormal intestinal motility and esophageal achalasia, due to lower esophageal sphincter (LES) hypercontraction [370]. Spry2" mice demonstrate hyperplasticity in the enteric nervous plexus (ENS) and have elevated levels of muscarinic (M2)-acetylcholine receptors in the neuro-muscular junctions of esophagus, which may be responsible for the hypercontraction of LES. GDNF/Ret signaling, which is necessary for the survival of ENS [386] is increased in neuronal cells of Spry2/‘ mice, suggesting that the ability of Spry2 to repress GDNF signaling is important for preventing ENS hyperplasia, and esophageal achalasia. In fact, anti-GDNF antibodies can correct ENS hyperplasia and reduce the esophageal dilation in Spry2” mice [370]. Sprouty proteins in cancer The ability of Spry proteins to regulate RTK makes them good candidates as cancer genes. The expression of Spry2 gene is altered in several types of cancer. Spry1 and Spry2 are down regulated in breast cancer [393], and expression of Spry2 in the breast cancer cells (MCF7) reduces the ability of these cells to form 2Y5“, a mutant that cannot repress growth factor tumors in athymic mice. Spry induced ERK activation, fails to abrogate the tumor forming ability of MCF7 cells [393]. 98 Spry1 and Spry2 are also down regulated in prostate cancer. Spry1 was found to be decreased in 40% of prostate cancers, when these were compared with matched normal prostate [394]. Spry2 expression is decreased in invasive prostate cancer and clinical prostate cancer, when these were compared to benign prostatic hyperplasia (BPH) [395]. This reduced expression is attributed to epigenetic inactivation, through hypermethylation of the Spry2 promoter. The Spry2 gene contains a large CpG island spanning positions ~500 and +950 relative to the putative transcription start site. This region is hypermethylated in the high grade clinical prostate cancers, whereas in BPH this region is relatively non-methylated [395]. Interestingly, Spry2 expression is elevated in melanomas with BRafVSOOE 061R mutations or in melanomas with NRas mutations, compared to cells without mutations [396]. An independent study found that Spry2 was expressed at higher fV599E levels in melanomas with BRa compared to melanomas with wild type BRaf and normal melanocytes [372]. Although additional studies are needed to determine the precise role of Spry in cancer, initial evidence supports the hypothesis that Spry proteins repress malignant transformation, through their ability to repress RTK signaling. Nevertheless the increased expression of Spry2 in some types of cancer, as well 99 as its ability to sustain EGFR signaling, suggests that Spry2 promotes cancer progression. Whether this is true remains to be seen. 100 References 10. 11. 12. 13. 14. Jemal, A., Murray, T., Samuels, A., Ghafoor, A., Ward, E. and Thun, M.J., Cancer statistics, 2003. CA Cancer J Clin, 2003. 53(1): p. 5-26. Parkin, D.M., Global cancer statistics in the year 2000. Lancet Oncol, 2001 . 2(9): p. 533-543. Kumar, V., Abbas, A. and Fausto, N., Robbins and Cotran Pathologic Basis of Disease. 7th Edition. 2005: Elsevier Saunders. 269-342. Macaluso, M., Paggi, MG. and Giordano, A., Genetic and epigenetic alterations as hallmarks of the intricate road to cancer. Oncogene, 2003. 22(42): p. 6472-6478. Nowell, P.C., The clonal evolution of tumor cell populations. Science, 1976. 194(4260): p. 23-28. Fearon, ER. and Vogelstein, B., A genetic model for colorectal tumorigenesis. Cell, 1990. 61(5): p. 759-767. Renan, M.J., How many mutations are required for tumorigenesis? Implications from human cancer data. Mol Carcinog, 1993. 7(3): p. 139- 146. Bishop, J.M., Enemies within: the genesis of retrovirus oncogenes. Cell, 1981. 23(1): p. 5-6. Weinberg, R.A., Tumor suppressor genes. Science, 1991. 254(5035): p. 1 138-1 146. Ponder, B.A., Cancer genetics. Nature, 2001. 411(6835): p. 336-341. Parker, R.C., Varmus, HE. and Bishop, J.M., Cellular homologue (c-src) of the transforming gene of Rous sarcoma virus: isolation, mapping, and transcriptional analysis of c-src and flanking regions. Proc Natl Acad Sci U S A, 1981. 78(9): p. 5842-5846. Pelengaris, S., Khan, M. and Evan, G., c-MYC: more than just a matter of life and death. Nat Rev Cancer, 2002. 2(10): p. 764-776. Yeatman, T.J., A renaissance for SRC. Nat Rev Cancer, 2004. 4(6): p. 470-480. Prober, DA. and Edgar, B.A., Growth regulation by oncogenes-mew insights from model organisms. Curr Opin Genet Dev, 2001. 11(1): p. 19- 26. 101 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. Todd, R. and Wong, D.T., Oncogenes. Anticancer Res, 1999. 19(6A): p. 4729-4746. Anderson, M.W., Reynolds, S.H., You, M. and Maronpot, R.M., Role of proto-oncogene activation in carcinogenesis. Environ Health Perspect, 1992. 98: p. 13-24. Parada, L.F., Tabin, C.J., Shih, C. and Weinberg, R.A., Human EJ bladder carcinoma oncogene is homologue of Harvey sarcoma virus ras gene. Nature, 1982. 297(5866): p. 474-478. Feinberg, AP, The epigenetics of cancer etiology. Semin Cancer Biol, 2004. 14(6): p. 427-432. Alitalo, K., Amplification of cellular oncogenes in cancer cells. Med Biol, 1984. 62(6): p. 304-317. Ren, R., Mechanisms of BCR-ABL in the pathogenesis of chronic myelogenous leukaemia. Nat Rev Cancer, 2005. 5(3): p. 172-183. Spencer, CA. and Groudine, M., Control of c-myc regulation in normal and neoplastic cells. Adv Cancer Res, 1991. 56: p. 1-48. Cowell, J.K., One hundred years of retinoblastoma research. From the clinic to the gene and back again. Ophthalmic Paediatr Genet, 1989. 10(2): p. 75-88. Gennett, IN. and Cavenee, W.K., Molecular genetics in the pathology and diagnosis of retinoblastoma. Brain Pathol, 1990. 1(1): p. 25-32. Knudson, A.G., Jr., Mutation and cancer: statistical study of retinoblastoma. Proc Natl Acad Sci U S A, 1971. 68(4): p. 820-823. Friend, S.H., Bemards, R., Rogelj, 8., Weinberg, R.A., Rapaport, J.M., Albert, D.M. and Dryja, TR, A human DNA segment with properties of the gene that predisposes to retinoblastoma and osteosarcoma. Nature, 1986. 323(6089): p. 643-646. Fung, Y.K., Murphree, A.L., T'Ang, A., Qian, J., Hinrichs, SH. and Benedict, W.F., Structural evidence for the authenticity of the human retinoblastoma gene. Science, 1987. 236(4809): p. 1657-1661. Lee, W.H., Shew, J.Y., Hong, F.D., Sery, T.W., Donoso, L.A., Young, L.J., Bookstein, R. and Lee, E.Y., The retinoblastoma susceptibility gene encodes a nuclear phosphoprotein associated with DNA binding activity. Nature, 1987. 329(6140): p. 642-645. 102 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. Classon, M. and Harlow, E., The retinoblastoma tumour suppressor in development and cancer. Nat Rev Cancer, 2002. 2( 12): p. 910-917. Macleod, K., Tumor suppressor genes. Curr Opin Genet Dev, 2000. 10(1): p. 81-93. Knudson, A.G., Two genetic hits (more or less) to cancer. Nat Rev Cancer, 2001. 1(2): p. 157-162. Kinzler, K.W. and Vogelstein, 8., Lessons from hereditary colorectal cancer. Cell, 1996. 87(2): p. 159-170. Fodde, R., The multiple functions of tumour suppressors: it's all in APC. Nat Cell Biol, 2003. 5(3): p. 190-192. Powell, S.M., Zilz, N., Beazer-Barclay, Y., Bryan, T.M., Hamilton, S.R., Thibodeau, S.N., Vogelstein, B. and Kinzler, K.W., APC mutations occur early during colorectal tumorigenesis. Nature, 1992. 359(6392): p. 235- 237. Bronner, C.E., Baker, S.M., Morrison, P.T., Warren, G., Smith, L.G., Lescoe, M.K., Kane, M., Earabino, C., Lipford, J., Lindblom, A. and et al., Mutation in the DNA mismatch repair gene homologue hMLH1 is associated with hereditary non-polyposis colon cancer. Nature, 1994. 368(6468): p. 258-261. Fishel, R., Lescoe, M.K., Rao, M.R., Copeland, N.G., Jenkins, N.A., Garber, J., Kane, M. and Kolodner, R., The human mutator gene homolog MSH2 and its association with hereditary nonpolyposis colon cancer. Cell, 1993. 75(5): p. 1027-1038. Lengauer, C., Kinzler, K.W. and Vogelstein, 3., Genetic instability in colorectal cancers. Nature, 1997. 386(6625): p. 623-627. Egger, G., Liang, G., Aparicio, A. and Jones, P.A., Epigenetics in human disease and prospects for epigenetic therapy. Nature, 2004. 429(6990): p. 457-463. Claus, R. and Lubbert, M., Epigenetlc targets in hematopoietic malignancies. Oncogene, 2003. 22(42): p. 6489-6496. Hanahan, D. and Weinberg, RA, The hallmarks of cancer. Cell, 2000. 100(1): p. 57-70. Schlessinger, J., Cell signaling by receptor tyrosine kinases. Cell, 2000. 103(2): p. 211-225. 103 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. Tari, AM. and Lopez-Berestein, G., GRBZ: a pivotal protein in signal transduction. Semin Oncol, 2001. 28(5 Suppl 16): p. 142-147. Birge, RB. and Hanafusa, H., Closing in on SH2 specificity. Science, 1993. 262(5139): p. 1522-1524. ‘Avruch, J., Khokhlatchev, A., Kyriakis, J.M., Luo, Z., Tzivion, G., Vawas, D. and Zhang, X.F., Ras activation of the Raf kinase: tyrosine kinase recruitment of the MAP kinase cascade. Recent Prog Horm Res, 2001. 56: p. 127-155. Margolis, B. and Skolnik, E.Y., Activation of Ras by receptor tyrosine kinases. J Am Soc Nephrol, 1994. 5(6): p. 1288-1299. Chardin, P., Camonis, J.H., Gale, N.W., van Aelst, L., Schlessinger, J., Wigler, M.H. and Bar-Sagi, D., Human 8031: a guanine nucleotide exchange factor for Ras that binds to GRBZ. Science, 1993. 260(5112): p. 1338-1343. Gibbs, J.B., Sigal, l.S., Poe, M. and Scolnick, E.M., Intrinsic GTPase activity distinguishes normal and oncogenic ras p21 molecules. Proc Natl Acad Sci U S A, 1984. 81(18): p. 5704-5708. Sweet, R.W., Yokoyama, S., Kamata, T., Feramisco, J.R., Rosenberg, M. and Gross, M., The product of ras is a GTPase and the T24 oncogenic mutant is deficient in this activity. Nature, 1984. 311(5983): p. 273-275. Satoh, T., Nakafuku, M. and Kaziro, Y., Function of Ras as a molecular switch in signal transduction. J Biol Chem, 1992. 267(34): p. 24149-24152. Vojtek, AB. and Der, C.J., Increasing complexity of the Ras signaling pathway. J Biol Chem, 1998. 273(32): p. 19925-19928. Aktas, H., Cai, H. and Cooper, G.M., Ras links growth factor signaling to the cell cycle machinery via regulation of cyclin D1 and the Cdk inhibitor p27KlP1. Mol Cell Biol, 1997. 17(7): p. 3850-3857. Chambers, AF. and Tuck, A.B., Ras-responsive genes and tumor metastasis. Crit Rev Oncog, 1993. 4(2): p. 95-114. Giehl, K., Oncogenic Ras in tumour progression and metastasis. Biol Chem, 2005. 386(3): p. 193-205. Pantazis, P., Pelicci, P.G., Dalla-Favera, R. and Antoniades, H.N., Synthesis and secretion of proteins resembling platelet-derived growth factor by human glioblastoma and fibrosarcoma cells in culture. Proc Natl Acad Sci U S A, 1985. 82(8): p. 2404-2408. 104 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. Suzuki, K. and Yamada, S., Ascites sarcoma 180, a tumor associated with hypercalcemia, secretes potent bone-resorbing factors including transforming growth factor alpha, interleukin-1 alpha and interleukin-6. Bone Miner, 1994. 27(3): p. 219-233. Koh, J.S., Trent, J., Chen, L., El-Naggar, A., Hunt, K., Pollock, R. and Zhang, W., Gastrointestinal stromal tumors: overview of pathologic features, molecular biology, and therapy with imatinib mesylate. Histol Histopathol, 2004. 19(2): p. 565-574. Yarden, Y. and Ullrich, A., Growth factor receptor tyrosine kinases. Annu Rev Biochem, 1988. 57: p. 443-478. 803, J.L., ras oncogenes in human cancer: a review. Cancer Res, 1989. 49(17): p. 4682-4689. Melo, J.V., The molecular biology of chronic myeloid leukaemia. Leukemia, 1996. 10(5): p. 751-756. Krengel, U., Schlichting, L., Scherer, A., Schumann, R., Frech, M., John, J., Kabsch, W., Pal, BF. and Wittinghofer, A., Three-dimensional structures of H-ras p21 mutants: molecular basis for their inability to function as signal switch molecules. Cell, 1990. 