Characterization of Split ends function during Drosophila eye development by David Bagon Doroquez Bachelor of Science in Biology Santa Clara University, 1999 Submitted to the Department of Biology in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biology at the Massachusetts Institute of Technology May 2007 Copyright © 2007 David Bagon Doroquez. All rights reserved. The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part. Signature of Author: Certified by: Accepted by: David Bagon Doroquez Ilaria Rebay Associate Professor of Biology Thesis Supervisor Stephen P. Bell Professor Biology Chairman, Committee on Graduate Students Characterization of Split ends function during Drosophila eye development David Bagon Doroquez Submitted to the Biology Department on May 25, 2007 in partial fulfillment of the Requirements for the degree of Doctor of Philosophy in Biology Conserved signal transduction pathways coordinate all aspects of metazoan development, including cell fate specification, differentiation, and growth. Rather than functioning as completely independent modules, signaling pathways interface to create a web of specific interactions that a cell integrates in a spatial and temporal manner. The Epidermal Growth Factor Receptor (EGFR) and Notch pathways are evolutionarily conserved signal transduction mechanisms that interact intimately to regulate a broad spectrum of developmental processes in metazoa. The molecular bases underlying cross-talk and signal integration between these two pathways are just beginning to be elucidated. In this thesis, we have focused on the role of Split ends (Spen), the founding member of a family of transcriptional co-repressors, as a node of cross-talk between the EGFR and Notch signaling pathways during Drosophila eye development. At the morphogenetic furrow (MF), which marks the wave of differentiation that passes through the imaginal disc eye primordium, Notch signaling is required for establishing and refining the expression of proneural Atonal (Ato). Ato promotes the EGFR pathway’s reiterative signaling for progressive and sequential recruitment of cells within each ommatidial facet of the eye. Previous studies found that spen functioned within or in parallel to the EGFR pathway during midline glial cell development in the embryonic central nervous system. In vertebrates, Spen orthologs function in repressor complexes that antagonize the transcription of Notch pathway targets. The involvement of Spen proteins in EGFR and Notch signaling in these systems thus motivated us to explore the consequences of loss of spen function with respect to each pathway during eye development. Here we report that Spen acts as both a positive regulator of EGFR signaling and as an antagonist of Notch signaling in the eye. We find that loss of spen results in hyper-activation of the Notch pathway via upregulation of the Notch activator Scabrous. This results in loss of Ato and activated MAP kinase at the MF, therefore antagonizing output from the EGFR signaling pathway. As a consequence, there is a failure of cell fate specification in spen mutant ommatidia. These observations suggest that Spen modulates output from the Notch and EGFR pathways to ensure appropriate patterning during eye development. Additionally, we have characterized a transgene encoding nuclear localization sequencetagged Spen C-terminus that functions as a dominant-negative (SpenDN). The Spen C-terminus contains the evolutionarily conserved SPOC domain that is required for transcriptional repression. In order to identify components related to Spen function and to understand the processes in which Spen may be involved, we performed a genetic screen to identify dominant modifiers of the rougheye associated with eye-expressed SpenDN. Our results confirm interactions with the EGFR and Notch pathways, but also suggest functions for Spen in chromatin regulation and programmed cell death. As Spen-like proteins are involved in human development and disease, it will be important to elucidate the underlying molecular mechanism behind Spen function. Thesis Advisor: Ilaria Rebay Title: Associate Professor of Biology Dedicated to Lesley Hisayo Yamaki and Eufrosina Bagon and Quintin Lambino Doroquez and to the loving memory of Yutaka Yamaki, John Quintin Doroquez, and Aaron Lee Doroquez Acknowledgements I am truly blessed to be part of a great community that has supported me during my graduate career. I thank God for my colleagues, friends, and family who have provided me a firm foundation on which I can stand and perform the work described in this thesis. Without their undying encouragement and support, I would not be the person who I am today. I would first like to thank my loving wife Lesley Yamaki. She is my best friend, my cornerstone, my life-line. She has brought balance to my life and perspective. It is so great to have an awesome and understanding spouse like her who has been there through thick-and-thin from college and now through grad school. She has been my nurse in getting through hardships and difficulties. I thank her also for our cats Jolee and Eek! who have brought us such great joy. I thank my parents, Eufrosina and Quintin, who have provided me so much guidance, support, and encouragement throughout my academic career. They have taught me that faith and hard work can go miles. I also thank Michael and Jun/Melissa/Michael for their support. I offer my appreciation and my gratitude to William & Darleen Yamaki, Lorna Chang, and the Yamaki and Chang families for their encouragement and support. I would like to especially thank my advisor and mentor Ilaria Rebay. Even from 1,000 miles away she has always been there for me. She is an awesome, hard-working scientist that has been a role model. Although she was not physically present the past couple of years, when I needed a bit of guidance at the bench, I would often ask ‘WWID?’ (‘What would Ilaria do?’) I think it has helped my experiments. I often think of Ilaria as a coach, encouraging me to reach my utmost potential while maintaining my excitement. I thank her for her patience and her focus and attention for those in her lab. I express my gratitude to Ilaria for accommodating me in internet conference lab meetings. I feel like I was sitting with the lab (but I wish I had some of the bagels and goodies they had during the meeting). I thank all of those individuals in the Rebay Lab—current and former—for being great grad school companions and colleagues. They have all been there when I needed help. Who could ask for anything more? I offer my appreciation to Terry Orr-Weaver who has opened up her lab to me. I thank her for welcoming me to the lab and I am glad that I have been a part of it. I thank the OrrWeaver Lab for offering their companionship, welcoming me, and providing me with their interest and feedback. It has been exciting to learn about Drosophila signaling and cell cycle from these two labs. I am blessed and honored to have been a part of two fantastic laboratories during my graduate career. I am proud that I have walked the path of graduate school with them. I thank those who contributed to my academic development: Rick Fehon and lab, Jackie Lees, Paul Garrity, Mary-Lou Pardue, Peter Reddien, and Lizabeth Perkins. I thank my undergraduate advisor Leilani Miller (SCU) for her advice over the years. She showed me the great power of genetics in model systems. My experiences here have strengthened my understanding of that power. I also thank Jeanine Wiener-Kronish and Teiji Sawa (UCSF) for giving me my first chance at laboratory research. I would like to offer my gratitude to my friends, local and far: JFK, SCU, MIT Biograd2000, and the Elm St. Crew. They have provided healthy diversions for one to get through grad school. I offer my thanks to Fr. Clancy, Fr. Reynolds, and the Tech Catholic Community for helping me along my spiritual journey. Finally, I would like to thank the Oakland Raiders for providing me with a few football seasons of enjoyment for an East Bay native living in New England. Go Raiders! (...and, I guess, Go Pats! as long they are not playing Oakland.) Table of Contents Abbreviations and Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 Chapter 1. Signal integration during development: mechanisms of EGFR and Notch pathway function and cross-talk . . . . . . . . . 15 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 The Epidermal Growth Factor Receptor Pathway . . . . . . . . . . . . . . . . . . . . . 17 DER Ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 DER: The Drosophila EGF Receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 Activation of Ras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 Activation of the Raf/MEK/MAPK Cascade . . . . . . . . . . . . . . . . . . . . . . . 28 Regulation of Nuclear Targets via Activated MAPK . . . . . . . . . . . . . . . . . . . 30 The Notch Signaling Pathway in Drosophila . . . . . . . . . . . . . . . . . . . . . . . . 33 Structure of Notch and its Ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 Notch is processed to generate heterodimeric receptor . . . . . . . . . . . . . . . . . . 36 Notch discriminates between ligands through differential glycosylation . . . . . . . . . 37 Notch-Ligand interaction results in extracellular truncation . . . . . . . . . . . . . . . 38 Endocytosis regulates ligand-receptor interactions in the Notch pathway . . . . . . . . 41 Intramembrane cleavage by gamma-secretase releases intracellular Notch . . . . . . . 43 Transcriptional repression in the absence of signaling . . . . . . . . . . . . . . . . . . 45 Transactivation of target genes follows proteolytic cleavage . . . . . . . . . . . . . . . 48 Context-specific variations in Notch signaling . . . . . . . . . . . . . . . . . . . . . . 50 Interplay of the DER and Notch signaling pathways . . . . . . . . . . . . . . . . . . . . 54 Relationships between DER and Notch signaling during development . . . . . . . . . . 54 The Drosophila eye . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 DER and Notch regulate the Retinal Determination Gene Network . . . . . . . . . . . 57 DER and Notch at the morphogenetic furrow . . . . . . . . . . . . . . . . . . . . . . 60 Notch and DER in R8 founder cell determination . . . . . . . . . . . . . . . . . . . . 63 DER and Notch are required for cell cycle regulation during eye formation . . . . . . . 72 Combinatorial factors in the determination and differentiation of ommatidial cell types75 Planar polarity regulates R3 versus R4 identity . . . . . . . . . . . . . . . . . . . . . 76 R7 develops through orchestrated signaling by DER, Sev, and Notch . . . . . . . . . . 78 The DER and Notch pathways determine cone cell fate . . . . . . . . . . . . . . . . . 83 DER, Notch, Programmed Cell Death and the development of pigment cells . . . . . . 87 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 Chapter 2. Split ends: molecular function and role in development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 spen is linked to defects in Notch and EGFR signaling . . . . . . . . . . . . . . . . . . 122 Spen is required for multiple processes during Drosophila development . . . . . . . . . 124 Structure and molecular function of Spen family proteins . . . . . . . . . . . . . . .125 Spen family proteins are involved in developmental pathways and disease . . . . . . . . 133 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 Chapter 3. Split ends antagonizes the Notch and potentiates the EGFR signaling pathways during Drosophila eye development 143 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 Fly Strains and Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 Immunohistochemistry and Immunofluorescence . . . . . . . . . . . . . . . . . . . . 148 Quantitative RT-PCR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150 spen is required for proper proneural gene expression at the morphogenetic furrow . 150 Loss of spen leads to elevated levels of Sca protein and transcript . . . . . . . . . . . 151 Loss of spen results in elevated Notch signaling at the morphogenetic furrow . . . . . 153 Reduction of spen suppresses Notch haploinsufficiency phenotypes in the wing . . . 154 Loss of spen reduces dpERK levels in the developing eye . . . . . . . . . . . . . . . 154 Transcriptional output from the EGFR pathway is perturbed in the absence of spen . 156 yan is not transcriptionally regulated by Spen . . . . . . . . . . . . . . . . . . . . . 158 spen is required for patterning of neuronal and non-neuronal cell types during eye development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 Spen antagonizes Notch signaling . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 Elevated Notch pathway output leads to reduced EGFR signaling in spen mutant tissue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 Spen modulates EGFR and Notch signaling posterior to the morphogenetic furrow . 169 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 Supplementary Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 Chapter 4. A screen for genetic modifiers of the rough eye associated with dominant-negative Spen in Drosophila . . . . . . 183 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186 Fly Strains and Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186 Pilot F1 Deficiency Dominant Modifier Screen . . . . . . . . . . . . . . . . . . . . . 188 X-ray mutagenesis F1 dominant modifier screen . . . . . . . . . . . . . . . . . . . . 188 P Lethal and EP F1 Dominant Modifier Kit Screens . . . . . . . . . . . . . . . . . . 190 Histology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 Over-expression of the Spen C-terminus is dominant-negative . . . . . . . . . . . . . 190 Potential genetic interactors arise from pilot sev>spenDN modifier screen . . . . . . . 194 A P-element screen for modifiers of sev-spenDN . . . . . . . . . . . . . . . . . . . . 203 X-ray screen identifies modifiers including new alleles of spen and pointed . . . . . . 205 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206 Eye-specific over-expression of the Spen C-terminus is dominant-negative . . . . . . 206 EGFR and Notch pathway components interact genetically with dominant-negative spen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210 Spen may regulate chromatin state . . . . . . . . . . . . . . . . . . . . . . . . . . . 211 Spen may be a regulator of cell death . . . . . . . . . . . . . . . . . . . . . . . . . . 212 Conclusions and state of the sev-spenDN screen . . . . . . . . . . . . . . . . . . . . . 213 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214 Chapter V. Discussion and Future Directions . . . . . . . . . . . . . . . 219 The requirement of spen for the antagonism of Notch signaling . . . . . . . . . . . . . 220 Spen opposition to Notch’s antagonism of the EGFR pathway . . . . . . . . . . . . . . 226 Notch-independent role for Spen in EGFR signaling . . . . . . . . . . . . . . . . . . . 228 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232 Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235 Appendix I. Split ends is a tissue/promoter specific regulator of Wingless signaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240 Drosophila strains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240 Isolation of new spen alleles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242 Whole-mount staining and microscopy . . . . . . . . . . . . . . . . . . . . . . . . . 242 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243 Mutations in spen dominantly suppress a sensitized Wg phenotype in the eye . . . . . 243 spen is required for Wg signaling in the wing and leg . . . . . . . . . . . . . . . . . 249 spen acts downstream of Arm stabilization . . . . . . . . . . . . . . . . . . . . . . . 253 spen is not required for Wg signaling in the embryo . . . . . . . . . . . . . . . . . . 255 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256 Spen is required for normal Wg signaling in imaginal discs . . . . . . . . . . . . . . 256 Spen is not required for embryonic Wg signaling . . . . . . . . . . . . . . . . . . . . 258 Spen may have a redundant partner . . . . . . . . . . . . . . . . . . . . . . . . . . 259 Mechanism of Spen action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262 Appendix II. Further insights into Spen’s requirement during Drosophila eye development . . . . . . . . . . . . . . . . . . . . . . . . . . 267 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268 spen mutant clones lack Spen protein . . . . . . . . . . . . . . . . . . . . . . . . . . 268 spen is required for proper adult eye patterning . . . . . . . . . . . . . . . . . . . . 268 Loss of spen leads to reduced dpERK and elevated Yan but not yan transcript . . . . 272 Elevated Delta and Rhomboid observed in spen mutant clones . . . . . . . . . . . . 277 Transcript level analysis in spen mutants: Notch, ato, rho, and MAPK phosphatases . 277 Increased cell adhesion observed with loss of spen . . . . . . . . . . . . . . . . . . . 278 Spen does not affect cell cycle regulation . . . . . . . . . . . . . . . . . . . . . . . . 283 Spen and Programmed Cell Death . . . . . . . . . . . . . . . . . . . . . . . . . . . 285 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285 Fly Strains and Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285 Immunohistochemistry and Immunofluorescence . . . . . . . . . . . . . . . . . . . . 286 Western Blot Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287 In Situ Hybridization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287 Quantitative RT-PCR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290 Appendix III. Characterization of dominant-negative Spen: in vivo and in vitro . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294 spenDN expression in CNS midline glial cells acts dominant-negatively . . . . . . . . 294 A candidate approach to identify dominant modifiers of the sev-spenDN rough-eye phenotype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295 SpenDN may regulate Notch pathway targets in Drosophila cell culture . . . . . . . . 300 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301 Fly Strains and Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301 Immunohistochemistry and Histology . . . . . . . . . . . . . . . . . . . . . . . . . . 302 Transcription Assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303 10 Appendix IV. Phenotypic analysis of Spen over-expression . . . . . 305 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306 A spen EP line that allows for Spen over-expression . . . . . . . . . . . . . . . . . . 306 Eye-specific expression of the spen EP disrupts eye development . . . . . . . . . . . 307 Spen over-expression leads to reduced Yan expression and ectopic BarH1 expression 308 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312 Fly Strains and Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312 Histology, Immunohistochemistry and Immunofluorescence . . . . . . . . . . . . . . 312 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313 Appendix V. Characterization of dominant-negative Spen∆C during Drosophila eye development and in Drosophila cultured cells . . . . . . . . . . . . . . . . . . . . . . . . . . . 315 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316 Over-expression of SpenDC results in elevated Yan protein levels . . . . . . . . . . . . 316 SpenDC does not regulate Yan at the post-transcriptional level via the yan 3’UTR . . . 316 Spen affects activation of Ets targets . . . . . . . . . . . . . . . . . . . . . . . . . . 318 Spen does not affect Yan nuclear export mediated by activated RasV12 in S2 cells . . . 321 spen RNAi knock-down does not affect Yan protein levels but dpERK levels are elevated . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323 Fly Strains and Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323 Luciferase-UTR/Transcription Assays . . . . . . . . . . . . . . . . . . . . . . . . . 324 Immunohistochemistry and Immunofluorescence . . . . . . . . . . . . . . . . . . . . 324 RNAi interference and semi-quantitative RT-PCR . . . . . . . . . . . . . . . . . . . 324 Yan Nuclear Export Assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325 Western Blot Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326 11 Abbreviations and Acronyms Abb/Acr Ac ACT ADAM AMKL AML Ank Aop Aos AP Arm AS-C Ase Ato ban E-cat E-GAL bHLH Brd Brm Casp-3* CBF ChIP Chn CIR CKID CNS Co-Act co-IP Co-R CRD CRYBP1 CtBP CtIP CycA/B/E CyO da dArk DE-Cad DER Df DIO Dl Dll dll dMkp3 DN Dp Dpp Drk Dsh Dsor1 dsRNA DV E(spl) EBS 12 full Achaete activated A Disintegrin And Metalloproteinase Acute megakaryocytic leukemia acute myeloid leukemia Ankyrin Anterior open Argos anterior-posterior Armadillo Achaete-Scute complex Asense Atonal bantam E-catenin E-Galactosidase basic helix-loop-helix Bearded Brahma cleaved, activated Caspase-3 C-promoter binding factor chromatin immunoprecipitation Charlatan CBF-interacting repressor Casein Kinase ID central nervous system co-activators co-immunoprecipitation co-repressors Cys-rich domain crystallin binding protein 1 C-terminal binding protein CtBP-interacting protein Cyclin A/B/E Curly O daughterless Apaf-1 related killer Drosophila E-Cadherin Drosophila Epidermal Growth Factor Receptor Deficiency Death inducer-obligator Delta Delta-like distalless Drosophila MAPK phosphatase 3 dominant-negative Duplication Decapentaplegic Downstream of receptor kinase Disheveled Dowstream of Raf1 double-stranded RNA dorsal-ventral Enhancer of split Ets-binding site Abb/Acr EC ECD ed EGF EGFR Elp Emc EMSA en EP Esdn Ets eve EXT Ey Eya eyg Faf FGF FLP Fng FRT Fz GAP GFP GMR Grk Gro grp H h HAT HDAC HES Hh hid Hox hs HSC HTH ICD Ig IG intra IPCs Iswi Jag Klu Krn Kul Kuz lch LNR Lqf L'sc full Extracellular Extracellular domain echinoid Epidermal Growth Factor Epidermal Growth Factor Receptor Ellipse Extramacrochaetae Electrophoretic mobility shift assay engrailed Enhancer-Promoter Enhancer of spen[DN] E twenty-six even-skipped Extracellular truncation Eyeless Eyes absent eyegone Fat facets Fibroblast growth factor FLP recombinase Fringe FLPase recombination target Frizzled GTPase activating protein green fluorescent protein glass multimerized repeat Gurken Groucho grapes Hairless hairy Histone acetyltransferase Histone decacetylase Human enhancer of split Hedgehog head involution defective homeobox heat-shock hematopoietic stem cell helix-turn-helix intracellular domain Immunoglobulin intermediate group intracellular interommatidial precursor cells Imitation Switch Jagged Klumpfuss Keren Kuzbanian-like Kuzbanian lateral chordotonal LIN-12/Notch Repeats Liquid Facets Lethal of scute Abb/Acr luci Lz Mae Mal MAPK MEK MF MGC Mib Mint miR Mkl Mor Msk N NCoR neo Neur Nito NLS NRSF NuRD OC OCFRE Ofut Ott Pak PCD PCP PEST PH3 Phyl PKA PNS PntP1/P2 PPN prd Pros Psq PTB ptc PTP-ER pygo RBD Rbm RBP-J RDGN REM Rho RID RNAi Ro rpr RRM RTE RTK full luciferase Lozenge Modulator of the activity of Ets Megakaryocytic Acute Leukemia Mitogen Activated Protein Kinase MAP/ERK Kinase morphogenetic furrow midline glial cell Mind bomb Msx2-interacting nuclear target micro-RNA Megakaryoblastic leukemia Moira Moleskin Notch Nuclear receptor co-repressor neomycin Neuralized Spenito nuclear localization signal Neuronal restricted silencing nucleosome remodeling and histone deacetylase osteocalcin osterocalcin FGF response element O-fucosyltransferase One Twenty-Two p2-activated kinase programmed cell death planar cell polarity Pro-Glu-Ser-Thr phosphorylated Histone H3 Phyllopod Protein Kinase A peripheral nervous system PointedP1/P2 pre-proneural paired Prospero Pipsqueak phosphotyrosine