62(3): p. 539-548. Macaluso, M., Russo, G., Cinti, C., Bazan, V., Gebbia, N. and Russo, A., Ras family genes: an interesting link between cell cycle and cancer. J Cell Physiol, 2002. 192(2): p. 125-130. Tong, L., Milbum, M.V., de Vos, AM. and Kim, S.H., Structure of ras proteins. Science, 1989. 245(4915): p. 244. Van Etten, R.A., Cycling, stressed—out and nervous: cellular functions of c- Abl. Trends Cell Biol, 1999. 9(5): p. 179-186. Shtivelman, E., Lifshitz, B., Gale, RP. and Canaani, E., Fused transcript of abl and bcr genes in chronic myelogenous leukaemia. Nature, 1985. 315(6020): p. 550-554. Muller, A.J., Young, J.C., Pendergast, A.M., Pondel, M., Landau, N.R., Littman, DR. and Witte, O.N., BCR first exon sequences specifically activate the BCR/ABL tyrosine kinase oncogene of Philadelphia chromosome-positive human leukemias. Mol Cell Biol, 1991. 11(4): p. 1785-1792. Ren, S.Y., Xue, F., Feng, J. and Skorski, T., Intrinsic regulation of the interactions between the SH3 domain of p85 subunit of phosphatidylinositol-3 kinase and the protein network of BCR/ABL oncogenic tyrosine kinase. Exp Hematol, 2005. 33(10): p. 1222-1228. 105 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. Pluk, H., Dorey, K. and Superti-Furga, G., Autoinhibition of c—Abl. Cell, 2002. 108(2): p. 247-259. Darnell, J.E., Jr., Transcription factors as targets for cancer therapy. Nat Rev Cancer, 2002. 2(10): p. 740-749. Nebert, D.W., Transcription factors and cancer: an overview. Toxicology, 2002. 181-182: p. 131-141. Jenkins, R.B., Qian, J., Lieber, MM. and Bostwick, D.G., Detection of c- myc oncogene amplification and chromosomal anomalies in metastatic prostatic carcinoma by fluorescence in situ hybridization. Cancer Res, 1997. 57(3): p. 524-531. Mimori, K., Mori, M., Shiraishi, T., Tanaka, S., Haraguchi, M., er, H., Shirasaka, C. and Akiyoshi, T., Expression of ornithine decarboxylase mRNA and c-myc mRNA in breast tumours. Int J Oncol, 1998. 12(3): p. 597-601. Schlagbauer—Wadl, H., Griffioen, M., van Elsas, A., Schrier, P.l., Pustelnik, T., Eichler, H.G., Wolff, K., Pehamberger, H. and Jansen, 8., Influence of increased c-Myc expression on the growth characteristics of human melanoma. J Invest Dermatol, 1999. 112(3): p. 332-336. Elend, M. and Eilers, M., Cell growth: downstream of Myc - to grow or to cycle? Curr Biol, 1999. 9(24): p. R936-938. Evan, G.l., Wyllie, A.H., Gilbert, C.S., Littlewood, T.D., Land, H., Brooks, M., Waters, C.M., Penn, L2. and Hancock, D.G., Induction of apoptosis in fibroblasts by c-myc protein. Cell, 1992. 69(1): p. 119-128. La Rocca, S.A., Crouch, DH. and Gillespie, D.A., c-Myc inhibits myogenic differentiation and myoD expression by a mechanism which can be dissociated from cell transformation. Oncogene, 1994. 9(12): p. 3499- 3508. Massague, J., G1 cell-cycle control and cancer. Nature, 2004. 432(7015): p. 298-306. Hartwell, L.H. and Kastan, M.B., Cell cycle control and cancer. Science, 1994. 266(5192): p. 1821-1828. Murray, A.W., Recycling the cell cycle: cyclins revisited. Cell, 2004. 116(2): p. 221-234. Morgan, 0.0., Principles of CDK regulation. Nature, 1995. 374(6518): p. 131-134. 106 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. Schwartz, G.K., CDK inhibitors: cell cycle arrest versus apoptosis. Cell Cycle, 2002. 1(2): p. 122-123. Trimarchi, J.M. and Lees, J.A., Sibling rivalry in the E2F family. Nat Rev Mol Cell Biol, 2002. 3(1): p. 11-20. Ohtani, K., Implication of transcription factor E2F in regulation of DNA replication. Front Biosci, 1999. 4: p. D793-804. Seville, L.L., Shah, N., Westwell, AD. and Chan, W.C., Modulation of pRB/E2F functions in the regulation of cell cycle and in cancer. Curr Cancer Drug Targets, 2005. 5(3): p. 159-170. Bates, S. and Peters, G., Cyclin D1 as a cellular proto-oncogene. Semin Cancer Biol, 1995. 6(2): p. 73-82. Hosokawa, Y. and Arnold, A., Mechanism of cyclin D1 (CCND1, PRAD1) overexpression in human cancer cells: analysis of allele-specific expression. Genes Chromosomes Cancer, 1998. 22(1): p. 66-71. Moroy, T. and Geisen, C., Cyclin E. Int J Biochem Cell Biol, 2004. 36(8): p. 1424-1439. Sauer, K. and Lehner, CF, The role of cyclin E in the regulation of entry into S phase. Prog Cell Cycle Res, 1995. 1: p. 125-139. Morris, L., Allen, KB and La Thangue, N.B., Regulation of E2F transcription by cyclin E-Cdk2 kinase mediated through p300/CBP co- activators. Nat Cell Biol, 2000. 2(4): p. 232-239. Yam, C.H., Fung, T.K. and Poon, R.Y., Cyclin A in cell cycle control and cancer. Cell Mol Life Sci, 2002. 59(8): p. 1317-1326. Pei, X.H. and Xiong, Y., Biochemical and cellular mechanisms of mammalian CDK inhibitors: a few unresolved issues. Oncogene, 2005. 24(17): p. 2787-2795. Sherr, C.J., The lNK4a/ARF network in tumour suppression. Nat Rev Mol Cell Biol, 2001. 2(10): p. 731-737. Kastan, MB. and Bartek, J., Cell-cycle checkpoints and cancer. Nature, 2004. 432(7015): p. 316-323. Gottifredi, V. and Prives, C., The S phase checkpoint: when the crowd meets at the fork. Semin Cell Dev Biol, 2005. 16(3): p. 355-368. O'Connell, M.J., Walworth, NC. and Carr, A.M., The GZ-phase DNA- damage checkpoint. Trends Cell Biol, 2000. 10(7): p. 296-303. 107 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. Peeper, D.S., van der Eb, A.J. and Zantema, A., The G1/S cell-cycle checkpoint in eukaryotic cells. Biochim Biophys Acta, 1994. 1198(2-3): p. 215-230. Hicks, G.G., Egan, S.E., Greenberg, AH. and Mowat, M., Mutant p53 tumor suppressor alleles release ras-induced cell cycle growth arrest. Mol Cell Biol, 1991. 11(3): p. 1344-1352. Shaw, P., Bovey, R., Tardy, S., Sahli, R., Sordat, B. and Costa, J., Induction of apoptosis by wild-type p53 in a human colon tumor-derived cell line. Proc Natl Acad Sci U S A, 1992. 89(10): p. 4495-4499. Marx, J., New link found between p53 and DNA repair. Science, 1994. 266(5189): p. 1321-1322. Levine, A.J., p53, the cellular gatekeeper for growth and division. Cell, 1997. 88(3): p. 323-331. Midgley, CA. and Lane, D.P., p53 protein stability in tumour cells is not determined by mutation but is dependent on Mdm2 binding. Oncogene, 1997. 15(10): p. 1179-1189. Vousden, K.H. and Lu, X., Live or let die: the cell's response to p53. Nat Rev Cancer, 2002. 2(8): p. 594-604. Li, Y., Jenkins, C.W., Nichols, MA. and Xiong, Y., Cell cycle expression and p53 regulation of the cyclin-dependent kinase inhibitor p21. Oncogene, 1994. 9(8): p. 2261-2268. el-Deiry, W.S., Harper, J.W., O'Connor, P.M., Velculescu, V.E., Canman, C.E., Jackman, J., Pietenpol, J.A., Burrell, M., Hill, D.E., Wang, Y. and et al., WAF1/ClP1 is induced in p53-mediated G1 arrest and apoptosis. Cancer Res, 1994. 54(5): p. 1169-1174. Hayflick, L. and Moorhead, PS, The serial cultivation of human diploid cell strains. Exp Cell Res, 1961. 25: p. 585-621. Wright, W.E., Pereira-Smith, OM. and Shay, J.W., Reversible cellular senescence: implications for immortalization of normal human diploid fibroblasts. Mol Cell Biol, 1989. 9(7): p. 3088-3092. Collins, K., Mammalian telomeres and telomerase. Curr Opin Cell Biol, 2000. 12(3): p. 378-383. Harley, C.B., Futcher, AB. and Greider, C.W., Telomeres shorten during ageing of human fibroblasts. Nature, 1990. 345(6274): p. 458-460. 108 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. Mathon, N.F. and Lloyd, A.C., Cell senescence and cancer. Nat Rev Cancer, 2001. 1(3): p. 203-213. Hayflick, L., Mortality and immortality at the cellular level. A review. Biochemistry (Mosc), 1997. 62(11): p. 1180-1190. Shay, J.W. and Bacchetti, S., A survey of telomerase activity in human cancer. Eur J Cancer, 1997. 33(5): p. 787-791. Bryan, TM. and Cech, T.R., Telomerase and the maintenance of chromosome ends. Curr Opin Cell Biol, 1999. 11(3): p. 318-324. Vaziri, H. and Benchimol, S., Reconstitution of telomerase activity in normal human cells leads to elongation of telomeres and extended replicative life span. Curr Biol, 1998. 8(5): p. 279-282. Henson, J.D., Neumann, A.A., Yeager, TR. and Reddel, R.R., Alternative lengthening of telomeres in mammalian cells. Oncogene, 2002. 21(4): p. 598-610. Hengartner, MO, The biochemistry of apoptosis. Nature, 2000. 407(6805): p. 770-776. lgney, PH. and Krammer, P.H., Death and anti-death: tumour resistance to apoptosis. Nat Rev Cancer, 2002. 2(4): p. 277-288. Schmitz, l., Kirchhoff, S. and Krammer, P.H., Regulation of death receptor- mediated apoptosis pathways. Int J Biochem Cell Biol, 2000. 32(11-12): p. 1 123-1 136. Sprick, M.R., Weigand, M.A., Rieser, E., Rauch, C.T., Juo, P., Blenis, J., Krammer, PH. and Walczak, H., FADD/MORT1 and caspase-8 are recruited to TRAIL receptors 1 and 2 and are essential for apoptosis mediated by TRAIL receptor 2. Immunity, 2000. 12(6): p. 599-609. Thornberry, NA, The caspase family of cysteine proteases. Br Med Bull, 1997. 53(3): p. 478-490. Kischkel, F.C., Hellbardt, S., Behnnann, l., Gerrner, M., Pawlita, M., Krammer, PH. and Peter, M.E., Cytotoxicity-dependent APO-1 (Fas/CDQ5)-associated proteins form a death-inducing signaling complex (DISC) with the receptor. Embo J, 1995. 14(22): p. 5579-5588. lrmler, M., Thome, M., Hahne, M., Schneider, P., Hofmann, K., Steiner, V., Bodmer, J.L., Schroter, M., Burns, K., Mattmann, C., Rimoldi, 0., French, LE. and Tschopp, J., Inhibition of death receptor signals by cellular FLIP. Nature, 1997. 388(6638): p. 190-195. 109 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. Rich, T., Allen, R.L. and Wyllie, A.H., Defying death after DNA damage. Nature, 2000. 407(6805): p. 777-783. Martinou, JC. and Green, D.R., Breaking the mitochondrial barrier. Nat Rev Mol Cell Biol, 2001. 2(1): p. 63-67. Verhagen, A.M., Ekert, P.G., Pakusch, M., Silke, J., Connolly, L.M., Reid, G.E., Moritz, R.L., Simpson, R.J. and Vaux, D.L., Identification of DlABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell, 2000. 102(1): p. 43-53. Soengas, M.S., Capodieci, P., Polsky, D., Mora, J., Esteller, M., Opitz- Araya, X., McCombie, R., Herman, J.G., Gerald, W.L., Lazebnik, Y.A., Cordon-Cardo, C. and Lowe, S.W., Inactivation of the apoptosis effector Apaf—1 in malignant melanoma. Nature, 2001. 409(6817): p. 207-211. Bertram, J.S., The molecular biology of cancer. Mol Aspects Med, 2000. 21(6): p. 167-223. Risinger, MA. and Groden, J., Cross/inks and crosstalk: human cancer syndromes and DNA repair defects. Cancer Cell, 2004. 6(6): p. 539-545. Fleck, O. and Schar, P., Translesion DNA synthesis: little fingers teach tolerance. Curr Biol, 2004. 14(10): p. R389-391. Robbins, J.H., Kraemer, K.H. and Flaxman, B.A., DNA repair in tumor cells from the variant form of xeroderma pigmentosum. J Invest Dermatol, 1975. 64(3): p. 150-155. Berneburg, M. and Lehmann, A.R., Xeroderma pigmentosum and related disorders: defects in DNA repair and transcription. Adv Genet, 2001. 43: p. 71-102. Maher, V.M., Rowan, L.A., Silinskas, K.C., Kateley, SA. and McCormick, J.J., Frequency of UV-induced neoplastic transformation of diploid human fibroblasts is higher in xeroderma pigmentosum cells than in normal cells. Proc Natl Acad Sci U S A, 1982. 79(8): p. 2613-2617. Masutani, C., Kusumoto, R., Yamada, A., Dohmae, N., Yokoi, M., Yuasa, M., Araki, M., lwai, S., Takio, K. and Hanaoka, F., The XPV (xeroderma pigmentosum variant) gene encodes human DNA polymerase eta. Nature, 1999. 399(6737): p. 700-704. Zhivotovsky, B. and Kroemer, G., Apoptosis and genomic instability. Nat Rev Mol Cell Biol, 2004. 5(9): p. 752-762. Michor, F., Chromosomal instability and human cancer. Philos Trans R Soc Lond B Biol Sci, 2005. 360(1455): p. 631-635. 110 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. Hanahan, D. and Folkman, J., Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis. Cell, 1996. 86(3): p. 353-364. Hanahan, D., Christofori, G., Naik, P. and Arbeit, J., Transgenic mouse models of tumour angiogenesis: the angiogenic switch, its molecular controls, and prospects for preclinical therapeutic models. Eur J Cancer, 1996. 