binding patched Protein tyrosine phosphatase-ERK/enhancer of Ras1 pygopus Ras-binding domain RNA-binding protein motif Recombination signal sequence-binding protein J retinal determination gene network Ras exchanger motif Rhomboid receptor interaction domain RNA interference Rough reaper RNA-recognition motif RNA transport elements Receptor Tyrosine Kinase Abb/Acr S S2 Sal SAM Sc Sca SCS SEM Sens Ser Sev Sgg SH2/3 Sharp Sim Skip SME Smr SMRT SMW Sno So SOP Sos Spen Spi spl SPOC SRA SRC- Ssdn Stg STS Su(H) TACE TAD TAF0 TCF TGF-D TGF-E TM TM6B tou Toy ts Ttk88 Tub twi UAS ubi Upd vg Vn VP16 wg Zw3 full Star Schneider 2 Spalt Sterile D motif Scute Scabrous sev-spen[DN] on CyO scanning electron microscopy Senseless Serrate Sevenless Shaggy Src-homology 2/3 SMRT/HDAC1 associated repressor protein Single-minded Ski-interacting protein Lz[sma] minimal enhancer Smrter Silencing mediator for retinoid and thyroid receptor second mitotic wave Strawberry notch Sine oculis sensory organ precursor Son of sevenless Split ends Spitz split Spen paralog and ortholog C-terminus Steroid receptor activator Steroid receptor co-activator Suppressor of spen[DN] String sev-spen[DN] on TM6B Suppressor of Hairless Tumor Necrosis Factor-a Converting Enzyme Trasncriptional Activation Domain TATA-binding protein-associated factor 0 T-cell factor Transforming Growth Factor-D Tissue Growth Factor-E transmembrane Third multiple 6B toutatis Twin of eyeless temperature sensitive Tramtrack 88 Tubulin twist Upstream activating sequence ubiquitin Unpaired vestigial Vein Herpes simplex Virus Protein 16 wingless Zeste white 3 13 14 Chapter 1. Signal integration during development: mechanisms of EGFR and Notch pathway function and cross-talk David B. Doroquez and Ilaria Rebay Critical Reviews in Biochemistry and Molecular Biology. 2006. 41(6): 339-85 15 Abstract Metazoan development relies on a highly regulated network of interactions between conserved signal transduction pathways to coordinate all aspects of cell fate specification, differentiation and growth. In this review, we discuss the intricate interplay between the Epidermal Growth Factor Receptor (EGFR; Drosophila EGFR/DER) and the Notch signaling pathways as a paradigm for signal integration during development. First, we describe the current state of understanding of the molecular architecture of the EGFR and Notch signaling pathways that has resulted from synergistic studies in vertebrate, invertebrate and cultured cell model systems. Then, focusing specifically on the Drosophila eye, we discuss how cooperative, sequential, and antagonistic relationships between these pathways mediate the spatially- and temporallyregulated processes that generate this sensory organ. The common themes underlying the coordination of the EGFR and Notch pathways appear to be broadly conserved and should therefore be directly applicable to elucidating mechanisms of information integration and signaling specificity in vertebrate systems. Introduction Development in metazoans provides a fruitful backdrop in which to study how a limited collection of signaling networks accurately regulates a diverse array of cellular events including cell fate specification, morphogenesis, proliferation, differentiation, polarity establishment, programmed cell death, and motility. Somehow, the spatial and temporal mechanisms that coordinate the function of these pathways must impart specificity. A major part of the answer, which we are just beginning to decipher, is that rather than functioning as completely independent and insulated modules, signaling pathways interface in intricate ways to create a web of specific interactions that the cell integrates and interprets in a spatially and temporally appropriate manner (Shilo, 2005; Sundaram and Han, 1996; Sundaram, 2005; Tan and Kim, 1999; Voas and Rebay, 2004). The evolutionarily conserved Epidermal Growth Factor Receptor (EGFR) pathway and 16 the Notch pathway represent two such signal transduction mechanisms. Recent studies have emphasized that the interaction between the EGFR and Notch pathways is intimate and critical during developmental processes in metazoa (Shilo, 2005; Sundaram, 2005). The reports of EGFR and Notch signaling interactions are storied, but the molecular mechanisms underlying the cross-talk and signal integration between these two pathways are just beginning to be revealed (Sundaram, 2005). In the following review, our goal is to compare and contrast the various types of relationships reported between the EGFR and Notch signaling pathway in Drosophila melanogaster. In this system, EGFR and Notch may interact antagonistically or cooperatively over specific time and space. First, incorporating results from both genetic and biochemical analysis, we will describe the biochemical and molecular interactions involved within the individual EGFR and Notch signaling pathways. Then, we focus on the interaction of these pathways in the context of progressive development of the Drosophila eye, a well-characterized, genetically tractable-system to study cell-cell signaling integration and cross-talk. The Epidermal Growth Factor Receptor Pathway The EGFR signaling pathway was described initially through cooperative genetic analyses in Drosophila melanogaster and Caenorhabditis elegans and biochemical studies in mammalian cell culture (Perrimon and Perkins, 1997; Shilo, 2005; Voas and Rebay, 2004; Wassarman et al., 1995). In Drosophila, the EGFR pathway is used reiteratively throughout development, contributing to processes as diverse as adult abdominal dorsoventral patterning, wing vein determination and oogenesis (Shilo, 2003; Shilo, 2005). In the following sections, we review the signaling machinery that drives the EGFR pathway in Drosophila. In particular, we highlight the mechanisms that convert extracellular signals into transcriptional responses. It is important to note that in presenting a broad overview of EGFR signaling, we have incorporated information from multiple contexts and have down-played the tissue-to-tissue variations in the signaling machinery. 17 DER Ligands Signal transduction mechanisms require that molecular cues initiated at the cell surface be relayed to the nucleus. In the EGFR signaling pathway, these signals take the form of small protein ligands. There are at least eight different EGFR ligands in mammals, which include EGF, Transforming Growth Factor-α (TGF-α), and the Neuregulins (Riese and Stern, 1998). EGFR ligands are synthesized as transmembrane precursors where they either function as membrane-anchored proteins for juxtacrine signaling or are cleaved proteolytically to release a soluble ligand (Riese and Stern, 1998). Four activating ligands and one inhibitory ligand have been identified for the Drosophila Epidermal Growth Factor Receptor (DER) (Fig. 1-1A) (Shilo, 2003): (1) the Neuregulin Vein; (2) the TGF-α ligands Spitz, Keren and Gurken; and (3) the repressive ligand Argos. The Neuregulin Vein Ligand. Vein (Vn), a Neuregulin-class secreted ligand with a weak activation capacity, is utilized in tissues that require low EGFR activation (Schnepp et al., 1996; Shilo, 2003). Neuregulins have an immunoglobulin-like (Ig) domain, multiple potential glycosylation sites, and a cysteine-rich EGF domain (Garratt et al., 2000). The Ig domain of Vn may facilitate protein-protein interaction through dimerization or interactions with other extracellular proteins (Fig. 1-1A) (Schnepp et al., 1996). Developmentally, Vn is necessary during embryogenesis for the patterning of the ventral ectoderm and specification of muscle precursors (Shilo, 2003). The role of Vn during eye development is less clear. Although it is not required genetically, Vn may stimulate DER signaling in the contexts of cell proliferation and survival, and possibly differentiation (Spencer et al., 1998). Because Vn does not appear to play a major role in the eye (Shilo, 2003), it will not be considered further in this review. Rather, the ensuing discussion will focus on Spi, the DER ligand that has a key role during eye development. The TGF-α Family of DER Ligands. The primary DER ligand, Spi, is responsible for activation of the DER pathway in most tissues during Drosophila development, beginning in oogenesis and continuing through adult male spermatogenesis (Shilo, 2003). Spi, like the other 18 Fig. 1-1. The Drosophila Epidermal Growth Factor (EGF) Receptor (DER) and its Ligands. (A) DER has five ligands. Spi, Grk, and Krn are cleaved for secretion at their transmembrane domain (TMD; arrow). (B) The two different isoforms of DER differ at their Ntermini. See text for domain details. Adapted from Shilo, 2003. 19 two TGF-α family ligands, Keren (Krn) and Gurken (Grk), is generated as a transmembrane precursor with an N-terminal signal peptide, an EGF domain in the extracellular region, a transmembrane domain and a cytosolic C-terminus (Fig. 1-1A) (Kumar et al., 1995; NeumanSilberberg and Schupbach, 1993; Reich and Shilo, 2002; Rutledge et al., 1992). In the past several years, the mechanisms by which Spi is processed have begun to be elucidated (Klambt, 2000; Klambt, 2002; Shilo, 2003). Based on strong genetic interactions with DER pathway components, Rhomboid (Rho) and Star (S), which encode novel transmembrane proteins, were initially proposed to facilitate DER signaling via cell-surface processing and presentation of the Spi ligand (Sturtevant et al., 1993). More recently, Rho and Star were found to mediate processing of Spi within the cell (Fig. 1-2) (Bang and Kintner, 2000; Lee et al., 2001; Tsruya et al., 2002; Urban et al., 2001). Rho encodes a seven-transmembrane protein, while Star encodes a Type II transmembrane protein; each has a cytoplasmic N-terminus and a lumenal C-terminus (Bier et al., 1990; Kolodkin et al., 1994; Lee et al., 2001). Biochemical and cell culture experiments showed that Star colocalizes with Spi, such that Star translocates the Spi precursor protein from the Endoplasmic Reticulum (ER) to the Golgi apparatus by either blocking an ER retention signal in Spi or actively exporting Spi to the Golgi (Fig. 1-2) (Lee et al., 2001). The lumenal domain of Star mediates this trafficking through interaction with the lumenal domain of the Spi precursor (Lee et al., 2001). Following translocation, Spi and Star sequentially bind Rho (Tsruya et al., 2002) and Spi is then cleaved within its transmembrane domain in a Rho dependent manner (Fig. 1-2) (Lee et al., 2001; Tsruya et al., 2002; Urban and Freeman, 2002; Urban et al., 2001). Once cleaved, the mature Spi ligand is transported from the Golgi and out of the cell to bind its cognate receptor. Rho encodes a novel intramembrane serine protease whose catalytic activity is required for Spi cleavage at the motif ASIASGA (Lee et al., 2001; Urban and Freeman, 2003; Urban et al., 2001). Rho is specifically sensitive to serine protease inhibitors, supporting its placement within the serine protease superfamily (Urban et al., 2001). Rho is well-conserved across species, indicating that other Rhomboid proteins may also have proteolytic function (Urban and 20 Fig. 1-2. Spi ligand processing, DER activation, and Spi-inhibition by Aos. (A) Membranebound Spitz (mSpi) is transported to the Golgi by Star, (B) cleaved by Rhomboid (Rho) and (C) secreted, where it (D) activates DER or (E) is sequestered by Aos, preventing DER activation. Adapted from Shilo, 2003. Freeman, 2002; Urban et al., 2001; Urban et al., 2002b). Indeed, the other three Drosophila Rhomboids (Rho2-4) are functional proteases, although unlike Rho, they do not require Starmediated translocation for their function (Urban et al., 2002a). The existence of multiple Rhomboids suggests there may be tissue-specific regulation of proteolytic activation of DER ligands. Interestingly, the other DER ligands Grk and Krn are also activated by