32A(14): p. 2386-2393. Dameron, K.M., Volpert, O.V., Tainsky, MA. and Bouck, N., Control of angiogenesis in fibroblasts by p53 regulation of thrombospondin-1. Science, 1994. 265(5178): p. 1582-1584. Volpert, O.V., Dameron, KM. and Bouck, N., Sequential development of an angiogenic phenotype by human fibroblasts progressing to tumorigenicity. Oncogene, 1997. 14(12): p. 1495-1502. Webb, CF. and Vande Woude, G.F., Genes that regulate metastasis and angiogenesis. J Neurooncol, 2000. 50(1-2): p. 71-87. Beck, L., Jr. and D'Amore, P.A., Vascular development: cellular and molecular regulation. Faseb J, 1997. 11(5): p. 365-373. Veikkola, T. and Alitalo, K., VEGFs, receptors and angiogenesis. Semin Cancer Biol, 1999. 9(3): p. 211-220. Rundhaug, J.E., Matrix metalloproteinases and angiogenesis. J Cell Mol Med, 2005. 9(2): p. 267-285. Hicklin, D.J. and Ellis, L.M., Role of the vascular endothelial growth factor pathway in tumor growth and angiogenesis. J Clin Oncol, 2005. 23(5): p. 1011-1027. Risau, W., Angiogenesis and endothelial cell function. Arzneimittelforschung, 1994. 44(3A): p. 416-417. Dvorak, H.F., VPF/VEGF and the angiogenic response. Semin Perinatol, 2000. 24(1): p. 75-78. McMahon, G., VEGF receptor signaling in tumor angiogenesis. Oncologist, 2000. 5 Suppl 1: p. 3-10. Norrby, K., Vascular endothelial growth factor and de novo mammalian angiogenesis. Microvasc Res, 1996. 51(2): p. 153-163. Dery, M.A., Michaud, MD. and Richard, D.E., Hypoxia-inducible factor 1: regulation by hypoxic and non-hypoxic activators. Int J Biochem Cell Biol, 2005. 37(3): p. 535-540. 111 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. Maynard, MA. and Ohh, M., Von Hippel-Lindau tumor suppressor protein and hypoxia-inducible factor in kidney cancer. Am J Nephrol, 2004. 24(1): p. 1-13. Kliche, S. and Waltenberger, J., VEGF receptor signaling and endothelial function. lUBMB Life, 2001. 52(1-2): p. 61-66. Matsumoto, T. and Claesson-Welsh, L., VEGF receptor signal transduction. Sci STKE, 2001. 2001(112): p. RE21. Cairns, R.A., Khokha, R. and Hill, R.P., Molecular mechanisms of tumor invasion and metastasis: an integrated view. Curr Mol Med, 2003. 3(7): p. 659-671. Hood, JD. and Cheresh, D.A., Role of integrins in cell invasion and migration. Nat Rev Cancer, 2002. 2(2): p. 91-100. Bracke, M.E., Van Roy, PM. and Mareel, M.M., The E-cadherin/catenin complex in invasion and metastasis. Curr Top Microbiol Immunol, 1996. 213 (PH): p. 123-161. Ziober, B.L., Lin, CS. and Kramer, R.H., Laminin-binding integrins in tumor progression and metastasis. Semin Cancer Biol, 1996. 7(3): p. 119-128. Rudek, M.A., Venitz, J. and Figg. W.D., Matrix metalloproteinase inhibitors: do they have a place in anticancer therapy? Pharmacotherapy. 2002. 22(6): p. 705-720. Bissell, M.J. and Radisky, D., Putting tumours in context. Nat Rev Cancer, 2001. 1(1): p. 46-54. Mareel, M.M., Van Roy, PM. and Bracke, M.E., How and when do tumor cells metastasize? Crit Rev Oncog, 1993. 4(5): p. 559-594. Ruoslahti, E., Specialization of tumour vasculature. Nat Rev Cancer, 2002. 2(2): p. 83-90. Muller, A., Homey, 3., Soto, H., Ge, N., Catron, D., Buchanan, M.E., McClanahan, T., Murphy, E., Yuan, W., Wagner, S.N., Barrera, J.L., Mohar, A., Verastegui, E. and Zlotnik, A., Involvement of chemokine receptors in breast cancer metastasis. Nature, 2001. 410(6824): p. 50-56. McCormick, J.J. and Maher, V.M., Towards an understanding of the malignant transformation of diploid human fibroblasts. Mutat Res, 1988. 199(2): p. 273-291. 112 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. McCormick, J.J. and Maher, V.M., Malignant transformation of mammalian cells in culture, including human cells. Environ Mol Mutagen, 1989. 14 Suppl 16: p. 105-113. McCormick, J.J. and Maher, V.M., Analysis of the multistep process of carcinogenesis using human fibroblasts. Risk Anal, 1994. 14(3): p. 257- 263. Ryan, P.A., Maher, V.M. and McCormick, J.J., Failure of infinite life span human cells from different immortality complementation groups to yield finite life span hybrids. J Cell Physiol, 1994. 159(1): p. 151-160. McCormick, J.J., Fry, D.G., Hurlin, P.J., Morgan, T.L., Wilson, D.M. and Maher, V.M., Malignant transformation of human fibroblasts by oncogene transfection or carcinogen treatment. Prog Clin Biol Res, 1990. 3400: p. 195-205. Morgan, T.L., Yang, D.J., Fry, D.G., Hurlin, P.J., Kohler, S.K., Maher, V.M. and McCormick, J.J., Characteristics of an infinite life span diploid human fibroblast cell strain and a near-diploid strain arising from a clone of cells expressing a transfected v-myc oncogene. Exp Cell Res, 1991. 197(1): p. 125-136. Hahn, WC. and Weinberg, R.A., Modelling the molecular circuitry of cancer. Nat Rev Cancer, 2002. 2(5): p. 331-341. Cerni, C., Telomeres, telomerase, and myc. An update. Mutat Res, 2000. 462(1): p. 31-47. Gil, J., Kerai, P., Lleonart, M., Bernard, 0., Cigudosa, J.C., Peters, G., Camero, A. and Beach, 0., lmmortalization of primary human prostate epithelial cells by c-Myc. Cancer Res, 2005. 65(6): p. 2179-2185. Kim, H.S., Shin, J.Y., Yun, J.Y., Ahn, UK and Le, J.Y., lmmortalization of human embryonic fibroblasts by overexpression of c-myc and simian virus 40 large Tantigen. Exp Mol Med, 2001. 33(4): p. 293-298. Wu, K.J., Grandori, C., Amacker, M., Simon-Vermot, N., Polack, A., Lingner, J. and Dalla-Favera, R., Direct activation of TERT transcription by c-MYC. Nat Genet, 1999. 21(2): p. 220-224. Kyo, S., Takakura, M., Taira, T., Kanaya, T., Itoh, H., Yutsudo, M., Ariga, H. and lnoue, M., Sp1 cooperates with c-Myc to activate transcription of the human telomerase reverse transcriptase gene (h TERT). Nucleic Acids Res, 2000. 28(3): p. 669-677. 113 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 181. 182. Park, N.H., Guo, W., Kim, H.R., Kang, MK. and Park, N.H., c-Myc and Sp 1/3 are required for transactivation of hamster telomerase catalytic subunit gene promoter. Int J Oncol, 2001. 19(4): p. 755-761. Yang, 0., Kohler, S.K., Maher, V.M. and McCormick, J.J., v-sis oncogene- induced transformation of human fibroblasts into cells capable of forming benign tumors. Carcinogenesis, 1994. 15(10): p. 2167-2175. Wilson, D.M., Yang, D.J., Dillberger, J.E., Dietrich, S.E., Maher, V.M. and McCormick, J.J., Malignant transformation of human fibroblasts by a transfected N-ras oncogene. Cancer Res, 1990. 50(17): p. 5587-5593. Hurlin, P.J., Maher, V.M. and McCormick, J.J., Malignant transformation of human fibroblasts caused by expression of a transfected T24 HRAS oncogene. Proc Natl Acad Sci U S A, 1989. 86(1): p. 187-191. Lin, C., Maher, V.M. and McCormick, J.J., Malignant transformation of human fibroblast strain MSU-1.1 by v-fes requires an additional genetic Change. Int J Cancer, 1995. 63(1): p. 140-147. Fry, D.G., Milam, L.D., Dillberger, J.E., Maher, V.M. and McCormick, J.J., Malignant transformation of an infinite life span human fibroblast cell strain by transfection with v-Ki-ras. Oncogene, 1990. 5(9): p. 1415-1418. Lin, C., Wang, Q., Maher, V.M. and McCormick, J.J., Malignant transformation of a human fibroblast cell strain by transfection of a v-fes oncogene but not by transfection of a gag-human c-fes construct. Cell Growth Differ, 1994. 5(12): p. 1381-1387. Benito, M. and Lorenzo, M., Platelet derived growth factor/tyrosine kinase receptor mediated proliferation. Growth Regul, 1993. 3(3): p. 172-179. Stice, L.L., Vaziri, C. and Faller, D.V., Regulation of platelet-derived growth factor signaling by activated p21Ras. Front Biosci, 1999. 4: p. 072-86. Dikic, l. and Blaukat, A., Protein tyrosine kinase-mediated pathways in G protein-coupled receptor signaling. Cell Biochem Biophys, 1999. 30(3): p. 369-387. Lennartsson, J., Blume-Jensen, P., Herrnanson, M., Ponten, E., Carlberg, M. and Ronnstrand, L., Phosphorylation of Shc by Src family kinases is necessary for stem cell factor receptor/c-kit mediated activation of the Ras/MAP kinase pathway and c-fos induction. Oncogene, 1999. 18(40): p. 5546-5553. Li, J. and Smithgall, T.E., Co-expression with BCR induces activation of the FES tyrosine kinase and phosphorylation of specific N-terminal BCR tyrosine residues. J Biol Chem, 1996. 271(51): p. 32930-32936. 114 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. Boley, S.E., McManus, T.P., Maher, V.M. and McCormick, J.J., Malignant transformation of human fibroblast cell strain MSU-1.1 by N-methyl-N- nitrosourea: evidence of elimination of p53 by homologous recombination. Cancer Res, 2000. 60(15): p. 4105-4111. O'Reilly, 8., Walicka, M., Kohler, S.K., Dunstan, R., Maher, V.M. and McCormick, J.J., Dose-dependent transformation of cells of human fibroblast cell strain MSU-1.1 by cobalt-60 gamma radiation and characterization of the transformed cells. Radiat Res, 1998. 150(5): p. 577- 584. Yang, 0., Louden, C., Reinhold, D.S., Kohler, S.K., Maher, V.M. and McCormick, J.J., Malignant transformation of human fibroblast cell strain MSU-1.1 by (+-)-7 beta,8 alpha-dihydroxy-Q alpha,10 alpha-epoxy- 7,8,9,10-tetrahydrobenzo [a]pyrene. Proc Natl Acad Sci U S A, 1992. 89(6): p. 2237-2241. Lou, Z., O'Reilly, S., Liang, H., Maher, V.M., Sleight, SD. and McCormick, J.J., Down-regulation of overexpressed sp1 protein in human fibrosarcoma cell lines inhibits tumor formation. Cancer Res, 2005. 65(3): p. 1007-1017. Liang, H., O'Reilly, 8., Liu, Y., Abounader, R., Laterra, J., Maher, V.M. and McCormick, J.J., Sp1 regulates expression of MET, and ribozyme-induced down-regulation of MET in fibrosarcoma-derived human cells reduces or eliminates their tumorigenicity. Int J Oncol, 2004. 24(5): p. 1057-1067. Qing, J., Maher, V.M., Tran, H., Argraves, W.S., Dunstan, R.W. and McCormick, J.J., Suppression of anchorage-independent growth and matrigel invasion and delayed tumor formation by elevated expression of fibulin-1D in human fibrosarcoma-derived cell lines. Oncogene, 1997. 15(18): p. 2159-2168. Colicelli, J., Human RAS superfamily proteins and related GTPases. Sci STKE, 2004. 2004(250): p. RE13. Rodenhuis, S., ras and human tumors. Semin Cancer Biol, 1992. 3(4): p. 241-247. Malumbres, M. and Pellicer, A., RAS pathways to cell cycle control and cell transformation. Front Biosci, 1998. 3: p. d887-912. Harvey, J.J., An Unidentified Virus Which Causes The Rapid Production Of Tumours In Mice. Nature, 1964. 204: p. 1104-1105. Hall, A., Marshall, C.J., Spurr, MK and Weiss, R.A., Identification of transforming gene in two human sarcoma cell lines as a new member of the ras gene family located on chromosome 1. Nature, 1983. 303(5916): p. 396-400. 115 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. Der, C.J., Krontiris, TC. and Cooper, G.M., Transforming genes of human bladder and lung carcinoma cell lines are homologous to the ras genes of Harvey and Kirsten sarcoma viruses. Proc Natl Acad Sci U S A, 1982. 79(1 1 ): p. 3637-3640. Parada, L.F. and Weinberg, R.A., Presence of a Kirsten murine sarcoma virus ras oncogene in cells transformed by 3-methylcholanthrene. Mol Cell Biol, 1983. 3(12): p. 2298-2301. Scolnick, E.M., Rands, E., Williams, D. and Parks, W.P., Studies on the nucleic acid sequences of Kirsten sarcoma virus: a model for formation of a mammalian RNA-containing sarcoma virus. J Virol, 1973. 12(3): p. 458- 463. Santos, E., Tronick, S.R., Aaronson, S.A., Pulciani, S. and Barbacid, M., T24 human bladder carcinoma oncogene is an activated form of the normal human homologue of BALB- and Harvey-MSV transforming genes. Nature, 1982. 298(5872): p. 343-347. Reddy, E.P., Reynolds, R.K., Santos, E. and Barbacid, M., A point mutation is responsible for the acquisition of transforming properties by the T24 human bladder carcinoma oncogene. Nature, 1982. 300(5888): p. 149-152. Tabin, C.J., Bradley, S.M., Bargmann, C.|., Weinberg, R.A., Papageorge, A.G., Scolnick, E.M., Dhar, R., Lowy, DR. and Chang, E.H., Mechanism of activation ofa human oncogene. Nature, 1982. 300(5888): p. 143-149. Taparowsky, E., Suard, Y., Fasano, 0., Shimizu, K., Goldfarb, M. and Wigler, M., Activation of the T24 bladder carcinoma transforming gene is linked to a single amino acid change. Nature, 1982. 300(5894): p. 762-765. Hurley, J.B., Simon, M.|., Teplow, D.B., Robishaw, JD. and Gilman, A.G., Homologies between signal transducing G proteins and ras gene products. Science, 1984. 226(4676): p. 860-862. McCormick, F., Coupling of ras p21 signalling and GTP hydrolysis by GTPase activating proteins. Philos Trans R Soc Lond B Biol Sci, 1992. 336( 1276): p. 43-47; discussion 47-48. Boume, H.R., Sanders, DA. and McCormick, F., The GTPase superfamily: conserved structure and molecular mechanism. Nature, 1991. 349(6305): p. 117-127. Perez-Sela, D. and Rebollo, A., Novel aspects of Ras proteins biology: regulation and implications. Cell Death Differ, 1999. 6(8): p. 722-728. 116 205. 206. 207. 208. 209. 210. 211. 212. 213. 214. 215. 216. Milbum, M.V., Tong, L., deVos, A.M., Brunger, A., Yamaizumi, Z., Nishimura, S. and Kim, S.H., Molecular switch for signal transduction: structural differences between active and inactive forms of protooncogenic ras proteins. Science, 1990. 247(4945): p. 939-945. Herrrnann, C., Horn, G., Spaargaren, M. and Wittinghofer, A., Differential interaction of the ras family GTP-binding proteins H-Ras, Rap 1A, and R- Ras with the putative effector molecules Raf kinase and Ral-guanine nucleotide exchange factor. J Biol Chem, 1996. 271(12): p. 6794-6800. Neal, S.E., Eccleston, J.F., Hall, A. and Webb, M.R., Kinetic analysis of the hydrolysis of GTP by p21N-ras. The basal GTPase mechanism. J Biol Chem, 1988. 263(36): p. 19718-19722. Quilliam, L.A., Rebhun, J.F. and Castro, A.F., A growing family of guanine nucleotide exchange factors is responsible for activation of Ras-family GTPases. Prog Nucleic Acid Res Mol Biol, 2002. 71: p. 391-444. Feig, LA. and Cooper, G.M., Inhibition of NIH 3T3 cell proliferation by a mutant ras protein with preferential affinity for GDP. Mol Cell Biol, 1988. 8(8): p. 3235-3243. Chen, S.Y., Huff, S.Y., Lai, C.C., Der, OJ. and Powers, 8., Ras-15A protein shares highly similar dominant-negative biological properties with Ras-17N and forms a stable, guanine-nucleotide resistant complex with CDCZ5 exchange factor. Oncogene, 1994. 9(9): p. 2691-2698. Boguski, MS. and McCormick, F., Proteins regulating Ras and its relatives. Nature, 1993. 366(6456): p. 643-654. Donovan, 8., Shannon, KM. and Bollag, G., GTPase activating proteins: critical regulators of intracellular signaling. Biochim Biophys Acta, 2002. 1602(1): p. 23-45. Scheffzek, K., Ahmadian, MR. and Wittinghofer, A., GTPase-activating proteins: helping hands to complement an active site. Trends Biochem Sci, 1998. 23(7): p. 257-262. Bollag, G. and McCormick, F., Regulators and effectors of ras proteins. Annu Rev Cell Biol, 1991. 7: p. 601-632. Nimnual, A. and Bar-Sagi, D., The two hats of SOS. Sci STKE, 2002. 2002(145): p. PE36. Buday, L. and Downward, J., Epidermal growth factor regulates p21ras through the formation of a complex of receptor, Grb2 adapter protein, and 803 nucleotide exchange factor. Cell, 1993. 73(3): p. 611-620. 117 217. 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. Cadena, D.L. and Gill, G.N., Receptor tyrosine kinases. Faseb J, 1992. 6(6): p. 2332-2337. Ceresa, BR and Pessin, J.E., Insulin regulation of the Ras activation/inactivation cycle. Mol Cell Biochem, 1998. 182(1-2): p. 23-29. Crespo, P., Xu, N., Simonds, W.F. and Gutkind, J.S., Ras-dependent activation of MAP kinase pathway mediated by G-protein beta gamma subunits. Nature, 1994. 369(6479): p. 418-420. Luttrell, L.M., Hawes, B.E., van Biesen, T., Luttrell, D.K., Lansing, T.J. and Lefkowitz, R.J., Role of c-Src tyrosine kinase in G protein-coupled receptor- and Gbetagamma subunit-mediated activation of mitogen- activated protein kinases. J Biol Chem, 1996. 271(32): p. 19443-19450. Douville, E. and Downward, J., EGF induced SOS phosphorylation in PC12 cells involves P90 RSK-2. Oncogene, 1997. 15(4): p. 373-383. Jefferson, A. B., Klippel, A. and Williams, L. T. Inhibition of mSOS-activity by binding of phosphatidylinositol 4, 5-P2 to the mSOS pleckstrin homology domain. Oncogene, 1998. 16(18): p. 2303- 2310. Cullen, P.J. and Lockyer, P.J., Integration of calcium and Ras signalling. Nat Rev Mol Cell Biol, 2002. 3(5): p. 339-348. Famsworth, C.L., Freshney, N.W., Rosen, L.B., Ghosh, A., Greenberg, ME. and Feig, L.A., Calcium activation of Ras mediated by neuronal exchange factor Ras-GRF. Nature, 1995. 376(6540): p. 524-527. Gideon, P., John, J., Frech, M., Lautwein, A., Clark, R., Scheffler, J.E. and Wittinghofer, A., Mutational and kinetic analyses of the GTPase-activating protein (GAP)-p21 interaction: the C-terminal domain of GAP is not sufficient for full activity. Mol Cell Biol, 1992. 12(5): p. 2050-2056. Willumsen, B.M., Christensen, A., Hubbert, N.L., Papageorge, AG. and Lowy, D.R., The p21 ras C-terminus is required for transformation and membrane association. Nature, 1984. 310(5978): p. 583-586. Niv, H., Gutman, O., Kloog, Y. and Henis, Y.|., Activated K-Ras and H-Ras display different interactions with saturable nonraft sites at the surface of live cells. J Cell Biol, 2002. 157(5): p. 865-872. Prior, l.A., Muncke, C., Patton, RC. and Hancock, J.F., Direct visualization of Ras proteins in spatially distinct cell surface microdomains. J Cell Biol, 2003. 160(2): p. 165-170. 118 229. 230. 231. 232. 233. 234. 235. 236. 237. 238. 239. 240. 241. Reiss, Y., Goldstein, J.L., Seabra, M.C., Casey, P.J. and Brown, M.S., Inhibition of purified p21ras farnesylsprotein transferase by Cys-AAX tetrapeptides. Cell, 1990. 62(1): p. 81-88. Boyartchuk, V.L., Ashby, MN. and Rine, J., Modulation of Ras and a-factor function by carboxyl-terminal proteolysis. Science, 1997. 275(5307): p. 1796-1800. Dai, Q., Choy, E., Chiu, V., Romano, J., Slivka, S.R., Steitz, S.A., Michaelis, S. and Philips, M.R., Mammalian prenylcysteine carboxyl methyltransferase is in the endoplasmic reticulum. J Biol Chem, 1998. 273(24): p. 15030-15034. Apolloni, A., Prior, l.A., Lindsay, M., Parton, R.G. and Hancock, J.F., H-ras but not K-ras traffics to the plasma membrane through the exocytic pathway. Mol Cell Biol, 2000. 20(7): p. 2475-2487. Repasky, G.A., Chenette, E.J. and Der, C.J., Renewing the conspiracy theory debate: does Raf function alone to mediate Ras oncogenesis? Trends Cell Biol, 2004. 14(11): p. 639-647. Herrmann, C., Ras-effector interactions: after one decade. Curr Opin Struct Biol, 2003. 13(1): p. 122-129. Wellbrock, C., Karasarides, M. and Marais, R., The RAF proteins take centre stage. Nat Rev Mol Cell Biol, 2004. 5(11): p. 875-885. Wittinghofer, A. and Nassar, N., How Ras-related proteins talk to their effectors. Trends Biochem Sci, 1996. 21(12): p. 488-491. Drugan, J.K., Khosravi-Far, R., White, M.A., Der, C.J., Sung, Y.J., Hwang, Y.W. and Campbell, S.L., Ras interaction with two distinct binding domains in Raf-1 may be required for Ras transformation. J Biol Chem, 1996. 271(1): p. 233-237. Stokoe, D., Macdonald, S.G., Cadwallader, K., Symons, M. and Hancock, J.F., Activation of Raf as a result of recruitment to the plasma membrane. Science, 1994. 264(5164): p. 1463-1467. Morrison, OK. and Cutler, R.E., The complexity of Raf-1 regulation. Curr Opin Cell Biol, 1997. 9(2): p. 174-179. Johnson, G.L. and Lapadat, R., Mitogen-activated protein kinase pathways mediated by ERK, JNK, and p38 protein kinases. Science, 2002. 298(5600): p. 1911-1912. Hilger, R.A., Scheulen, ME. and Strumberg, D., The Ras-Raf-MEK-ERK pathway in the treatment of cancer. Onkologie, 2002. 25(6): p. 51 1-518. 119 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. 252. 253. 254. Hazzalin, CA. and Mahadevan, L.C., MAPK-regulated transcription: a continuously variable gene switch? Nat Rev Mol Cell Biol, 2002. 3(1): p. 30-40. Carpenter, CL. and Cantley, L.C., Phosphoinositide 3-kinase and the regulation of cell growth. Biochim Biophys Acta, 1996. 1288(1): p. M11-16. Franke, T.F., Kaplan, DR. and Cantley, L.C., PI3K: downstream AKTion blocks apoptosis. Cell, 1997. 88(4): p. 435-437. Okkenhaug, K. and Vanhaesebroeck, B., New responsibilities for the PI3K regulatory subunit p85 alpha. Sci STKE, 2001. 2001(65): p. PE1. Otsu, M., Hiles, l., Gout, |., Fry, M.J., Ruiz-Larrea, F., Panayotou, G., Thompson, A., Dhand, R., Hsuan, J., Totty, N. and et al., Characterization of two 85 kd proteins that associate with receptor tyrosine kinases, middle- T/pp60c-src complexes, and Pl3-kinase. Cell, 1991. 65(1): p. 91-104. Rodriguez-Viciana, P., Wame, P.H., Vanhaesebroeck, B., Waterfield, MD. and Downward, J., Activation of phosphoinositide 3-kinase by interaction with Ras and by point mutation. Embo J, 1996. 15(10): p. 2442-2451. Maffucci, T. and Falasca, M., Specificity in pleckstrin homology (PH) domain membrane targeting: a role for a phosphoinositide-protein co- operative mechanism. FEBS Lett, 2001. 506(3): p. 173-179. Gillooly, D.J., Simonsen, A. and Stenmark, H., Cellular functions of phosphatidylinositol 3-phosphate and FYVE domain proteins. Biochem J, 2001. 355(Pt 2): p. 249-258. Alessi, D.R., Andjelkovic, M., Caudwell, B., Cron, P., Morrice, N., Cohen, P. and Hemmings, B.A., Mechanism of activation of protein kinase B by insulin and IGF-1. Embo J, 1996. 15(23): p. 6541-6551. Datta, S.R., Brunet, A. and Greenberg, M.E., Cellular survival: a play in three Akts. Genes Dev, 1999. 13(22): p. 2905-2927. Datta, S.R., Dudek, H., Tao, X., Masters, 8., Fu, H., Gotoh, Y. and Greenberg, M.E., Akt phosphorylation of BAD coup/es survival signals to the cell-intrinsic death machinery. Cell, 1997. 91(2): p. 231-241. Kane, L.P., Shapiro, V.S., Stokoe, D. and Weiss, A., Induction of NF- kappaB by the Akt/PKB kinase. Curr Biol, 1999. 9(11): p. 601-604. Burgering, 8M. and Medema, R.H., Decisions on life and death: FOXO Forkhead transcription factors are in command when PKB/Akt is off duty. J Leukoc Biol, 2003. 73(6): p. 689-701. 120 255. 256. 257. 258. 259. 260. 261. 262. 263. 264. 265. 266. 267. Mayo, L.D., Dixon, J.E., Durden, D.L., Tonks, MK. and Donner, D.B., PTEN protects p53 from Mdm2 and sensitizes cancer cells to chemotherapy. J Biol Chem, 2002. 277(7): p. 5484-5489. Mayo, LD. and Donner, D.B., A phosphatidylinositol 3-kinase/Akt pathway promotes translocation of Mdm2 from the cytoplasm to the nucleus. Proc Natl Acad Sci U S A, 2001. 98(20): p. 11598-11603. Doble, SW. and Woodgett, J.R., GSK-3: tricks of the trade for a multi- tasking kinase. J Cell Sci, 2003. 116(Pt 7): p. 1175-1186. Cross, D.A., Alessi, D.R., Cohen, R, Andjelkovich, M. and Hemmings, B.A., Inhibition of glycogen synthase kinase-3 by insulin mediated by protein kinase 8. Nature, 1995. 378(6559): p. 785-789. Bar-Sagi, D. and Hall, A., Rae and Rho GTPases: a family reunion. Cell, 2000. 103(2): p. 227-238. Lambert, J.M., Lambert, Q.T., Reuther, G.W., Malliri, A., Siderovski, D.P., Sondek, J., Collard, J.G. and Der, C.J., Tiam1 mediates Ras activation of Rac by a Pl(3)K-independent mechanism. Nat Cell Biol, 2002. 4(8): p. 621- 625. Welch, H.C., Coadwell, W.J., Stephens, LR. and Hawkins, P.T., Phosphoinositide 3-kinase-dependent activation of Rac. FEBS Lett, 2003. 546(1): p. 93-97. Michiels, F., Habets, G.G., Stam, J.C., van der Kammen, RA. and Collard, J.G., A role for Rac in Tiam1-induced membrane ruffling and invasion. Nature, 1995. 375(6529): p. 338-340. Fleming, |.N., Elliott, C.M., Buchanan, F.G., Downes, GP. and Exton, J.H., Ca2+/calmodulin-dependent protein kinase II regulates Tiam1 by reversible protein phosphorylation. J Biol Chem, 1999. 