Star-mediated translocation and Rho proteolytic cleavage (Reich and Shilo, 2002; Urban et al., 2002a). However in vertebrates, Rho-mediated cleavage of TGF-α ligands appears not to be the predominant mechanism of EGFR activation. For example, although the vertebrate Rhomboid 21 p100hRho can associate with TGF-α ligands in cultured mammalian cells, the relevance of the interaction to TGF-α processing in vivo remains unclear (Nakagawa et al., 2005). Rather, mounting evidence suggests that other proteases, such as the Tumor Necrosis Factor-α Converting Enzyme (TACE)/A Disintegrin And Metalloproteinase (ADAM) 17, mediate TGF-α cleavage and ectodomain secretion (Black, 2002; Hinkle et al., 2003; Hinkle et al., 2004; Juanes et al., 2005; Killar et al., 1999; Lee et al., 2003; Merlos-Suarez et al., 2001; Peschon et al., 1998; Sunnarborg et al., 2002). In summary, the ER-Golgi translocation and proteolytic cleavage of Spi by the transmembrane proteins Star and Rho, while obligate in Drosophila, may reflect a specialized mechanism of EGFR ligand processing and presentation. The Repressive ‘Ligand’ Argos. DER signal transduction is negatively regulated at the ligand-receptor level. The gene that encodes the DER regulator protein Argos (Aos) was initially identified based on lethal mutations that disrupted development of the Drosophila eye (Freeman et al., 1992). Molecular cloning revealed that aos encodes a novel secreted protein with a cysteine-rich region similar to an EGF repeat (Fig. 1-1A) (Freeman et al., 1992). Genetic analysis and cell culture experiments led to the hypothesis and eventual demonstration that this protein functions antagonistically to the DER pathway (Freeman, 1994a; Freeman et al., 1992; Sawamoto et al., 1996; Sawamoto et al., 1994; Schweitzer et al., 1995). Interestingly, high levels of DER signaling potentiate aos expression, suggesting the importance of a negative feedback loop to attenuate high levels of signaling output (Golembo et al., 1996; Wasserman and Freeman, 1998). Structure-function analysis shows that the Aos atypical EGF domain, the N-terminal EGF-flanking regions, and the C-terminus, are all critical for Aos’ inhibitory ability (Fig. 11A) (Howes et al., 1998). In domain-swap experiments in which the EGF domain of the DER activator ligands Vn or Spi was replaced with the Aos atypical EGF domain, the Aos EGF domain was sufficient to convert Vn or Spi into an inhibitor (Jin et al., 2000; Schnepp et al., 1998). These experiments show that the Aos EGF domain is essential for abrogation of DER signaling. 22 There are conflicting reports that describe the mechanism by which Aos represses DER signaling. Initial experiments in Drosophila S2 cultured cells exposed to the activating ligand Spi demonstrated that Aos expression reduced DER phosphorylation levels (Schweitzer et al., 1995), indicating that Spi and Aos are mutually antagonistic in their regulation of DER. However, in the presence of overexpressed DER, expression of Aos was sufficient to repress DER activation, suggesting Aos may act independent of Spi (Schweitzer et al., 1995). Subsequent co-immunoprecipitation (co-IP) experiments showed that DER and Aos could interact and that the extracellular domain of DER and the EGF/C-terminal region of Aos were necessary both for the interaction and for DER repression (Jin et al., 2000; Vinos and Freeman, 2000). Cross-linking experiments also suggested that Aos-DER interaction prevented DER dimerization, thereby blocking activation (Jin et al., 2000). Further, when DER and Spi were expressed along with increasing amounts of Aos, co-IP of DER and Spi was reduced. Thus, it was concluded that Aos competes with Spi for binding to DER (Jin et al., 2000). Recently, the model that arose from the above Aos/Spi/DER analysis has been challenged (Fig. 1-2E). Biochemical studies using surface plasmon resonance (SPR), analytical ultracentrifugation, and cell culture experiments demonstrated that Aos did not bind DER to inhibit signaling. Rather, Aos bound to Spi with a 1:1 stoichiometry and sequestered Spi from DER, suggesting Aos functions as a ‘ligand sink’ to bind soluble Spi (Klein et al., 2004). In contrast to an earlier report, Aos did not disrupt bound Spi-DER complexes, according to orderof-addition experiments performed in cell culture. However, formation of Spi-Aos complexes prior to addition or expression of DER, led to the failure of Spi to bind receptor (Klein et al., 2004; Vinos and Freeman, 2000). The model of ligand sequestration by Aos follows with in vivo observations, where Aos does not function over a long distance, but acts at a short range to negatively regulate Spitz activity (Fig. 1-2E) (Klein et al., 2004; Reeves et al., 2005). In conclusion, the secreted protein Aos inhibits DER signaling at the ligand-receptor level, not by acting directly as an antagonistic ligand, but likely by sequestering activating-ligand away from the DER receptor (Fig. 1-2E). 23 DER: The Drosophila EGF Receptor EGFR/ErbB proteins belong to a subset of the Receptor Tyrosine Kinase (RTK) superfamily that serves as receptors for growth factors, differentiation factors, and metabolic stimulation factors (van der Geer et al., 1994). EGFR/ErbB, was identified as a viral oncogene (v-ErbB) in the genome of the avian erythroblastosis virus (AEV), transduced from the chicken genome’s Egfr gene, c-ErbB (Downward et al., 1984; Roussel et al., 1979; Vennstrom and Bishop, 1982). Since its early characterization in oncogenesis, EGFR signaling was found to play fundamental roles in proliferation, differentiation, and development (Olayioye et al., 2000). The Drosophila genome encodes a single locus of the EGFR/ErbB family: Drosophila Epidermal growth factor Receptor (DER) (Livneh et al., 1985). C. elegans also possesses a single EGFR: LET-23 (Aroian et al., 1990). Humans, on the other hand, have four EGFRs: ErbB1/EGFR/HER1, HER2/Neu, ErbB3/HER3, and ErbB4/HER4 (Klapper et al., 2000; Olayioye et al., 2000). DER was isolated and cloned in a screen to identify homologs of chicken v-ErbB in the Drosophila genome (Livneh et al., 1985). The DER locus encodes alternativelyspliced isoforms whose encoded proteins (isoform I and II) differ at their amino terminus (Fig. 1-1B) (Schejter et al., 1986). These two isoforms may have distinct functions. For example, a constitutively-activated form of isoform I, but not of isoform II, was shown to be sufficient to induce ectopic differentiation in the developing Drosophila eye (Kumar and Moses, 2001b). The nature of the difference between the two isoforms is still unclear and remains to be elucidated. DER, like other RTKs, possesses three domains that facilitate intracellular transduction of signals initiated at the cell surface: a glycosylated extracellular ligand-binding domain, a single hydrophobic Type I transmembrane domain, and a cytoplasmic tyrosine kinase catalytic domain (Fig. 1-1B) (Bazley and Gullick, 2005; Ullrich and Schlessinger, 1990; van der Geer et al., 1994). DER belongs to Subclass I of the RTK family, characterized by a pair of ligand bindingdomains (L1 and L2) and Cys-rich repeat sequences (C1 and C2) located in its extracellular domain (Ullrich and Schlessinger, 1990). While conservation in the kinase and ligand binding 24 domains is too strong to allow classification of DER within the EGFR family, comparision of the Asn-linked glycosylation sites indicates DER is more similar to HER1 than to Neu (Livneh et al., 1985; Schejter et al., 1986). Specifically, DER shares twelve Asn with HER1, but only nine with Neu (Schejter et al., 1986). These sites may prove important for function, because N-linked glycosylation was reported to be essential for proper EGFR plasma membrane translocation, dimerization, ligand-binding specificity, and activation (Olson and Lane, 1989; Slieker and Lane, 1985; Soderquist and Carpenter, 1984; Tsuda et al., 2000; Whitson et al., 2005). It is interesting that Drosophila and C. elegans each only have a single EGFR. In mammals, EGFRs are not only able to homodimerize but also heterodimerize with other EGFRs (Schlessinger, 2000). Heterodimerization of receptors permits receptor diversity and specificity of function, signaling level, and response (Schlessinger, 2000). For example, in some contexts, heterodimerization may result in stronger signaling than achieved from homodimeric receptor pairs (Moghal and Sternberg, 1999). This begs the question of whether or not this type of diversity is required in Drosophila or C. elegans. If variation is necessary, then diversity at the receptor level may occur with variable N-linked glycosylation, with a repertoire of ligands, or with the differential localization or targeting of the receptor (Moghal and Sternberg, 1999; Whitson et al., 2005). The stoichiometry of ligand-to-receptor binding is varied across classes of receptors. As a member of the RTK Subclass I, each DER protein binds to a monomeric ligand (Ullrich and Schlessinger, 1990). Binding of ligand to the DER extracellular domain stimulates conformational alteration and receptor dimerization (Fig. 1-3B) (Schlessinger, 1988). According to biophysical studies, binding of monomeric ligand with monomeric receptor forms a 1:1 EGF: EGFR dimer that facilitates heterotetramerization. That is, a 2:2 EGF:EGFR ratio is required for activation of the cytoplasmic domain (Lemmon et al., 1997). Dimerization stabilizes protein interactions, promotes juxtaposition of the cytoplasmic domains, and induces a conformational change that leads to activation of its kinase function (Fig. 1-3B). The activated kinase domain then leads to trans- and auto-phosphorylation of tyrosine residues on the cytoplasmic tail of the 25 receptor (Schlessinger, 1988; van der Geer et al., 1994; Zhang et al., 2006). Activation of the DER dimer facilitates its ability to catalyze phosphorylation of cytoplasmic substrates and/or to recruit phosphotyrosine binding (PTB) or Src-homology 2 (SH2) domain-containing proteins that have increased affinity for activated receptor (Fig. 1-3B) (Mohammadi et al., 1993; Pawson and Scott, 1997; Sherrill, 1997; van der Geer et al., 1994). In summary, DER phosphorylation following ligand reception initiates nucleation of the subsequent cytosolic signaling complexes. Activation of Ras The primary consequence of EGFR activation in Drosophila is assembly of a membrane- localized protein complex that recruits and activates Ras (Fig. 1-3B). Specific tyrosine phosphorylation of DER promotes docking of the cytoplasmic adaptor protein Downstream of Receptor Kinase (Drk), the homolog of mammalian Grb2 and C. elegans SEM-5, via its single SH2 domain (Fig. 1-3B). Drk converts this phosphotyrosine input to a signaling output by serving as the cornerstone for the protein complex that ultimately recruits and activates Ras (Pawson et al., 2001). Drk is a flexible protein with two Src-homology 3 (SH3) domains flanking the phosphotyrosine-binding SH2 domain (Olivier et al., 1993; Yuzawa et al., 2001). These SH3 domains bind to proline-rich motifs in a separate set of proteins, thus building a protein complex around activated receptors (Pawson and Scott, 1997; Ren et al., 1993). Drk complexes constitutively with the Guanine nucleotide exchange factor (GEF) Son of sevenless (Sos) by binding to its C-terminal Pro-rich region (Fig. 1-3B) (Buday and Downward, 1993; Hunter, 2000; Olivier et al., 1993; Schlessinger, 1993; Simon et al., 1991). Recruitment of the Drk-Sos complex to activated DER promotes the association of Sos with Ras (Fig. 1-3B). Ras is a proto-oncogene localized to the cytoplasmic side of the plasma membrane by farnesylation and palmitoylation (Magee and Marshall, 1999; Neuman-Silberberg et al., 1984; Simon et al., 1991). It serves as a guanine nucleotide-binding protein that cycles between inactive GDP-containing and active GTP-containing conformations (Fig. 1-3). GEFs such as Son-of-sevenless (Sos) facilitate the exchange of GDP to GTP, thereby activating Ras. Providing 26 Fig. 1-3. The DER signaling pathway. (A) In non-stimulated cells, Ras exists in a GDPbound, inactive state. 