274(18): p. 12753- 12758. Mertens, A.E., Roovers, RC. and Collard, J.G., Regulation of Tiam1-Rac signalling. FEBS Lett, 2003. 546(1): p. 11-16. Feig, L.A., Urano, T. and Cantor, 8., Evidence for a Ras/Ral signaling cascade. Trends Biochem Sci, 1996. 21(11): p. 438-441. Geyer, M., Herrmann, C., Wohlgemuth, S., Wittinghofer, A. and Kalbitzer, H.R., Structure of the Ras-binding domain of RalGEF and implications for Ras binding and signalling. Nat Struct Biol, 1997. 4(9): p. 694-699. Wolthuis, RM. and 303, J.L., Ras caught in another affair: the exchange factors for Ral. Curr Opin Genet Dev, 1999. 9(1): p. 112-117. 121 268. 269. 270. 271. 272. 273. 274. 275. 276. 277. 278. Voss, M., Weemink, P.A., Haupenthal, S., Moller, U., Cool, R.H., Bauer, 8., Camonis, J.H., Jakobs, K.H. and Schmidt, M., Phospholipase D stimulation by receptor tyrosine kinases mediated by protein kinase C and a Ras/Ral signaling cascade. J Biol Chem, 1999. 274(49): p. 34691-34698. Cantor, S.B., Urano, T. and Feig, L.A., Identification and characterization of Ral-binding protein 1, a potential downstream target of Ral GTPases. Mol Cell Biol, 1995. 15(8): p. 4578-4584. Jullien-Flores, V., Dorseuil, O., Romero, F., Letourneur, F., Saragosti, S., Berger, R., Tavitian, A., Gacon, G. and Camonis, J.H., Bridging Ral GTPase to Rho pathways. RLIP76, a Ral effector with CDC42/Rac GTPase-activating protein activity. J Biol Chem, 1995. 270(38): p. 22473- 22477. Russell, M., Lange-Carter, CA. and Johnson, G.L., Direct interaction between Ras and the kinase domain of mitogen-activated protein kinase kinase kinase (MEKK1). J Biol Chem, 1995. 270(20): p. 11757-11760. Uhlik, M.T., Abell, A.N., Cuevas, B.D., Nakamura, K. and Johnson, G.L., Wiring diagrams of MAPK regulation by MEKK 1, 2, and 3. Biochem Cell Biol, 2004. 82(6): p. 658-663. Tall, G.G., Barbieri, M.A., Stahl, PD. and Horazdovsky, B.F., Ras- activated endocytosis is mediated by the Rab5 guanine nucleotide exchange activity of RIN1. Dev Cell, 2001. 1(1): p. 73-82. Barbieri, M.A., Kong, C., Chen, P.l., Horazdovsky, BF. and Stahl, P.D., The SRC homology 2 domain of Rin1 mediates its binding to the epidermal growth factor receptor and regulates receptor endocytosis. J Biol Chem, 2003. 278(34): p. 32027-32036. Hu, H., Bliss, J.M., Wang, Y. and Colicelli, J., RIN1 is an ABL tyrosine kinase activator and a regulator of epithelial-cell adhesion and migration. Curr Biol, 2005. 15(9): p. 815-823. Kuriyama, M., Harada, N., Kuroda, S., Yamamoto, T., Nakafuku, M., Iwamatsu, A., Yamamoto, D., Prasad, R., Croce, C., Canaani, E. and Kaibuchi, K., Identification of AF-6 and canoe as putative targets for Ras. J Biol Chem, 1996. 271(2): p. 607-610. Radziwill, G., Erdmann, R.A., Margelisch, U. and Moelling, K., The Bcr kinase downregulates Ras signaling by phosphorylating AF-6 and binding to its PDZ domain. Mol Cell Biol, 2003. 23(13): p. 4663-4672. Boettner, B., Harjes, P., lshimaru, S., Heke, M., Fan, H.Q., Qin, Y., Van Aelst, L. and Gaul, U., The AF-6 homolog canoe acts as a Rap1 effector 122 279. 280. 281. 282. 283. 284. 285. 286. 287. 288. during dorsal closure of the Drosophila embryo. Genetics, 2003. 165(1): p. 1 59-169. Su, L., Hattori, M., Moriyama, M., Murata, N., Harazaki, M., Kaibuchi, K. and Minato, N., AF-6 controls integrin-mediated cell adhesion by regulating Rap1 activation through the specific recruitment of Rap1GTP and SPA-1. J Biol Chem, 2003. 278(17): p. 15232-15238. Aoyama, Y., Avruch, J. and Zhang, X.F., Nore1 inhibits tumor cell growth independent of Ras or the MST1/2 kinases. Oncogene, 2004. 23(19): p. 3426-3433. Hesson, L., Bieche, |., Krex, D., Criniere, E., Hoang-Xuan, K., Maher, ER. and Latif, F., Frequent epigenetic inactivation of RASSF 1A and BLU genes located within the critical 3p21.3 region in gliomas. Oncogene, 2004. 23(13): p. 2408-2419. Li, Z., Zhang, H., McManus, T.P., McCormick, J.J., Lawrence, CW. and Maher, V.M., hRE V3 is essential for error-prone translesion synthesis past UV or benzo[a]pyrene diol epoxide-induced DNA lesions in human fibroblasts. Mutat Res, 2002. 510(1-2): p. 71-80. Khokhlatchev, A., Rabizadeh, 8., Xavier, R., Nedwidek, M., Chen, T., Zhang, X.F., Seed, B. and Avruch, J., Identification of a novel Ras- regulated proapoptotic pathway. Curr Biol, 2002. 12(4): p. 253-265. Mulcahy, L.S., Smith, MR. and Stacey, D.W., Requirement for ras proto- oncogene function during serum-stimulated growth of NIH 3T3 cells. Nature, 1985. 313(5999): p. 241-243. Peeper, D.S., Upton, T.M., Ladha, M.H., Neuman, E., Zalvide, J., Bemards, R., DeCaprio, J.A. and Ewen, M.E., Ras signalling linked to the cell-cycle machinery by the retinoblastoma protein. Nature, 1997. 386(6621): p. 177-181. Filmus, J., Robles, A.l., Shi, W., Wong, M.J., Colombo, LL. and Conti, C.J., Induction of cyclin D1 overexpression by activated ras. Oncogene, 1994. 9(12): p. 3627-3633. Westwick, J.K., Lambert, Q.T., Clark, G.J., Symons, M., Van Aelst, L., Pestell, R.G. and Der, C.J., Rac regulation of transformation, gene expression, and actin organization by multiple, PAK-independent pathways. Mol Cell Biol, 1997. 17(3): p. 1324-1335. Muise-Helmericks, R.C., Grimes, H.L., Bellacosa, A., Malstrom, S.E., Tsichlis, RN. and Rosen, N., Cyclin D expression is controlled post- transcriptionally via a phosphatidylinositol 3-kinase/Akt-dependent pathway. J Biol Chem, 1998. 273(45): p. 29864-29872. 123 289. 290. 291. 292. 293. 294. 295. 296. 297. 298. Diehl, J.A., Cheng, M., Roussel, M.F. and Sherr, C.J., Glycogen synthase kinase-3beta regulates cyclin D1 proteolysis and subcel/ular localization. Genes Dev, 1998. 12(22): p. 3499-3511. Joyce, D., Bouzahzah, B., Fu, M., Albanese, C., D'Amico, M., Steer, J., Klein, J.U., Lee, R.J., Segall, J.E., Westwick, J.K., Der, OJ. and Pestell, R.G., Integration of Rac-dependent regulation of cyclin D1 transcription through a nuclear factor-kappaB-dependent pathway. J Biol Chem, 1999. 274(36): p. 25245-25249. Delmas, C., Manenti, S., Boudjelal, A., Peyssonnaux, C., Eychene, A. and Darbon, J.M., The p42/p44 mitogen-activated protein kinase activation triggers p27Kip1 degradation independently of CDK2/cyc/in E in NIH 3T3 cells. J Biol Chem, 2001. 276(37): p. 34958-34965. Kops, G.J., de Ruiter, N.D., De Vries-Smits, A.M., Powell, D.R., Bos, J.L. and Burgering, B.M., Direct control of the Forkhead transcription factor AFX by protein kinase 8. Nature, 1999. 398(6728): p. 630-634. Weber, J.D., Hu, W., Jefcoat, S.G., Jr., Raben, D.M. and Baldassare, J.J., Ras-stimulated extracellular signal-related kinase 1 and RhoA activities coordinate platelet-derived growth factor-induced G1 progression through the independent regulation of cyclin D1 and p27. J Biol Chem, 1997. 272(52): p. 32966-32971. Serrano, M., Lin, A.W., McCurrach, M.E., Beach, D. and Lowe, S.W., Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell, 1997. 88(5): p. 593-602. Woods, D., Pan'y, D., Chervvinski, H., Bosch, E., Lees, E. and McMahon, M., Raf-induced proliferation or cell cycle arrest is determined by the level of Raf activity with arrest mediated by p21Cip1. Mol Cell Biol, 1997. 17(9): p. 5598-5611. Aoki, M., Blazek, E. and Vogt, P.K., A role of the kinase mTOR in cellular transformation induced by the oncoproteins P3k and Akt. Proc Natl Acad Sci U S A, 2001. 98(1): p. 136-141. Aspuria, P.J. and Tamanoi, F., The Rheb family of GTP-binding proteins. Cell Signal, 2004. 16(10): p. 1105-1112. Garami, A., Zwartkruis, F.J., Nobukuni, T., Joaquin, M., Roccio, M., Stocker, H., Kozma, S.C., Hafen, E., Bos, J.L. and Thomas, G., Insulin activation of Rheb, a mediator of mTOR/S6K/4E-BP signaling, is inhibited by T801 and 2. Mol Cell, 2003. 11(6): p. 1457-1466. 124 299. 300. 301. 302. 303. 304. 305. 306. 307. 308. 309. Fingar, D.C., Salama, S., Tsou, C., Harlow, E. and Blenis, J., Mammalian cell size is controlled by mTOR and its downstream targets S6K1 and 4EBP1/elF4E. Genes Dev, 2002. 16(12): p. 1472-1487. Tee, AR. and Blenis, J., mTOR, translational control and human disease. Semin Cell Dev Biol, 2005. 16(1): p. 29-37. Sen, P., Mukherjee, 8., Ray, D. and Raha, S., Involvement of the Akt/PKB signaling pathway with disease processes. Mol Cell Biochem, 2003. 253(1- 2): p. 241-246. Chang, F., Lee, J.T., Navolanic, P.M., Steelman, L.S., Shelton, J.G., Blalock, W.L., Franklin, RA. and McCubrey, J.A., Involvement of Pl3K/Akt pathway in cell cycle progression, apoptosis, and neoplastic transformation: a target for cancer chemotherapy. Leukemia, 2003. 17(3): p. 590-603. Sulciner, D.J., Irani, K., Yu, Z.X., Ferrans, V.J., Goldschmidt-Clermont, P. and Finkel, T., rac1 regulates a cytokine-stimulated, redox-dependent pathway necessary for NF-kappaB activation. Mol Cell Biol, 1996. 16(12): p. 7115-7121. Kauffmann-Zeh, A., Rodriguez-Viciana, P., Ulrich, E., Gilbert, C., Coffer, P., Downward, J. and Evan, G., Suppression of c-Myc-induced apoptosis by Ras signalling through Pl(3)K and PKB. Nature, 1997. 385(6616): p. 544-548. Chang, F., Steelman, L.S., Shelton, J.G., Lee, J.T., Navolanic, P.M., Blalock, W.L., Franklin, R. and McCubrey, J.A., Regulation of cell cycle progression and apoptosis by the Ras/RaIVMEK/ERK pathway (Review). Int J Oncol, 2003. 22(3): p. 469-480. Troppmair, J. and Rapp. U.R., Raf and the road to cell survival: a tale of bad spells, ring bearers and detours. Biochem Pharmacol, 2003. 66(8): p. 1341-1345. Ries, 8., Biederer, C., Woods, 0., Shifman, O., Shirasawa, S., Sasazuki, T., McMahon, M., Oren, M. and McCormick, F., Opposing effects of Ras on p53: transcriptional activation of mdm2 and induction of p19ARF. Cell, 2000. 103(2): p. 321-330. Burridge, K. and Wennerberg, K., Rho and Rac take center stage. Cell, 2004. 116(2): p. 167-179. Shields, J.M., Pruitt, K., McFall, A., Shaub, A. and Der, C.J., Understanding Ras: 'it ain't over 'til it’s over'. Trends Cell Biol, 2000. 10(4): p. 147-154. 125 310. 311. 312. 313. 314. 315. 316. 317. 318. 319. 320. 321. Allen, W.E., Zicha, D., Ridley, A.J. and Jones, G.E., A role for Cdc42 in macrophage chemotaxis. J Cell Biol, 1998. 141(5): p. 1147-1157. Banyard, J., Anand-Apte, B., Symons, M. and Zetter, B.R., Motility and invasion are differentially modulated by Rho family GTPases. Oncogene, 2000. 19(4): p. 580-591. Doanes, A.M., Irani, K., Goldschmidt-Clermont, P.J. and Finkel, T., A requirement for rac1 in the PDGF-stimulated migration of fibroblasts and vascular smooth cells. Biochem Mol Biol Int, 1998. 45(2): p. 279-287. Nobes, CD. and Hall, A., Rho GTPases control polarity, protrusion, and adhesion during cell movement. J Cell Biol, 1999. 144(6): p. 1235-1244. Fox, P.L., Sa, 6., Dobrowolski, SF. and Stacey, D.W., The regulation of endothelial cell motility by p21 ras. Oncogene, 1994. 9(12): p. 3519-3526. Cheresh, D.A., Leng, J. and Klemke, R.L., Regulation of cell contraction and membrane ruffling by distinct signals in migratory cells. J Cell Biol, 1999. 146(5): p. 1107-1116. McGlade, J., Brunkhorst, B., Anderson, D., Mbamalu, G., Settleman, J., Dedhar, S., Rozakis-Adcock, M., Chen, LB. and Pawson, T., The N- terminal region of GAP regulates cytoskeletal structure and cell adhesion. Embo J, 1993. 12(8): p. 3073-3081. Kulkami, S.V., Gish, G., van der Geer, P., Henkemeyer, M. and Pawson, T., Role of p120 Ras-GAP in directed cell movement. J Cell Biol, 2000. 149(2): p. 457-470. Kranenburg, O., Gebbink, M.F. and Voest, E.E., Stimulation of angiogenesis by Ras proteins. Biochim Biophys Acta, 2004. 1654(1): p. 23-37. Collier, I.E., Wilhelm, S.M., Eisen, A.Z., Mariner, B.L., Grant, G.A., Seltzer, J.L., Kronberger, A., He, C.S., Bauer, EA. and Goldberg, G.l., H-ras oncogene-transformed human bronchial epithelial cells ( TBE- 1) secrete a single metalloprotease capable of degrading basement membrane collagen. J Biol Chem, 1988. 263(14): p. 6579-6587. Westermarck, J., Li, S.P., Kallunki, T., Han, J. and Kahari, V.M., p38 mitogen-activated protein kinase—dependent activation of protein phosphatases 1 and 2A inhibits MEK1 and MEK2 activity and collagenase 1 (MMP-1) gene expression. Mol Cell Biol, 2001. 21(7): p. 2373-2383. Yang, J.Q., Zhao, W., Duan, H., Robbins, M.E., Buettner, G.R., Oberley, L.W. and Domann, F .E., v-Ha-RaS oncogene upregu/ates the 92-kDa type 126 322. 323. 324. 325. 326. 327. 328. 329. 330. 331. 332. IV collagenase (MMP-9) gene by increasing cellular superoxide production and activating NF-kappaB. Free Radic Biol Med, 2001. 31(4): p. 520-529. Furge, K.A., Kiewlich, D., Le, P., Vo, M.N., Faure, M., Howlett, A.R., Lipson, K.E., Woude, GP. and Webb, C.P., Suppression of Ras-mediated tumorigenicity and metastasis through inhibition of the Met receptor tyrosine kinase. Proc Natl Acad Sci U S A, 2001. 98(19): p. 10722-10727. Webb, CR, Taylor, G.A., Jeffers, M., Fiscella, M., Oskarsson, M., Resau, J.H. and Vande Woude, G.F., Evidence for a role of Met-HGF/SF during Ras-mediated tumorigenesis/metastasis. Oncogene, 1998. 17(16): p. 2019-2025. Hacohen, N., Kramer, S., Sutherland, D., Hiromi, Y. and Krasnow, M.A., sprouty encodes a novel antagonist of FGF signaling that patterns apical branching of the Drosophila airways. Cell, 1998. 92(2): p. 253-263. Warburton, D., Schwarz, M., Tefft, D., Flores-Delgado, G., Anderson, KB. and Cardoso, W.V., The molecular basis of lung morphogenesis. Mech Dev, 2000. 92(1): p. 55-81. Placzek, M. and Skaer, H., Airway patterning: A paradigm for restricted signalling. Curr Biol, 1999. 9(14): p. R506-510. Casci, T., Vinos, J. and Freeman, M., Sprouty, an intracellular inhibitor of Ras signaling. Cell, 1999. 96(5): p. 655-665. Reich, A., Sapir, A. and Shilo, B., Sprouty is a general inhibitor of receptor tyrosine kinase signaling. Development, 1999. 126(18): p. 4139-4147. Gross, |., Bassit, B., Benezra, M. and Licht, J.D., Mammalian sprouty proteins inhibit cell growth and differentiation by preventing ras activation. J Biol Chem, 2001. 276(49): p. 46460-46468. Mailleux, A.A., Tefft, D., Ndiaye, D., Itoh, N., Thiery, J.P., Warburton, D. and Bellusci, 8., Evidence that SPROUTY2 functions as an inhibitor of mouse embryonic lung growth and morphogenesis. Mech Dev, 2001. 102(1-2): p. 81-94. Tefft, D., Lee, M., Smith, 8., Crowe, D.L., Bellusci, S. and Warburton, D., mSprouty2 inhibits FGF10-activated MAP kinase by differentially binding to upstream target proteins. Am J Physiol Lung Cell Mol Physiol, 2002. 283(4): p. L700-706. Nutt, S.L., Dingwell, K.S., Holt, CE. and Amaya, E., Xenopus Sprouty2 inhibits FGF-mediated gastrulation movements but does not affect mesoderm induction and patterning. Genes Dev, 2001. 15(9): p. 1152- 1166. 127 333. 334. 335. 336. 337. 338. 339. 340. 341. 342. 343. 344. Minowada, G., Jarvis, L.A., Chi, C.L., Neubuser, A., Sun, X., Hacohen, N., Krasnow, MA. and Martin, G.R., Vertebrate Sprouty genes are induced by FGF signaling and can cause chondrodysplasia when overexpressed. Development, 1999. 126(20): p. 4465-4475. Kim, H.J. and Bar-Sagi, D., Modulation of signalling by Sprouty: a developing story. Nat Rev Mol Cell Biol, 2004. 5(6): p. 441-450. Lee, S.H., Schloss, D.J., Jarvis, L., Krasnow, MA. and Swain, J.L., Inhibition of angiogenesis by a mouse sprouty protein. J Biol Chem, 2001. 276(6): p. 4128-4133. Guy, G.R., Wong, E.S., Yusoff, P., Chandramouli, 8., Lo, T.L., Lim, J. and Fong, C.W., Sprouty: how does the branch manager work? J Cell Sci, 2003. 116(Pt 15): p. 3061-3068. Christofori, G., Split personalities: the agonistic antagonist Sprouty. Nat Cell Biol, 2003. 5(5): p. 377-379. Rubin, C., Litvak, V., Medvedovsky, H., Zwang, Y., Lev, S. and Yarden, Y., Sprouty fine-tunes EGF signaling through interlinked positive and negative feedback loops. Curr Biol, 2003. 13(4): p. 297-307. Hall, A.B., Jura, N., DaSilva, J., Jang, Y.J., Gong, D. and Bar-Sagi, D., hSpry2 is targeted to the ubiquitin-dependent proteasome pathway by c- Cbl. Curr Biol, 2003. 13(4): p. 308-314. Hanafusa, H., Torii, S., Yasunaga, T. and Nishida, E., Sprouty1 and SproutyZ provide a control mechanism for the Ras/MAPK signalling pathway. Nat Cell Biol, 2002. 4(11): p. 850-858. Lim, J., Wong, E.S., Ong, S.H., Yusoff, P., Low, 3.0. and Guy, G.R., Sprouty proteins are targeted to membrane ruffles upon growth factor receptor tyrosine kinase activation. Identification of a novel translocation domain. J Biol Chem, 2000. 275(42): p. 32837-32845. Rubin, C., Zwang, Y., Vaisman, N., Ron, D. and Yarden, Y., Phosphorylation of carboxyl-terminal tyrosines modulates the specificity of Sprouty-2 inhibition of different signaling pathways. J Biol Chem, 2005. 280(10): p. 9735-9744. Haglund, K., Schmidt, M.H., Wong, E.S., Guy, GR. and Dikic, l., Sprouty2 acts at the CbI/CIN85 interface to inhibit epidermal growth factor receptor downregulation. EMBO Rep, 2005. 6(7): p. 635-641. Kramer, 8., Okabe, M., Hacohen, N., Krasnow, MA. and Hiromi, Y., Sprouty: a common antagonist of FGF and EGF signaling pathways in Drosophila. Development, 1999. 126(11): p. 2515-2525. 128 345. 346. 347. 348. 349. 350. 351. 352. 353. 354. 355. Robert, C., Gagne, D., Bousquet, D., Barnes, F.L. and Sirard, M.A., Differential display and suppressive subtractive hybridization used to identify granulosa cell messenger rna associated with bovine oocyte developmental competence. Biol Reprod, 2001. 64(6): p. 1812-1820. Chambers, D. and Mason, 1., Expression of sprouty2 during early development of the chick embryo is coincident with known sites of FGF signalling. Mech Dev, 2000. 91(1-2): p. 361-364. Zhang, 8., Lin, Y., ltaranta, P., Yagi, A. and Vainio, 8., Expression of Sprouty genes 1, 2 and 4 during mouse organogenesis. Mech Dev, 2001. 109(2): p. 367-370. Gross, |., Morrison, D.J., Hyink, D.P., Georgas, K., English, M.A., Mericskay, M., Hosono, S., Sassoon, D., Wilson, P.D., Little, M. and Licht, J.D., The receptor tyrosine kinase regulator Sprouty1 is a target of the tumor suppressor WT1 and important for kidney development. J Biol Chem, 2003. 278(42): p. 41420-41430. Tefft, J.D., Lee, M., Smith, S., Leinwand, M., Zhao, J., Bringas, P., Jr., Crowe, D.L. and Warburton, D., Conserved function of mSpry-2, a murine homolog of Drosophila sprouty, which negatively modulates respiratory organogenesis. Curr Biol, 1999. 9(4): p. 219-222. Liu, H., Chen, J.Z., Gu, S.H., Dai, J.L., Zhao, E.P., Huang, L., Xu, W.X., Xie, Y. and Mao, Y.M., Assignment of human sprouty 4 gene to chromosome segment 5q32 approximately 33 and analysis of its pattern of expression. J Genet, 2003. 82(1-2): p. 23-26. Ozaki, K., Kadomoto, R., Asato, K., Tanimura, S., Itoh, N. and Kohno, M., ERK pathway positively regulates the expression of Sprouty genes. Biochem Biophys Res Commun, 2001. 285(5): p. 1084-1088. Abe, M. and Naski, M.C., Regulation of sprouty expression by PLCgamma and calcium-dependent signals. Biochem Biophys Res Commun, 2004. 323(3): p. 1040-1047. Ding, W., Bellusci, S., Shi, W. and Warburton, D., Functional analysis of the human Sprouty2 gene promoter. Gene, 2003. 322: p. 175-185. Lim, J., Yusoff, P., Wong, E.S., Chandramouli, S., Lao, D.H., Fong, CW. and Guy, G.R., The cysteine-rich sprouty translocation domain targets mitogen-activated protein kinase inhibitory proteins to phosphatidylinositol 4,5-bisphosphate in plasma membranes. Mol Cell Biol, 2002. 22(22): p. 7953-7966. Mason, J.M., Morrison, D.J., Bassit, B., Dimri, M., Band, H., Licht, JD. and Gross, l., Tyrosine phosphorylation of Sprouty proteins regulates their 129 356. 357. 358. 359. 360. 361. 362. 363. 364. 365. ability to inhibit growth factor signaling: a dual feedback loop. Mol Biol Cell, 2004. 15(5): p. 2176-2188. lmpagnatiello, M.A., Weitzer, S., Gannon, G., Compagni, A., Cotten, M. and Christofori, G., Mammalian sprouty-1 and -2 are membrane-anchored phosphoprotein inhibitors of growth factor signaling in endothelial cells. J Cell Biol, 2001. 152(5): p. 1087-1098. Cabrita, MA. and Christofori, G., Sprouty proteins: antagonists of endothelial cell signaling and more. Thromb Haemost, 2003. 90(4): p. 586- 590. Sasaki, A., Taketomi, T., Wakioka, T., Kato, R. and Yoshimura, A., Identification of a dominant negative mutant of Sprouty that potentiates fibroblast growth factor- but not epidermal growth factor-induced ERK activation. J Biol Chem, 2001. 276(39): p. 36804-36808. Egan, J.E., Hall, A.B., Yatsula, BA. and Bar-Sagi, D., The bimodal regulation of epidermal growth factor signaling by human Sprouty proteins. Proc Natl Acad Sci U S A, 2002. 99(9): p. 6041-6046. Wong, E.S., Lim, J., Low, B.C., Chen, Q. and Guy, G.R., Evidence for direct interaction between Sprouty and Cbl. J Biol Chem, 2001. 276(8): p. 5866-5875. Thien, CB. and Langdon, W.Y., Cbl: many adaptations to regulate protein tyrosine kinases. Nat Rev Mol Cell Biol, 2001. 2(4): p. 294-307. Fong, C.W., Leong, H.F., Wong, E.S., Lim, J., Yusoff, P. and Guy, G.R., Tyrosine phosphorylation of Sprouty2 enhances its interaction with c-Cbl and is crucial for its function. J Biol Chem, 2003. 278(35): p. 33456-33464. Wong, E.S., Fong, C.W., Lim, J., Yusoff, P., Low, B.C., Langdon, W.Y. and Guy, G.R., Sprouty2 attenuates epidermal growth factor receptor ubiquity/ation and endocytosis, and consequently enhances Ras/ERK signalling. Embo J, 2002. 21(18): p. 4796-4808. Li, X., Brunton, V.G., Burgar, H.R., Wheldon, L.M. and Heath, J.K., FRSZ- dependent SRC activation is required for fibroblast growth factor receptor- induced phosphorylation of Sprouty and suppression of ERK activity. J Cell Sci, 2004. 117(Pt 25): p. 6007-6017. ' Hanafusa, H., Torii, S., Yasunaga, T., Matsumoto, K. and Nishida, E., Shp2, an SH2-containing protein-tyrosine phosphatase, positively regulates receptor tyrosine kinase signaling by dephosphorylating and inactivating the inhibitor Sprouty. J Biol Chem, 2004. 279(22): p. 22992- 22995. 130 366. 367. 368. 369. 370. 371. 372. 373. 374. 375. Xiao, 8., Rose, D.W., Sasaoka, T., Maegawa, H., Burke, T.R., Jr., Roller, P.P., Shoelson, SE. and Olefsky, J.M., Syp (SH-PTP2) is a positive mediator of growth factor-stimulated mitogenic signal transduction. J Biol Chem, 1994. 269(33): p. 21244-21248. Neel, B.G., Gu, H. and Pao, L., The 'Shp’ing news: SH2 domain-containing tyrosine phosphatases in cell signaling. Trends Biochem Sci, 2003. 28(6): p. 284-293. Lee, 00., Putnam, A.J., Miranti, C.K., Gustafson, M., Wang, L.M., Vande Woude, GP. and Gao, C.F., Overexpression of sprouty 2 inhibits HGF/SF- mediated cell growth, invasion, migration, and cytokinesis. Oncogene, 2004. 23(30): p. 5193-5202. Yusoff, P., Lao, D.H., Ong, S.H., Wong, E.S., Lim, J., Lo, T.L., Leong, H.F., Fong, CW. and Guy, G.R., Sprouty2 inhibits the Ras/MAP kinase pathway by inhibiting the activation of Raf. J Biol Chem, 2002. 277(5): p. 3195- 3201. Taketomi, T., Yoshiga, D., Taniguchi, K., Kobayashi, T., Nonami, A., Kato, R., Sasaki, M., Sasaki, A., lshibashi, H., Moriyama, M., Nakamura, K.l., Nishimura, J. and Yoshimura, A., Loss of mammalian Sprouty2 leads to enteric neuronal hyperplasia and esophageal achalasia. Nat Neurosci, 2005. Sasaki, A., Taketomi, T., Kato, R., Saeki, K., Nonami, A., Sasaki, M., Kuriyama, M., Saito, N., Shibuya, M. and Yoshimura, A., Mammalian Sprouty4 suppresses Ras-independent ERK activation by binding to Raf1. Nat Cell Biol, 2003. 5(5): p. 427-432. Tsavachidou, D., Coleman, M.L., Athanasiadis, G., Li, S., Licht, J.D., Olson, M.F. and Weber, B.L., SPRY2 is an inhibitor of the ras/extracellular signal-regulated kinase pathway in melanocytes and melanoma cells with wild-type BRAF but not with the V599E mutant. Cancer Res, 2004. 64(16): p. 5556-5559. Yigzaw, Y., Cartin, L., Pierre, 8., Scholich, K. and Patel, T.B., The C terminus of sprouty is important for modulation of cellular migration and proliferation. J Biol Chem, 2001. 