14-3-3 binds to phosphorylated Raf and Ksr on two Ser residues, retaining each in the cytoplasm. (B) DER activation by Spi leads to its activated, GTP-bound state. PP2A dephosphorylates Raf and Ksr and displaces 14-3-3. This permits Raf and Ksr to re-localize with Ras near the membrane to activate the kinase cascade. Adapted from Raabe and Rapp, 2003. 27 balance, GTPase activating proteins (GAPs) promote the hydrolysis of GTP to GDP by Ras, thereby returning it to the inactive conformation (McCormick, 1994). This biochemical cycle allows Ras to operate as an ‘on/off ’ switch. Amino acid substitutions, like the Gly12Val (RasV12) mutation, that stabilize the GTP-bound state, create a constitutive ‘on’ state (Barbacid, 1987). In vivo genetic studies have taken advantage of targeted transgenic expression of activated RasV12 to dissect the mechanisms of signaling flow through the pathway in many different developmental contexts (Firth et al., 2005). Structural studies demonstrate that Ras conformation is dynamic, particularly in two loop regions known as Switch 1 and Switch 2. Sos tightly binds the Switch 2 region of Ras, induces conformational changes, and activates Ras by opening up a nucleotide-binding site in the Switch 1 region, facilitating the exchange of guanine nucleotides (Boriack-Sjodin et al., 1998; Hall et al., 2001). Following nucleotide exchange, Sos is released from Ras. Interestingly, it was demonstrated that the process of nucleotide exchange is potentiated through positive-feedback mechanisms whereby allosteric binding of Ras-GTP to the Ras exchanger motif (REM) of Sos forms a ternary complex and increases Sos-mediated nucleotide exchange of a separate Ras protein (Margarit et al., 2003). Activation of the Raf/MEK/MAPK Cascade The activation of Ras switches ‘on’ a series, or cascade, of protein kinases, such that as one kinase is phosphorylated, it subsequently phosphorylates the next. This culminates in phosphorylation and activation of the Mitogen Activated Protein Kinase (MAPK) (Fig. 13B). The kinase cascade is initiated as activated Ras binds to the proto-oncogene Raf with high-affinity, potentiating its catalytic activity as a serine/threonine kinase (Fig. 1-3B) (Chong et al., 2003). Raf contains several conserved domains: a Ras-binding domain (RBD), a Cysrich domain (CRD), a Ser/Thr-rich region, and a catalytic kinase domain (Chong et al., 2003). The RBD is required for translocation of Raf from the cytosol to the cell membrane where it associates with Ras. The CRD, which encodes a Zinc-finger domain, is required for activation 28 of Raf and also plays a role in binding to Ras (Avruch et al., 2001). Phosphorylation of Raf at the Ser/Thr-region and the catalytic domain may subject Raf to autoregulation or regulation by separate kinases; certain phosphorylation events may promote Raf function while others may antagonize its function (Avruch et al., 2001; Chong et al., 2003; Kolch, 2000). Physical interaction with Ras is hypothesized to elicit an activated Raf transition state by relieving intramolecular inhibitory regulation (Fig. 1-3A) (Avruch et al., 2001; Kolch, 2000). In resting cells, Raf exists in an inactive state such that its C-terminus negatively inhibits the kinase domain. This is mediated through serine phosphorylation by Protein Kinase A (PKA) and phospho-binding-protein 14-3-3 interaction (Fig. 1-3A) (Avruch et al., 2001; Dhillon et al., 2002). Activation of signaling relieves autoinhibition through Protein Phosphatase 2A (PP2A) mediated dephosphorylation, which displaces 14-3-3 and allows the RBD and CRD of Raf to interact with Ras near the plasma membrane (Fig. 1-3B) (Jaumot and Hancock, 2001). Raf is then phosphorylated at a separate serine residue by the Pak3 kinase. This prevents reassociation and autoinhibition by allowing 14-3-3 to take up residence at these newly phosphorylated sites to maintain stably-active Raf (Fig. 1-3B) (Avruch et al., 2001). Following activation, Raf binds via its kinase domain to the MAP/ERK Kinase (MEK)/ Downstream of Raf1 (Dsor1) and activates MEK by dual phosphorylation of Ser or Thr residues within an activation loop (Fig. 1-3B) (Crews et al., 1992; Crews and Erikson, 1992; Huang et al., 1993; Kolch, 2005; Macdonald et al., 1993; Pearson et al., 2001; Seger and Krebs, 1995; Tsuda et al., 1993). MEK is a dual-specificity kinase able to phosphorylate Ser, Thr, as well as Tyr residues (Huang et al., 1993; Pearson et al., 2001; Seger and Krebs, 1995). Activated MEK binds with its N-terminal docking site to MAPK and phosphorylates it first on Tyr and then on Thr residues located within MAPK’s flexible activation loop (Fig. 1-3B) (Pearson et al., 2001; Tanoue and Nishida, 2003; Zhang et al., 1994; Zhang et al., 1995). Dual phosphorylation is necessary for the activity of MAPK (Pearson et al., 2001). How is the activation and function of such a three-kinase signaling module regulated and coordinated? Recent studies have demonstrated the importance of scaffolding in this context. 29 Scaffolding of the MAPK cascade is mediated, at least in part, by the Kinase Suppressor of Ras (Ksr) protein, first isolated in a Drosophila genetic screen as an effector of Ras signaling (Fig. 13B) (Therrien et al., 1995; Therrien et al., 1996). Ksr facilitates assembly and activity of the Raf, MEK, and MAPK kinase cascade by mediating the individual protein-protein interactions and by localizing the complex to an appropriate subcellular site (Fig. 1-3) (Raabe and Rapp, 2003; Schaeffer and Weber, 1999). The complex choreography underlying these events remains an area of active research and its further discussion is beyond the scope of this review (Kolch, 2005; Morrison and Davis, 2003; Raabe and Rapp, 2003). Regulation of Nuclear Targets via Activated MAPK MAPK completes the transfer of information from cell surface to the nucleus by relocalizing from the cytoplasm to the nucleus upon activation. Translocation of activated, diphosphorylated MAPK (dpMAPK) into the nucleus may occur by multiple means including passive diffusion through the nuclear pore, interaction with the nuclear pore complex, association and co-shuttling with other nuclear-targeted proteins, or direct importin-mediated transport (Baker et al., 2002; Cyert, 2001; Khokhlatchev et al., 1998; Kondoh et al., 2005; Lorenzen et al., 2001). In Drosophila, Moleskin (Msk) and Ketel, the homologs of Importin-7 and Importin-β, respectively, are required for nuclear import of dpMAPK (Lorenzen et al., 2001). Recently, it was identified that Msk developmentally modulates the hold of dpMAPK in the cytoplasm or movement of dpMAPK into the nucleus (Fig. 1-4B) (Marenda et al., 2006; Vrailas et al., 2006; Vrailas and Moses, 2006). Thus, regulation of dpMAPK nuclear import appears critical in affecting output from the DER signaling pathway. Nuclear dpMAPK phosphorylates specific target proteins, including two Drosophila Ets transcription factors relevant to our discussion of DER signaling: Pointed-P2 (PntP2) and Yan/Anterior open (Aop) (Fig. 1-4B) (O'Neill et al., 1994). PntP2 is the homolog of the human Ets-1 transcription factor and functions as a transcriptional activator when phosphorylated by dpMAPK (Fig. 1-4B) (Watson et al., 1985). Yan is the homolog of the human Tel1 protein and represses target genes in the absence of 30 dpMAPK (Golub et al., 1994). Both PntP2 and Yan contain an Ets DNA-binding domain and compete for access to promoter regions of common downstream transcriptional targets, with level and duration of activity of these two key DER pathway effectors determining the net response to pathway stimulation. Work from multiple labs, combining both in vitro and in vivo experimental approaches, has led to a model in which Yan protein localization is dynamically regulated by protein-protein associations, phosphorylation, nuclear export and stability (Fig. 1-4) (Qiao et al., 2004; Rebay and Rubin, 1995; Song et al., 2005; Tootle et al., 2003). At its N-terminus, Yan contains a Sterile α Motif (SAM), found in a subset of Ets proteins, which mediates homotypic and heterotypic protein-protein interactions and is required for transcriptional repression. In vitro, Yan selfassociates to form head-to-tail polymers through its SAM domain (Fig. 1-4A) (Qiao et al., 2004). These polymers can be disrupted upon binding to Modulator of the activity of Ets (Mae)/Ets domain lacking (Edl), a SAM domain-containing protein that lacks an Ets DNA-binding domain (Fig. 1-4) (Baker et al., 2001; Qiao et al., 2004; Song et al., 2005; Yamada et al., 2003). Although the polymerization model awaits in vivo validation, the prediction is that in unstimulated cells, Yan polymers would be favored in an equilibrium with depolymerized Yan monomers bound to Mae, leading to transcriptional repression of target genes (Fig. 1-4A) (Song et al., 2005). Upon Ras activation and subsequent stimulation of dpMAPK, Yan S127 is phosphorylated (Rebay and Rubin, 1995). This breaks up the polymers, facilitates association with the exportin, Crm1, and effectively displaces Mae from depolymerized Yan (Song et al., 2005; Tootle et al., 2003). Yan is then exported from the nucleus to the cytoplasm where it is proposed to be degraded (Fig. 1-4B) (Rebay and Rubin, 1995; Tootle et al., 2003). This results in derepression of transcriptional targets and allows the cell to respond appropriately to the initial activating signal. Interestingly, among the validated transcriptional targets of Yan is mae itself (Vivekanand et al., 2004). In this context, accumulation of Mae has been suggested to promote a feedback loop resulting in further depolymerization, phosphorylation by dpMAPK, and export of Yan (Fig. 1-4B) (Song et al., 2005). 31 Fig. 1-4. Regulation of the Ets transcriptional regulators Yan and Pnt. (A) In non-stimulated cells, Yan polymers repress target gene transcription. (B) In stimulated cells, dpMAPK translocates into the nucleus promoting PntP2 activation of target gene transcription and export of Yan from the nucleus. Mae mediates Yan depolymerization and PntP2 activity. Adapted from Qiao et al., 2006; Song et al., 2005. 