276(25): p. 22742-22747. Yarden, Y. and Sliwkowski, M.X., Untangling the ErbB signalling network. Nat Rev Mol Cell Biol, 2001. 2(2): p. 127-137. Vaudry, D., Stork, P.J., Lazarovici, P. and Eiden, L.E., Signaling pathways for PC12 cell differentiation: making the right connections. Science, 2002. 296(5573): p. 1648-1649. 131 376. 377. 378. 379. 380. 381. 382. 383. 384. 385. 386. Stang, E., Blystad, F.D., Kazazic, M., Bertelsen, V., Brodahl, T., Raiborg, C., Stenmark, H. and Madshus, |.H., Obi-dependent ubiquitination is required for progression of E GF receptors into clathrin-coated pits. Mol Biol Cell, 2004. 15(8): p. 3591-3604. Leeksma, O.C., Van Achterberg, T.A., Tsumura, Y., Toshima, J., Eldering, E., Kroes, W.G., Mellink, C., Spaargaren, M., Mizuno, K., Pannekoek, H. and de Vries, C.J., Human sprouty 4, a new ras antagonist on 5q31, interacts with the dual specificity kinase TESK1. Eur J Biochem, 2002. 269(10): p. 2546-2556. Bamburg, J.R., Proteins of the ADF/cofi/in family: essential regulators of actin dynamics. Annu Rev Cell Dev Biol, 1999. 15: p. 185-230. Toshima, J., Toshima, J.Y., Amano, T., Yang, N., Narumiya, S. and Mizuno, K., Cofilin phosphorylation by protein kinase testicular protein kinase 1 and its role in integrin-mediated actin reorganization and focal adhesion formation. Mol Biol Cell, 2001. 12(4): p. 1131-1145. Tsumura, Y., Toshima, J., Leeksma, O.C., Ohashi, K. and Mizuno, K., Sprouty-4 negatively regulates cell spreading by inhibiting the kinase activity of testicular protein kinase. Biochem J, 2005. 387(Pt 3): p. 627- 637. Yigzaw, Y., Poppleton, H.M., Sreejayan, N., Hassid, A. and Patel, T.B., Protein-tyrosine phosphatase-1B (PTP1B) mediates the anti-migratory actions of Sprouty. J Biol Chem, 2003. 278(1): p. 284-288. Poppleton, H.M., Edwin, F., Jaggar, L., Ray, R., Johnson, LR. and Patel, T.B., Sprouty regulates cell migration by inhibiting the activation of Rac1 GTPase. Biochem Biophys Res Commun, 2004. 323(1): p. 98-103. Basson, M.A., Akbulut, S., Watson-Johnson, J., Simon, R., Carroll, T.J., Shakya, R., Gross, 1., Martin, G.R., Lufkin, T., McMahon, A.P., Wilson, P.D., Costantini, F.D., Mason, l.J. and Licht, J.D., Sprouty1 is a critical regulator of GDNF/RET-mediated kidney induction. Dev Cell, 2005. 8(2): p. 229-239. Shim, K., Minowada, G., Coling, DE. and Martin, G.R., Sprouty2, a mouse deafness gene, regulates cell fate decisions in the auditory sensory epithelium by antagonizing FGF signaling. Dev Cell, 2005. 8(4): p. 553- 564. Saxen, L. and Sariola, H., Early organogenesis of the kidney. Pediatr Nephrol, 1987. 1(3): p. 385-392. Moore, M.W., Klein, R.D., Farinas, |., Sauer, H., Arrnanini, M., Phillips, H., Reichardt, L.F., Ryan, A.M., Carver-Moore, K. and Rosenthal, A., Renal 132 387. 388. 389. 390. 391. 392. 393. 394. 395. 396. and neuronal abnormalities in mice lacking GDNF. Nature, 1996. 382(6586): p. 76-79. Vainio, S. and Lin, Y., Coordinating early kidney development: lessons from gene targeting. Nat Rev Genet, 2002. 3(7): p. 533-543. Takahashi, M., The GDNF/RET signaling pathway and human diseases. Cytokine Growth Factor Rev, 2001. 12(4): p. 361-373. Majumdar, A., Vainio, S., Kispert, A., McMahon, J. and McMahon, A.P., Wnt11 and Ret/Gdnf pathways cooperate in regulating ureteric branching during metanephric kidney development. Development, 2003. 130(14): p. 3175-3185. Chi, L., Zhang, 8., Lin, Y., Prunskaite-Hyyrylainen, R., Vuolteenaho, R., ltaranta, P. and Vainio, S., Sprouty proteins regulate ureteric branching by coordinating reciprocal epithelial Wnt11, mesenchymal Gdnf and stromal Fgf7 signalling during kidney development. Development, 2004. 131(14): p. 3345-3356. Pirvola, U., Ylikoski, J., Trokovic, R., Hebert, J.M., McConnell, SK. and Partanen, J., FGFR1 is required for the development of the auditory sensory epithelium. Neuron, 2002. 35(4): p. 671-680. Colvin, J.S., Bohne, B.A., Harding, G.W., McEwen, DO. and Omitz, D.M., Skeletal overgrowth and deafness in mice lacking fibroblast growth factor receptor 3. Nat Genet, 1996. 12(4): p. 390-397. Lo, T.L., Yusoff, P., Fong, C.W., Guo, K., McCaw, B.J., Phillips, W.A., Yang, H., Wong, E.S., Leong, H.F., Zeng, Q., Putti, TC. and Guy, G.R., The ras/mitogen-activated protein kinase pathway inhibitor and likely tumor suppressor proteins, sprouty 1 and sprouty 2 are deregulated in breast cancer. Cancer Res, 2004. 64(17): p. 6127-6136. Kwabi-Addo, B., Wang, J., Erdem, H., Vaid, A., Castro, R, Ayala, G. and Ittmann, M., The expression of Sprouty1, an inhibitor of fibroblast growth factor signal transduction, is decreased in human prostate cancer. Cancer Res, 2004. 64(14): p. 4728-4735. McKie, A.B., Douglas, D.A., Olijslagers, 8., Graham, J., Omar, M.M., Heer, R., Gnanapragasam, V.J., Robson, ON. and Leung, H.Y., Epigenetic inactivation of the human sprouty2 (hSPRY2) homologue in prostate cancer. Oncogene, 2005. 24(13): p. 2166-2174. Bloethner, S., Chen, 8., Hemminki, K., Muller-Berghaus, J., Ugurel, S., Schadendorf, D. and Kumar, R., Effect of common B-RAF and N-RAS mutations on global gene expression in melanoma cell lines. Carcinogenesis, 2005. 26(7): p. 1224-1232. 133 Chapter II. Sprouty 2 is necessary for tumor formation by HRas oncogene-transformed human fibroblasts Running title: Spry2 is necessary for HRas-transformation Piro Lito, Bryan D. Mets, Sandra O’Reilly, Veronica M. Maher and J. Justin McCormick' Carcinogenesis Laboratory, Department of Microbiology & Molecular Genetics and Department of Biochemistry & Molecular Biology, Michigan State University, East Lansing, Michigan 48824-1302, USA. *Correspondence: J. Justin McCormick, Carcinogenesis Laboratory, Food Safety and Toxicology Bldg., Michigan State University, East Lansing, MI 48824-1302, USA; Ph: (517) 353-7785; Fax: (517) 353-9004. E-mail address: mccormi1@msu.edu Keywords: Sprouty; HRas oncogene; malignant transformation; epidermal growth factor receptor 134 Abstract Sprouty 2 (Spry2) plays a regulatory role in the signaling pathways induced by a number of growth factors. One aspect of the function of Spry2 is to prevent the c-Cbl-induced degradation of epidermal growth factor receptor (EGFR). We report that human fibroblasts malignantly transformed by HRas"12 oncogene, exhibited an increase in the expression of Spry2, compared to the parental cells. To determine whether Spry2 plays a role in HRas-transformation, we down-regulated the expression of Spry2 using Spry2- specific shRNA. HRas-transformed cells with down-regulated levels of Spry2 failed to form tumors when injected into athymic mice, indicating that Spry2 in necessary for tumor formation by HRas-transformed cells. In cells expressing oncogenic HRas, Spry2 sustained not only the level of EGFR, but also the activation of ERK. What is more, HRas interacted with Spry2 in HRas-transformed cells, and HRas interacted with c-Cbl and CIN85 in a Spry2-dependent fashion, suggesting that HRas regulates the turnover of EGFR through Spry2. We also found that expression of Spry2 in immortalized human fibroblasts, did not affect EGFR levels, while it inhibited HRas and ERK activation. The effect on ERK was diminished when Spry2 was expressed at a higher level. These data show that Spry2 has distinct functions in pre-malignant and in malignant fibroblasts transformed by HRas. While in the former Spry2 can inhibit EGF signaling, in the latter, the inhibitory function of Spry2 is bypassed and the ability of Spry2 to sustain EGFR takes center stage. 135 Introduction Carcinogenesis is a multistep process by which cells acquire neoplastic characteristics through a series of genetic and/or epigenetic changes. To study this process, McCormick and colleagues developed a model system which mimics the pattern by which normal human fibroblasts become malignant [1]. In these experiments, normal foreskin-derived human skin fibroblasts were transfected with a v-Myc oncogene, giving rise to a clonal population of cells, which spontaneously acquired an infinite life span in culture [2]. This cell strain, which has a normal diploid karyotype, was designated MSU- 1.0. As MSU-1.0 cells were being propagated in culture, one cell underwent two chromosomal translocations, giving rise to a cell strain that is chromosomally stable, near-diploid, and partially growth factor independent [2]. This cell strain, designated, MSU-1.1, has been malignantly transformed by the overexpression of Ras oncogenes [2-5] or by exposure to a carcinogen, followed by selection of focus-forming cells [6, 7]. V12 When MSU-1.1 cells expressing high levels of HRas oncoprotein are injected subcutaneously into athymic mice, they form fibrosarcomas within three weeks. A malignant cell strain derived from such tumors, designated PH3MT, is completely growth factor independent and exhibits anchorage independent growth [4]. Sprouty was first identified in Drosophila as an inhibitor of fibroblast growth factor (FGF)- induced tracheal branching [8] and epidermal growth factor (EGF)—induced eye development [9]. Mammalian species express four isoforms of Sprouty (Spry1-4) [8, 10, 11], which act as inhibitors of growth factor-induced cellular differentiation, migration, and proliferation [12-15]. In addition to its inhibitory function, Spry2 also sustains the EGFR signaling [16-21]. This 136 function is mainly the result of the interaction of Spry2 with c-Cbl, an E3 ubiquitin ligase that catalyzes the ubiquitination of EGFR, targeting this receptor for lysosomal degradation [22]. By binding to c-Cbl, Spry2 prevents the interaction between c-Cbl and EGFR, and this interference blocks the degradation of the receptor. This in turn leads to sustained EGFR-induced ERK activity [23-26]. The expression of Spry2 gene is altered in several types of cancer. Spry2 is down - regulated in breast cancer [27], and expression of Spry2 in breast cancer cells (MCF7) reduces the ability of these cells to form tumors in athymic mice [27]. Spry2 is also down- regulated in prostate cancer, which is attributed to epigenetic inactivation, through hyper- methylation of the Spry2 promoter [28]. Spry2 expression is also elevated in some cancer subtypes, including melanomas that express activated Ras signaling pathways [29, 30], suggesting that Spry2 contributes to the malignant phenotype in cells expressing oncogenic Ras. The present study was designed to investigate the role of Spry2 in tumor formation in malignant cells with activated Ras signaling. We found that HRas-transformed human fibroblasts (PH3MT) expressed a higher level of Spry2 protein than that found in their parental cells (MSU-1.1). When we stably down-regulated the expression of Spry2 in PH3MT cells, we observed a complete loss of tumor-forming ability by these cells. In PH3MT cells Spry2 sustained the level of EGFR and ERK activation. Interestingly, independent expression of Spry2 in MSU-1.1 cells resulted in the inhibition of EGF signaling and was insufficient to malignantly transform these cells. 137 Results Determination of the Expression of Spry2 in HRas-transformed Cells. To detetermine the effect of HRas-transformation on the expression of Spry2 we examined the cells of the MSU lineage by Northern and Western blotting (Fig. 1A and B). Although compared to MSU-1.0 cells, MSU-1.1 cells exhibit only a modest increase in Spry2 expression, the cells malignantly transformed by HRasWZ oncogene (PH3MT) exhibit a significant increase in Spry2 expression. An independent HRas-transformed cell strain (PH2MT) was found to have the same high expression of Spry2 (Fig. 1C). These data suggest that oncogenic HRas is responsible for the high level of Spry2 protein found in PH3MT cells. To determine if the same was true for oncogenic NRas, we examined the levels of expression of Spry2 protein in MSU-1.1 cell strains malignantly transformed by the NRasV’Z oncogene, i.e., N-Ras-2T and N-Ras-3T [5]. NRas-transformed cells expressed Spry2 at a level similar to that present in the HRas-transformed cells, i.e., PH2MT and PH3MT (Fig. 10). We also examined the level of expression of Spry2 in a four patient-derived fibrosarcoma and seven patient-derived pancreatic carcinoma cell lines. As shown in Fig. 10, all four fibrosarcoma-derived cell lines expressed high levels of Spry2, compared to normal foreskin-derived fibroblasts. Of seven pancreatic carcinoma-derived cell lines analyzed for expression of Spry2, three cell lines 138 Figure 1. Expression profile of Spry2 in Ras-transformed cells and in patient derived cancer cells. Northern (A) and Western (B) blots showing the expression of Spry2 in immortalized human fibroblasts strains MSU-1.0 and MSU-1.1, and in MSU-1.1 cells malignantly transformed by the HRas oncogene (PH3MT). (C) The expression of Spry2 protein in MSU-1.1 cells malignantly transformed by HRas and NRas oncogenes. Cell strains that were derived by the malignant transformation of MSU-1.1 cells with HRas (PH2MT and PH3MT) or NRas (NRasZT and NRas3T) oncogenes, were analyzed by Western blotting with the indicated antibodies. (D) The expression of Spry2 protein in patient-derived fibrosarcoma cell lines and in normal, foreskin-derived, fibroblast cell lines (SL68 and SL89). (E) The expression of Spry2 protein in pancreatic carcinoma cell lines and in an infinite life span, pancreatic cell line (ps-1). 139 Spry2 ._._>_mIn_ w. 7322 o. rims. 511 “mm u —-—-. KUBO ._._>_mIn_ F. 7392 o. Tam—2 flu Spry2 u .d GAPDH _.O_mIn_ ._._2NT_n_ I... c. F. Tam: D. Spry2 _OZ 1 ._.n_i_> owe—FT. T O 1.0 1.1/1.0 PH3MT/1 0 1611313'3 of £33.31 U37139 88203476 086978 L25931 88014458 X64229 8L080146 88019987 88004770 U96131 026361 8F011468 U97067 Y00971 8L050353 L07493 014657 L07541 8F067656 X54942 U65410 ”87339 080008 8F053641 U14518 8F025441 017517 88926959 L25876 X56597 "68520 8J236876 X51688 038551 U30872 8F017790 L19183 084557 080000 L47276 J04031 88024704 8F047473 U94319 031885 "15796 X62534 8F047472 M31303 X51688 028423 X01060 "30448 8F039656 HT3165 K02581 J04988 013748 X57351 X02308 8F039843 Figure 1 6 .032 214 01064 6.095 1 1/1 0 FIGHT/1.0 °. r1 Figure 1 (Cont’) 215 patterns of change, we used the EPCLUSTER software1 to perform hierarchical, as well as K- means clustering of these genes. (Fig. 1 and 2). Differentially expressed genes were grouped together according to their functional characteristics (Table I). This was done in order to obtain a global view of the genotypic changes that take place as cell stains in the MSU-1 lineage became malignantly transformed. For a broader picture, some genes with marginal (less than 3-fold) changes between MSU 1.1 and MSU-1.0 cells were also included. The examination of these groups revealed some noteworthy traits. First, all of the genes that encode transcription factors or chromatin remodeling enzymes were up regulated in MSU-1.1 and PH3MT cells. Second, the type of change for the majority of the genes involved in cell cycle control supports a faster G1-S transition, and therefore a faster growth rate in the MSU-1.1 and PH3MT cell strains. This is consistent with prior studies that report the MSU-1.0 cell strain having a slovVer growth rate compared to the cell strains derived from it. Third, the majority of the genes encoding extracellular matrix (ECM) proteins, or proteins involved in cell adhesion and motility, were found to be down regulated in MSU-1.1 and PH3MT cells. Finally, genes involved in calcium signaling, or genes with calcium-binding domains, were down regulated in MSU-1.1 and PH3MT cells. ‘ This is available for use at the European Bioinformatics lnstitute’s web site: http:/lep.ebi.ac.uk/EP/ 216 Figure 2. K-means clustering of the differentially expressed genes. The y-axis represents the logarithmic fold-change in gene expression from the MSU-1.1 vs. MSU- 1.0 comparison (first point of discontinuity), and the PH3MT vs. MSU-1.0 comparison (second point of discontinuity). Each line in the graphs represents a gene. Red indicates an increase, while green indicates a decrease in gene expression compared to MSU-1.0 cells. Each graph shows genes with similar patterns of change. 217 N 0.59... :15 ‘15 11:1]. he 1" L. 89.86 2 86- Sufi: 2 86- 828 8963.2 86- 6m... . 8..- .9 8...- 8..- . ammmhz_ . Mom”. . OO.HI 04>: . 8..“l XND : S.HI Duo 6 238: $02: 8%“ 25.23 2356 . 2... . 8.5 . <28 II\ 8 a :4on82. . 8 6 Km: \ 8 6 FIND; . 8.N Uw.~zm._.200 . 8.N Nm 230>0 oo.N .wpzmezoo . 2 . 6228 .. 8 6 . 8 6 6.23200 86 39.83 .. 861 anon? . 861 «~03? . 86.. . E . - 3.8 . i . 8 T wmpm . 8 N 82528 8 ~1 . 86- 3 . 86- .05. 22 . . Nfiflm 2m“. 82.2 8 a obs: ~ Pan-Ras ’ "‘ PDzRaf-RBD “"- .. 4 Pan-Ras WCL ' B D latte HT1080 vc so C10 C12 [....... I:- ‘ Spry2 C E VleFT VC SC B1 I- ~ 4 Spry2 l - It I'. \ -‘.0 ',l41—r '3‘ Ir“. - I.‘ a“ "J Ku 80 Figure 1 p-ERK “Mm fluwflm“ ERK 238 Control cells and cells with down regulated Spry2 were then analyzed for their ability to form tumors in athymic mice. Parental HT1080 cells consistently form sarcomas in athymic mice (100 tumors/100 injected sites) with a latency of three to four weeks. HT1080 cells expressing the scrambled shRNA, formed tumors at a similar ratio as did the parental cells, i.e. 7 tumor/8 sites. HT1080 cells with down regulated Spry2 formed fewer tumors, i.e. 6 tumors/10 sites (Table l). Parental VleFT cells also form tumors in athymic mice with the same consistency as do parental VleFT cells (37 tumors/37 injected sites), although they exhibit a longer latency (approximately five weeks). VIP:FT cells expressing the scrambled shRNA formed tumors at a ratio resembling that of the parental cell line, i.e., 4 tumors/6 sites, with a latency of approximately five weeks. VleFT cells with down regulated Spry2, however, formed significantly fewer tumors upon injection in athymic mice, i.e. 6 tumors/16 sites. The latency of the tumors formed was the same as in the parental and control cells (Table II). These preliminary results suggest that both in the presence of wild type as well as in the presence of mutant Ras, Spry2 contributes to the ability of cells to form tumors in athymic mice. According to these findings, Spry2 plays a more prominent role in the tumorgenicity of VIPzFT cells (wild type Ras), than it does in HT1080 cells (NRasQSQ). In addition to expressing NRastQ, HT1080 cells, also overexpress PDGF. The latter may contribute significantly to the malignancy of these cells. Interestingly, Spry2 inhibits ERK activation and proliferation in 239 Table l The tumorigenicity of the HT1080 cell lines with down-regulated Spry2 Days for tumor to Cell Tumor reach 0.5 cm3 strain shRNA incidence" volume HT1080 - 100/100 21 -28 HT1080-SC scrambled 7/8 (87.5%) 35-63 HT1080-C10 spry2 6/10 (60%) 35-63 * Ratio of tumors formed to the number of sites injected subcutaneously. The mice were examined for tumor formation for at least 6 months after injection 240 response to PDGF stimulation, while it sustains ERK activation and growth in soft agar in HRas oncogene-transformed cells. These findings lend the preliminary hypothesis that in HT1080 cells, the ability of Spry2 to sustain RTK signaling in cells with activated Ras may be balanced, to some extent, by its ability to inhibit the same pathway in response to PDGF-stimulation, resulting in a diminished effect in tumor formation. Effect of Spry2 on EGF-induced cell cycle progression We have previously shown that Spry2 sustains the levels of EGFR and the signaling activity from this receptor, in fibrosarcomas containing HRasV12 mutations. To determine if Spry2 has a similar function in fibrosarcomas expressing mutant or wild type Ras, we examined the expression of EGFR following EGF-stimulation in the presence and absence of Spry2 expression in HT1080 and VIPFT cells. As expected, the levels of EGFR and the levels of active ERK were decreased in both HT1080 and VleFT cells with down regulated Spry2, compared to control cells (Fig. 1D and 1E). Since the down regulation of Spry2 in VIPzFT cells resulted in a more prominent decrease in tumorigenicity, compared to the effect of Spry2 down regulation in HT1080 cells, we focused our attention on function of Spry2 in these cells. Growth factor induced activation of ERK is an important step for cell cycle progression. To determine if the decrease in the activation of ERK, observed in the fibrosarcomas with down regulated Spry2, was associated with changes in 241 Table II The tumorigenicity of the VleFT cell lines with down- regulated Spry2 Days for tumor to Cell Tumor reach 0.5 cm3 strain shRNA incidence* volume VIPzFI' - 37/37 28-35 VleFT-SC scrambled 4/6 (60%)“ 28-84 VlP:FT—B1 spry2 6/16 (37.5%) 28-84 * Ratio of tumors formed to the number of sites injected subcutaneously. The mice were examined for tumor formation for at least 6 months after injection “The remaining 2/6 injections formed tumors that regressed to normal 242 cell cycle progression we examined the cell cycle distribution of VleFT cells expressing the scrambled shRNA to the distribution of VIP:FT cells expressing Spry2 shRNA. The cells were serum starved for a period of 60 hrs, and then stimulated with EGF to induced cell cycle progression. As soon as 2hrs after EGF stimulation, the proportion of control VleFT cells in G1 decreased, while the proportion in 8 increased, indicating progress from G1 to S phase (Fig. 2). Instead, in VIPzFT cells with down regulated Spry2, the proportion of cells in G1 remained unaltered 2 hrs after EGF stimulation, suggesting a delay in S phase entry. Furthermore, the population of control VleFT cells at G2 was prominent after ten hours of EGF-stimulation. We were unable to detect a significant population of VleFT cells with down regulated Spry2 that were at G2 at least until after 15 hrs of EGF stimulation. The proportion of cells at S phase was similar between the two cell strains. Importantly, the proportion of cells with DNA content below 2n, possibly apoptotic cells, was increased in the cells with down regulated Spry2. Discussion The data presented here suggest that Spry2 promotes tumor formation in human patient-derived fibrosarcoma cell lines, consistent with the role of Spry2 in PH3MT cells. Also, the ability of Spry2 to promote tumor formation by these cells is independent of the presence of activating Ras mutations. It should be noted that several months after their generation, the cell lines with down regulated Spry2 had regained expression of Spry2, most likely due to loss of expression of 243 Figure 2. Down regulation of Spry2 in VleFT cells delays progression through the cell cycle. VleFT cells expressing either scrambled shRNA or Spry2-specific shRNA were analyzed as described in Materials and Methods. 244 Will the at ShRNA WE VIPZFT Scrambled shRNA VIPZFT Spry2 shRNA Figure 2 245 the spry2-specific shRNA. This may account for the lesser effect that Spry2 had in the tumor forming ability of HT1080 and VIP:FT cellc compared to the effect of Spry2 on the tumor forming ability of PH3MT cells. Alternatively, Spry2 may have a more pronounced effect in HRas-transformed cells, because Spry2 may play a Ras isoform specific function. Material and methods Cell cycle analysis Cells plated at a density of 200,000 per 10cm-dish, were serum deprived for 48 to 72 hrs., and then stimulated with EGF (300 ng/mL) for varying time periods. After the period of stimulation was over, the cells were collected and fixed in 80% ethanol solution. FiXed cells were stained with propidium iodide (50 ng/mL) for 1 hr. at room temperature. The cell cycle analysis was performed by using FACS, using standard sttings. 246 llglljjglljjjljljjlljl