32 In contrast to its role in triggering the dismantlement of Yan-mediated repression, dpMAPK-mediated phosphorylation of PntP2 converts it into a potent activator (Fig. 1-4B). Like Yan, PntP2 bears a SAM domain. Interestingly, the SAM domain of PntP2 contains a docking site for dpMAPK similar to that observed in mammalian Ets-1 and Ets-2 (Seidel and Graves, 2002). Recently, it was shown that mutation of amino acids within this putative docking site leads to a reduction or loss in PntP2 phosphorylation and transcriptional activity, suggesting that dpMAPK docks at the SAM interface to phosphorylate and activate PntP2 (Fig. 1-4B) (Qiao et al., 2006). Like Yan, the PntP2 SAM domain can be bound by Mae (Tootle et al., 2003; Yamada et al., 2003). In vitro kinase assays support the hypothesis that the SAM-mediated MaePntP2 association interferes with PntP2 phosphorylation by dpMAPK, thereby downregulating the transcriptional activation ability of PntP2 (Qiao et al., 2006). As mentioned above, mae is a transcriptional target of PntP2 and Yan (Vivekanand et al., 2004). Thus, activation of mae by activated PntP2 leads to a negative feedback loop whereby Mae protein occludes the phosphorylation and activation of PntP2 (Fig. 1-4B). The Notch Signaling Pathway in Drosophila The Notch signaling pathway, like the DER pathway, is one of several fundamental and evolutionarily conserved mechanisms utilized in metazoans to control cell fate decisions. Notably, cell-cell signaling through the Notch pathway is involved in diverse processes including neurogenesis, proliferation, polarity/axes establishment, and morphogenesis (ArtavanisTsakonas et al., 1999; Baron, 2003; Kopan, 2002; Louvi and Artavanis-Tsakonas, 2006; Mumm and Kopan, 2000; Portin, 2002). The pathway has been characterized as simple and complex: ‘simple’ in that it does not require secondary messengers, but ‘complex’ because of the intricacies of how the relatively few components transduce the signal to the nucleus (Kopan, 2002). The general framework of the pathway is such that signaling is initiated when ligand binds the Notch receptor. Subsequently, Notch is proteolytically cleaved to release its intracellular domain. The Notch intracellular domain then translocates into the nucleus to 33 activate target genes by binding to a DNA-bound transcription factor. Like our presentation of the EGFR pathway, the discussion of the Notch pathway presented here represents a broad overview of information derived from various contexts. Thus, it is important to note that tissueto-tissue differences in deployment of the signaling machinery are not emphasized. Structure of Notch and its Ligands Notch signaling occurs predominantly via cell-cell contacts that allow the extracellular domain of a ligand in a signal-producing cell to interact directly with the extracellular domain of Notch on an adjacent signal-receiving cell (Fehon et al., 1990; Rebay et al., 1991). In addition to trans ligand-receptor interactions, where ligand from the signal-producing cell binds receptor of the signal-receiving cell, it has been observed that ligand and receptor produced within the same cell may autonomously bind. Although not entirely understood, this cis interaction of Notch and ligand is proposed to reduce signaling output of the pathway (Sakamoto et al., 2002; Schweisguth, 2004). In Drosophila, ligands for the Notch (N) receptor are encoded by two single-pass transmembrane proteins: Delta (Dl) and Serrate (Ser) (Fig. 1-5A) (Artavanis-Tsakonas et al., 1999; Fleming et al., 1990; Kopczynski et al., 1988; Vassin et al., 1987). Mammals possess five Notch ligands. These include proteins similar to Dl—Delta-like-1, -3, and -4 (DLL1/3/4)—as well as Ser-like proteins—Jagged-1 and -2 (Jag1/2) (Radtke and Raj, 2003). Dl and Ser encode structurally similar Type I transmembrane proteins with EGF-like homology in their extracellular domains. Dl contains nine continuous EGF-like motifs, a single transmembrane domain and a short cytosolic tail (Fig. 1-5A) (Kopczynski et al., 1988; Vassin et al., 1987). Ser is somewhat larger, with 14 copies of the EGF-like repeat interspersed with short linker sequences in its extracellular domain (Fig. 1-5A) (Fleming et al., 1990). The overall structure of the mammalian DLL1/3/4 proteins is similar to Drosophila Dl, however, DLL3 bears only six EGF-like repeats (Radtke and Raj, 2003). The mammalian Jagged1/2 proteins are similar to Ser, but contain 17 EGF-like repeats with fewer intervening linker sequences (Radtke and Raj, 2003). 34 35 Fig. 1-5. Notch and its ligands. (A) Dl and Ser have similar N-termini with a Notch-ligand N-terminus (Nn) and a Delta-Serrate-LAG-2 (DSL) domain. Ser differs by having a Cys-rich domain that follows the EGF repeats. (B) Notch encodes a large EGF repeat-rich receptor. See text for domain details. While there is a single Notch protein encoded by the Drosophila genome, the mammalian genome encodes four Notch receptors, Notch1-4 (Radtke and Raj, 2003). The Drosophila Notch locus encodes a 2,703 amino acid single-pass transmembrane receptor protein of ~300 kDa (Fig. 1-5B) (Kidd et al., 1986; Wharton et al., 1985a). At its N-terminal extracellular domain, N, has 36 EGF-like tandem repeats, followed by three highly-conserved Cys-rich LIN-12/Notch Repeats (LNRs) (Lieber et al., 1993). Within its cytoplasmic domain, Notch has an RBP-Jκ Association Module (RAM), a single nuclear localization signal (NLS) , seven ankyrin (Ank) repeats, a Transcriptional Activation Domain (TAD) that contains a glutamine-rich Opa repeat, and Pro-Glu-Ser-Thr (PEST) motifs (Fig. 1-5B) (Aster et al., 1997; Bork, 1993; Breeden and Nasmyth, 1987; Jarriault et al., 1995; Kidd et al., 1998; Kurooka et al., 1998; Tamura et al., 1995; Wharton et al., 1985b; Zweifel and Barrick, 2001). The mammalian Notch receptors differ slightly within their extracellular and cytoplasmic domains. In their ectodomains, Notch1-4 proteins contain 36, 36, 34, and 29 EGF repeats, respectively. The cytoplasmic TADs also vary between the receptors, such that Notch1 has a strong but Notch2 has a weak TAD. This is in marked contrast to Notch 3 and 4, which each lack a TAD (Radtke and Raj, 2003). Despite these variations, the high degree of structural and functional conservation of the Notch receptor and its ligands between Drosophila and mammals has permitted integration of a broad range of genetic and biochemical investigations into a unifying model of Notch signaling presented below. Notch is processed to generate heterodimeric receptor Notch proteins are synthesized as precursors that are cleaved during transport to the cell surface, generating NTM transmembrane-bound intracellular and NotchEC extracellular polypeptides (Logeat et al., 1998). Cleavage, at the site termed S1, leaves only 12 amino acids exposed to the cell surface of the NTM fragment (Logeat et al., 1998). Once cleaved, the products heterodimerize via the hydrophobic NotchEC-C-terminal domain to function as a mature receptor (Logeat et al., 1998; Sanchez-Irizarry et al., 2004). Heterodimerization is non-covalent and dependent on Ca2+ levels (Rand et al., 2000). Cleavage of newly-synthesized Notch proteins 36 occurs by the catalytic action of Furin-like Convertases in the trans-Golgi network (Logeat et al., 1998). Proprotein convertases, such as the Furins, function as ‘master switches’ in the activation of growth factors, neuropeptides, receptors, enzymes, glycoproteins, and toxins (Seidah and Prat, 2002). The requirement of Notch processing is different between Drosophila and vertebrates. In mammals, convertase-processing of Notch is constitutive and required for signaling (Logeat et al., 1998). However, in Drosophila, the membrane predominantly has Notch in its full-length, uncleaved form. Cleavage does proceed at low levels but is not required for Notch signaling function (Kidd and Lieber, 2002). Therefore, post-translational regulation of Notch expression is essentially absent in Drosophila. However, the subsequent cleavages, described in the following sections in which we discuss the regulation of and responses to ligand-receptor binding, are fundamental to Notch signaling in both invertebrates and vertebrates. Notch discriminates between ligands through differential glycosylation Following protein synthesis, Notch receptors are modified by the addition of sugar moieties. For example, Notch undergoes N-glycosylation at Asn residues, O-linked glucosylation on Ser residues and O-linked fucosylation on Ser or Thr residues. These reactions are catalyzed by different sets of enzymes. O-glucose and O-fucose may be additionally elongated by a set of glycosyltransferases. The function of O-glucosylation is unclear, but Ofucosylation and its subsequent elongation plays a critical role in ligand-receptor interactions in Notch signaling (Fig. 1-6) (Haines and Irvine, 2003; Haltiwanger, 2002; Haltiwanger and Stanley, 2002). O-glucosylation and O-fucosylation of Notch occur on Ser or Thr residues found in its repetitive extracellular EGF domains (Fig. 1-6) (Moloney et al., 2000; Wang et al., 1996; Wang et al., 2001; Wang and Spellman, 1998). In vivo studies suggest that O-fucosyltransferase (Ofut1) activity is developmentally regulated and required for Notch signaling (Okajima and Irvine, 2002). Specifically knockdown of Ofut1 via RNA interference (RNAi) in cell culture 37 led to a reduction of Notch O-fucosylation, impaired its interaction with ligand, and reduced signaling output from the pathway (Okajima and Irvine, 2002; Okajima et al., 2003). In vivo overexpression of Ofut1 antagonized lateral inhibition and inductive signaling (described later in this review) and led to differential signaling, such that Ser-Notch signaling was increased, but Dl-Notch signaling was inhibited (Okajima et al., 2003). Addition of O-fucosylation sites within the EGF-like repeats may result in ectopic Notch signaling (Li et al., 2003b). Together, these observations suggest that O-fucosylation regulates Notch signaling by influencing ligandreceptor interactions. O-fucosylated residues of Notch are further glycosylated by the β-1,3-N- acetylglucosaminyltransferase enzyme, Fringe (Fig. 1-6) (Haines and Irvine, 2003; Irvine and Wieschaus, 1994). Fng-mediated glycosylation potentiates Dl-Notch binding but inhibits SerNotch binding (Bruckner et al., 2000; Lei et al., 2003). Interestingly, the EGF-like repeats found in the Notch ligands are also O-fucosylated and then Fng-dependently glycosylated (Panin et al., 2002). Glycosylation of the ligands adds another level of complexity, regulation, and specificity in ligand-receptor interactions. In perspective, Fng activity is developmentally regulated and is only required for a subset of Notch signaling events (Haines and Irvine, 2003). It functions in inductive signaling events, but not during lateral inhibition events, each of which are described later in this review (Haines and Irvine, 2003; Irvine and Rauskolb, 2001; Irvine and Vogt, 1997). Notch-Ligand interaction results in extracellular truncation To initiate cell-cell communication, the ligand, Dl or Ser, binds to the extracellular domain of Notch via its own extracellular EGF repeats (Fehon et al., 1990; Rebay et al., 1991). This leads to highly coordinated and regulated proteolytic cleavage of both the ligand and the receptor. As described below, these proteolytic cleavages are essential for Notch signaling (Fig. 1-7) (Selkoe and Kopan, 2003). As mentioned above, the first proteolytic cleavage of Notch at site S1 occurs while the precursor protein transits through the Golgi to the plasma membrane, prior to interacting with 38 Fig. 1-6. Differential sugar modification of Notch affects ligand-binding. O-fucosylation by Ofut1 promotes binding of Ser. Dl interacts with Notch that has been glycosylated by Fng and further modified with galactose and sialic acid. Yellow: Potential EGF repeats for Ofucosylation. Red: EGF Repeat 12, 26, and 27 are highly conserved. Repeat 12 O-fucosylation mediates Dl-Notch signaling activation, whereas O-fucosylation on Repeats 26 and 27 have distinct roles, see following reference for more information: Rampal et al., 2005. Adapted from Haines and Irvine, 2003; Okajima and Irvine, 2002. 39 ligand. Ligand-receptor binding triggers the second cleavage, S2, performed by proteases of the ADAM family (Fig. 1-7). The Drosophila genome encodes five ADAM proteases: TACE/ ADAM17, Kuzbanian (Kuz)/ADAM10, Kuzbanian-like (Kul)/ADAM10, Meltrin/Meltrin α, and Mind-meld (Mmd)/Meltrin α (Sapir et al., 2005). TACE cleaves Notch extracellularly near the transmembrane domain to generate the Notch extracellular truncation (NEXT) product (Brou et al., 2000; Mumm and Kopan, 2000). The NEXT-generating cleavage promotes subsequent proteolytic cleavages by removing inhibitory extracellular sequences surrounding the S2 region, including the three LNR repeats (Fig. 1-5B,7) (Mumm et al., 2000; Sanchez-Irizarry et al., 2004). Additional observations suggest that other ADAM proteases may be involved in Notch signaling events. Kuz was genetically and molecularly implicated in the cleavage processing of the S2 site (Lieber et al., 2002; Pan and Rubin, 1997; Sotillos et al., 1997). In Drosophila, Kuz was shown to be necessary and sufficient for S2 cleavage in vivo and in vitro (Lieber et al., 2002). In contrast, TACE, not Kuz, appears both necessary and sufficient for the S2 cleavage of Notch in mammals (Brou et al., 2000) whereas in C. elegans, ADM-4/TACE and SUP-17/Kuz were reported to function redundantly (Jarriault and Greenwald, 2005). TACE is not required, but is sufficient for S2 cleavage in Drosophila (Lieber et al., 2002). Interestingly, recent studies suggest that ADAM metalloproteases may also have a broader role in processing ligands. For example, Kuz was shown to cleave Dl, as well as Jag/ Ser, independent of ligand-receptor interactions (LaVoie and Selkoe, 2003; Qi et al., 1999; Six et al., 2003). This requirement for ligand proteolytic processing conflicts with reports that suggest that Kuz activity is required exclusively in the signal-receiving cell (Klein, 2002; Sotillos et al., 1997; Wen et al., 1997). The ligands are further processed by the γ-secretase enzyme which is also necessary for release of the intracellular membrane of Notch, as will be described in the following section (LaVoie and Selkoe, 2003; Six et al., 2003). The biological activity of the ligand cleavage products is under debate. One model indicates that the Dl proteolytic event may release its extracellular domain (DlEC) to function as a secreted activating ligand (Qi et al., 1999). In contrast, another model implicates Kuzbanian-mediated proteolysis of ligand 40 as a signal-downregulating event to yield non-functional ligand (Mishra-Gorur et al., 2002). Recently, another ADAM protease, Kul, has been reported to function with Kuz to mediate Dl downregulation to promote unidirectional Notch signaling (Sapir et al., 2005). Regardless of whether the DlEC and Ser/JagEC products serve as active ligands, the intracellular products of γ-secretase activity, DlIC and Ser/JagIC, may themselves play a functional role. One model suggests that ligand cleavage results in localization of the intracellular domains to the nucleus to mediate bi-directional signaling events by facilitating the expression of target genes (Bland et al., 2003; LaVoie and Selkoe, 2003; Six et al., 2003). In addition, these studies demonstrate that such bi-directional signaling events may be antagonistic to one another (LaVoie and Selkoe, 2003). In summary, despite the ongoing controversies, it is clear that extensive proteolytic processing of both ligand and receptor occurs in vivo and that future work will be required to resolve the biological outcome of these events. Endocytosis regulates ligand-receptor interactions in the Notch pathway Notch signaling requires release of the Notch intracellular domain into the cytoplasm. However, amino acid sequences that lie in the ectodomain adjacent to the plasma membrane may create steric or conformational hindrance to intramembrane proteolytic events that release Notch (Fig. 1-7) (Sanchez-Irizarry et al., 2004). In recent years, a model has arisen such that relief of inhibition is mediated by endocytic events that involve both ligand and Notch (Chitnis, 2006; Gupta-Rossi et al., 2004; Le Borgne et al., 2005; Le Borgne and Schweisguth, 2003a; Pitsouli and Delidakis, 2005; Wilkin and Baron, 2005). Upon interaction with Dl or Ser, the ligand-bound cleaved Notch extracellular domain (NotchECD) is transendocytosed into the signalproducing cell (Fig. 1-7) (Parks et al., 2000). The pulling force produced from the endocytic event creates a conformational change and/or relief of steric hindrance of Notch, thus providing access for TACE to cleave at S2 and permitting subsequent Notch cleavage steps by γ-secretase (Parks et al., 2000). Endocytosis of ligand is regulated by the activity of the E3 ubiquitin ligase enzymes 41 Fig. 1-7. Notch proteolytic cleavages. Following binding of ligand, Notch is proteolytically cleaved on its extracellular side at site S2 by ADAM proteases. Neur and Mib promote transendocytosis of Dl and NotchEC. NotchEXT is then cleaved at sites S3 and S4 by the γsecretase complex (Psn/Nct/APH-1/PEN-2), releasing NotchICD into the cytoplasm and secreting the small Notchβ peptide. Adapted from Radtke et al., 2005. Neuralized (Neur) and Mind bomb (Mib) (Fig. 1-7) (Deblandre et al., 2001; Haddon et al., 1998; Itoh et al., 2003; Lai et al., 2001; Lai et al., 2005; Portin and Rantanen, 1991; Wang and Struhl, 2005; Yeh et al., 2001). Although Dl and Ser ligands can be polyubiquitylated and degraded (Deblandre et al., 2001; Lai et al., 2001), in the context of endocytic trafficking, mono- or multiubiquitylation is more likely involved (Haglund et al., 2003; Le Borgne et al., 2005). Such a model remains speculative because the ubiquitylation state (mono- versus multi-) of endocytosed Dl or Ser has not yet been elucidated. Also, it is not clear how the proteasome and the endocytic machinery differentiate between the various states of ubiquitylated proteins (Le Borgne et al., 42 2005). In certain cases, this may be facilitated by the activities of the epsin Liquid Facets (Lqf) and the de-ubiquitylating enzyme Fat facets (Faf) that regulate ligand endocytosis (Chen et al., 2002; Overstreet et al., 2004). In addition to the ligand-receptor pulling mechanism for activating Notch, other, not necessarily mutually exclusive, models for Notch activation have been proposed (Gupta-Rossi et al., 2004; Le Borgne et al., 2005). For example, one model suggests that ligand is produced in an inactive form that must be endocytosed and post-translationally modified prior to being recycled back to the plasma membrane to bind to the Notch receptor (Wang and Struhl, 2004). Alternatively, endocytosed ligand may lead to sorting and lysosomal degradation or to packaging into secreted exosome vesicles that deliver active ligand (Le Borgne and Schweisguth, 2003a; Mishra-Gorur et al., 2002). The details of these rather novel and debated models are beyond the scope of this discussion and are best reviewed in the following references (Le Borgne, 2006; Le Borgne et al., 2005; Le Borgne and Schweisguth, 2003a; Schweisguth, 2004). It seems likely that the current abundance of models that describe the processing and vesicular transport of ligand and receptor stems from the inherent mechanistic diversity and complexity of the Notch pathway signaling, and that many, if not all, of the currently envisioned paradigms reflect biologically relevant means of regulation used in different contexts. Intramembrane cleavage by gamma-secretase releases intracellular Notch Ligand-induced ectodomain shedding to generate NEXT permits subsequent proteolytic events to occur within the transmembrane domain of Notch, liberating intracellular Notch (NotchICD) (Selkoe and Kopan, 2003). Early studies showed the importance of NotchICD in nuclear localization and transcriptional activation (Greenwald, 1994; Kimble et al., 1998). Expression of NotchICD or Notch proteins whose extracellular domain was deleted (Notch∆E) resulted in constitutive signaling output (Lieber et al., 1993; Rebay et al., 1993; Roehl and Kimble, 1993; Struhl et al., 1993). Regulated Intramembrane Proteolysis (RIP) is catalyzed by γ-secretase activity, originally 43 characterized in the processing of the amyloid β-protein precursor (APP) to generate amyloid β-protein (Aβ) and defective in cases of Alzheimer’s disease (Selkoe and Kopan, 2003). γsecretase activity is attributed to the activity of the core protein, Presenilin (Psn), a nine-pass transmembrane protein, that acts in a multi-enzyme protease unit (Fig. 1-7) (De Strooper, 2003; Iwatsubo, 2004; Laudon et al., 2005; Periz and Fortini, 2004; Selkoe and Kopan, 2003). Presenilin physically interacts with Notch and is crucial for the release of NotchICD (De Strooper et al., 1999; Ray et al., 1999a; Ray et al., 1999b; Struhl and Greenwald, 1999; Ye et al., 1999). Loss of Psn in vivo produces a phenotype similar to Notch loss-of-function mutants, emphasizing its critical role in the pathway (Ye et al., 1999). Drosophila Psn is active in a homodimeric conformation, whereas the mammalian paralogs Psn-1 and -2 operate as heterodimers (Cervantes et al., 2004; Selkoe and Kopan, 2003). γ-secretase is composed of multiple factors that constitute the holoenzyme (Fig. 1-7). In addition to Psn, the proteolytic enzyme includes Nicastrin (Nct), a large Type I glycoprotein that binds homodimeric Psn and co-traffics to the plasma membrane via the secretory pathway (Goutte et al., 2000; Yu et al., 2000). Like Psn, Nct is required for liberation of NotchICD and loss of Nct results in reduced γ-secretase activity (Chung and Struhl, 2001; Hu et al., 2002; LopezSchier and St Johnston, 2002). Initially described in C. elegans, further factors that regulate the formation of the γ-secretase holoenzyme include APH-1, a putative seven-pass transmembrane protein, and PEN-2, a small two-pass transmembrane protein (Fig. 1-7) (Francis et al., 2002; Goutte et al., 2002). Although their activities remain unclear, Nct, APH-1, and PEN-2 facilitate the formation and maintenance of the Psn homodimer (Selkoe and Kopan, 2003). γ-secretase cleaves Notch within its transmembrane domain. Although it has not been shown with Drosophila Notch, murine Notch-1 is cleaved by γ-secretase at two sites, S3 and S4 (Fig. 1-7) (Ray et al., 1999a; Ray et al., 1999b). Proteolysis at these sites occurs sequentially such that S4 cleavage is dependent on S3 cleavage (Chandu et al., 2006). The S3 cleavage leads to the release of NotchICD, whereas the S3-S4 cleavage leads to release of a small hydrophobic Nβ peptide (Fig. 1-7) (Moehlmann et al., 2002; Okochi et al., 2002). Recent reports indicate that 44 the Nβ peptide product is secreted, although its function remains to be elucidated (Okochi et al., 2006). Like its ligands, Notch receptor endocytosis may be required for signaling events. Monoubiquitylation and clathrin-dependent endocytosis of Notch is required for γ-secretase activity in signal-receiving cells (Gupta-Rossi et al., 2004). It is unclear whether Notch is cleaved at the plasma membrane or within an intracellular endocytic compartment. One model suggests that following S2-cleavage, Notch is endocytosed, and that subsequent S3/4 cleavage occurs within lipid rafts where active γ-secretase has been observed (Gupta-Rossi et al., 2004; Le Borgne et al., 2005; Pasternak et al., 2004; Vetrivel et al., 2004). However, the data do not ruleout models in which proteolysis occurs at the plasma membrane (Struhl and Adachi, 2000). Recent work suggests that endocytosis may also lead to downregulation of Notch via lysosomal degradation (Le Borgne, 2006). Notch is targeted for endocytosis by the membraneassociated protein Numb and the AP2 clathrin-adaptor complex subunit protein α-adaptin (Berdnik et al., 2002; Guo et al., 1996). Degradation may also be mediated by a set of E3 ubiquitin ligases. For example, Numb may interact with the E3 ubiquitin ligases Itch, Nedd4 or Suppressor of Deltex [Su(Dx)] to promote ubiquitylation, endocytosis, and degradation of Notch (Cornell et al., 1999; Mazaleyrat et al., 2003; McGill and McGlade, 2003; Qiu et al., 2000; Sakata et al., 2004; Shaye and Greenwald, 2005; Wilkin et al., 2004). Adding another layer of complexity, rather than promoting Notch degradation as is most common for E3 ligases, Deltex prevents it from entering the degradative pathway, possibly through an endosomal sorting mechanism (Hori et al., 2004). Determining the validity of these models remains an active area of research, and additional information regarding the endocytic events that govern Notch can be found in the following references (Le Borgne, 2006; Le Borgne et al., 2005; Le Borgne and Schweisguth, 2003a; Schweisguth, 2004). Transcriptional repression in the absence of signaling The primary nuclear effector of Notch signaling in Drosophila is the bifunctional 45 transcriptional regulator Suppressor of Hairless [Su(H)] (Fig. 1-8A) (Lai, 2002b; Schweisguth and Posakony, 1992). Its requirement in Notch signaling is well-documented both genetically and biochemically (Bailey and Posakony, 1995; Fortini and Artavanis-Tsakonas, 1994; Furukawa et al., 1995; Lecourtois and Schweisguth, 1995; Schweisguth and Posakony, 1992). In the absence of Notch pathway activation, Su(H) functions as a transcriptional repressor (Fig. 1-8A), whereas in the presence of active Notch signaling, Su(H) becomes a transcriptional activator (Fig. 1-8B) (Dou et al., 1994; Hsieh et al., 1996; Waltzer et al., 1995). This dual role allows Su(H) to exert opposing effects on the same target genes in different contexts. For example, during specification of the embryonic mesectoderm, Su(H) acts as a downstream effector of Notch signaling to activate expression of the gene singleminded (sim), while simultaneously repressing sim expression in the adjacent neuroectoderm independent of Notch (Klein et al., 2000; Morel and Schweisguth, 2000). In the pathway ‘off’ state, Su(H) exists in a repressor complex. In vertebrates, the Su(H) ortholog, Recombination Signal Sequence-binding Protein Jκ (RBP-Jκ)/C-promoter Binding Factor 1 (CBF-1), mediates transcriptional repression by binding to a variety of co-repressors including Silencing Mediator for Retinoid and Thyroid Receptor (SMRT), Nuclear Receptor CoRepressor (NCoR) (Kao et al., 1998), CBF-interacting repressor (CIR) and KyoT2 (Hsieh et al., 1999; Qin et al., 2004; Taniguchi et al., 1998). Each of these co-repressors regulate repression with RBP-Jκ/CBF-1 in a large complex that includes Sin3A, SAP18, SAP30, RbAp46/48, Sharp/ Mint, and histone deacetylases (HDACs) (Kuroda et al., 2003; Lai, 2002b; Oswald et al., 2002). In Drosophila, Su(H) interacts with the SMRT homolog, Smrter (Smr), and chromatin regulators to facilitate transcriptional repression of target genes (Fig. 1-8A) (Tsai et al., 1999; Tsuda et al., 2002). In a distinct complex, Su(H) also mediates transcriptional repression through the co-repressor adaptor protein Hairless (H) (Fig. 1-8A) (Bang et al., 1995; Schweisguth and Posakony, 1994). With genetic loss of H, Notch signaling targets are ectopically activated (Barolo et al., 2000; Furriols and Bray, 2000; Klein et al., 2000; Morel et al., 2001). Hairless functions by binding to DNA-bound Su(H) along with the global co-repressors C-Terminal 46 Fig. 1-8. Su(H) mediates transcriptional repression and activation. (A) In unstimulated cells, Su(H) represses target genes mediated by a (i) Smr or a (ii) Hairless (H)/Gro repressor complex. (B) Stimulation of the pathway promotes conversion of Su(H) into an activator by NotchICD and Mam, recruiting an activator complex. Adapted from Lai, 2002b. 47 Binding Protein (CtBP) and Groucho (Gro) (Barolo et al., 2002). Recent studies indicate that H-binding to CtBP and Gro, in combination, is required to exert transcriptional repression of Notch signaling targets (Nagel et al., 2005). Like the mammalian CBF-1/RBP-Jκ repression complexes, HDACs are recruited to maintain a repressed state by the Drosophila Su(H)/H/CtBP/ Gro complex (Lai, 2002b). Transactivation of target genes follows proteolytic cleavage Proteolytic activation of Notch leads to activation of target genes of the pathway. Following liberation of NotchICD through proteolytic cleavage, the NotchICD is directed to the nucleus via its NLS motif (Fig. 1-8B) (Hsieh et al., 1996; Lieber et al., 1993). Once in the nucleus, NotchICD interacts directly with Su(H) through its RAM and Ank repeat domains (Kato et al., 1997; Kurooka et al., 1998; Roehl and Kimble, 1993). NotchICD binding to Su(H) displaces the co-repressor complexes and switches Su(H) from a repressor to an activator (Lubman et al., 2004; Mumm and Kopan, 2000). The conversion of Su(H) from a repressor to an activator is mediated, not only by NotchICD, but also by co-activators (Fig. 1-8B). This is facilitated by the bifunctional cofactor Ski-interacting Protein (Skip), which is constitutively bound to Su(H). In the absence of activated Notch, Skip is bound to the SMRT co-repressor complex (Leong et al., 2004; Zhou et al., 2000a; Zhou et al., 2000b). However, Skip has a greater affinity for NotchICD than SMRT, and because binding is mutually exclusive, the presence of NotchICD competes SMRT and associated co-repressors away from Skip and Su(H) (Leong et al., 2004; Zhou et al., 2000b). The NotchICD/Su(H)/Skip complex recruits co-activators such as the histone acetyl transferase (HAT) enzymes CBP/p300, PCAF, and GCN5 which all interact with NotchICD at its Ank repeats and TAD (Fig. 1-8B) (Kurooka and Honjo, 2000; Leong et al., 2004; Wallberg et al., 2002). Transactivation by NotchICD/Su(H) also requires the recruitment of Mastermind (Mam) (Fig. 1-8B) (Fryer et al., 2002; Petcherski and Kimble, 2000a; Petcherski and Kimble, 2000b; Portin and Rantanen, 1991; Wilson and Kovall, 2006; Wu et al., 2000; Wu and Griffin, 2004). 48 mam encodes a Gln-rich nuclear protein that contains an N-terminal Basic domain required to bind NotchICD Ank repeats, an N-terminal Acidic I domain that functions in p300 interactions, and a C-terminal Acidic II domain (Wu and Griffin, 2004). Recruitment of Mam completes the transcriptional complex that activates expression of NotchICD/Su(H) target genes (Wallberg et al., 2002). Recently, the crystal structure of the Notch/Su(H)/Mam ternary complex was solved suggesting that conversion of Su(H) from repressor to activator occurs via a conformational change induced by cooperative interaction of Su(H) with NotchICD and Mam (Barrick and Kopan, 2006; Nam et al., 2006; Wilson and Kovall, 2006). Several genes have been characterized as transcriptional targets of the Notch pathway. These include the Enhancer of Split Complex [E(spl)-C] that consists of thirteen Notchresponsive genes (Jennings et al., 1994; Knust et al., 1987a; Knust et al., 1987b): seven basic Helix-Loop-Helix (bHLH) transcription factors [E(spl)-mβ, -mγ, -mδ, -m3, -m5, -m7, and m8], four Bearded (Brd) family protein [E(spl)-mα, -m2, -m4, and –m6], the co-repressor Gro, and the Kazal-family protease inhibitor E(spl)-m1 (Delidakis and Artavanis-Tsakonas, 1992; Klambt et al., 1989; Knust et al., 1992; Lai et al., 2000; Paroush et al., 1994; Preiss et al., 1988; Wurmbach et al., 1999). The E(spl)-C encodes two families of transcriptional regulators. The E(spl)-bHLH class represses the transcription of proneural genes of the Achaete-Scute complex (AS-C) (Oellers et al., 1994). Little is known about the E(spl)-Brd class, although, Brd family proteins encoded by the Brd complex (Brd-C) locus were recently found to interfere with the interaction of Neur and Dl, thereby inhibiting Dl endocytosis and activation (Bardin and Schweisguth, 2006; Lai et al., 2000). Thus, it is possibile that the Brd family proteins within the E(spl)-C, like the Brd-C proteins, may exert negative feedback on the Notch signaling pathway by antagonizing ligand activation. In summary, we have described how the binding of ligand leads to the proteolytic release of activated NotchICD, conversion of Su(H) from a repressor to an activator, and transcription 49 of target genes such as the E(spl)-C. In the following section, the different general contexts for which Notch signaling is used will be presented. Context-specific variations in Notch signaling Numerous developmental processes require Notch signaling (Bray, 1998; Lai, 2004). However, the Notch signaling machinery is not utilized in exactly the same way in all contexts; instead, variations in the basic events described in the previous section are used to elicit an array of context-specific cell-to-cell signals. Three major contexts requiring Notch signaling will be discussed below: lateral inhibition, binary cell fate decisions, and boundary formation (Fig. 1-9). Lateral inhibition occurs when one cell is selected from a group of equivalent precursors and subsequently signals to its neighbors to prevent them from adopting that same fate (Fig. 19A). In the context of Notch signaling during Drosophila embryonic neurogenesis, the process starts with a field of homogeneous and equipotent cells, called the neurogenic region. These cells initially express ligand and receptor equally and have the ability to signal to one another (Artavanis-Tsakonas et al., 1999; Bray, 1998; Doe and Goodman, 1985; Lai, 2004; Moscoso del Prado and Garcia-Bellido, 1984; Wigglesworth, 1940). Expression of transcription factors of the bHLH family that include Achaete (Ac), Scute (Sc), Lethal of Scute (L’sc), and Asense (Ase) (Achaete-Scute Complex/AS-C), referred to collectively as the proneural genes and initially expressed throughout the neurogenic region, becomes refined to smaller proneural clusters from which a single neuroblast will be selected (Fig. 1-9A) (Skeath and Carroll, 1994). Dl and Notch expression is initially uniform within the proneural cluster, but an elevation of Dl in a single cell, the presumptive neuroblast, leads to the activation of Notch in the surrounding cells. Notch signaling activates the E(spl)-C genes whose encoded products repress expression of the proneural AS-C genes, resulting in downregulation of Dl, a downstream transcriptional target, and leading to molecular inhibition in Notch-activated cells (Fig. 1-9A). In turn, increased Dl expression in the presumptive neuroblast cell results in decreased Notch activation in the presumptive neuroblast, leading to reduced E(spl)-C expression and de-repression of the AS-C 50 Fig. 1-9. Three modes of Notch signaling function. (A) In lateral inhibition, (i) initially, cells in a field have equal potential, but stochastically or through a slight biased, one cell adopts a specific fate, e.g., a neuroblast (black cell), while inhibiting surrounding cells from taking on that same fate. (ii) This change is due to differential AS-C expression to activate Dl in the neuroblast versus upregulation of E(spl)-C in the surrounding cells to repress Dl expression. (B) During the determination of binary cell fates from an SOP, (i) a pair of cells have a bias towards one cell fate due to differential segregation of Neur and Numb components into pIIB. (ii) Neuralized promotes Dl-Notch transendocytosis in pIIB, while Numb inhibits activation of Notch. (C) During wing compartment formation, (i) Asymmetric ligand expression, dorsal Ser and ventral Dl along the dorsal-ventral midline, along with asymmetric dorsal Fng expression (light shading) results in high activated Notch expression (dark shading) and formation of a dorsal-ventral boundary. (ii) Due to asymmetric Fng-mediated Notch O-glycosylation, cells at the boundary have high Notch activity because of both Dl-Notch and Ser-Notch target activation. Adapted from Beatus and Lendahl, 1998; Haines and Irvine, 2003; Lai, 2004. 51 (Fig. 1-9A) (Singson et al., 1994). Recently, it was observed that Su(H)’s repressive function assures the downregulation of E(spl)-C in this process (Castro et al., 2005). Thus, at the conclusion of this process, lateral inhibition focuses ligand expression within a single cell that will adopt a neural fate, while the surrounding cells with active Notch signaling will become non-neuronal, epidermal cells (Fig. 1-9A) (Beatus and Lendahl, 1998). This inhibitory process is proposed to occur at a local level such that cell fates are determined through direct contact and short-range signaling between cells expressing proneural and neurogenic genes (ArtavanisTsakonas et al., 1999). The specification of a single precursor from within the proneural cluster was initially proposed to be a stochastic decision (Beatus and Lendahl, 1998; Simpson, 1997). However, it was subsequently observed that the SOP always occupies the same position within the cluster, suggesting that the expression of the proneural genes is somehow ‘pre-patterned’ to create a slight bias for selection of the sensory organ precursor (SOP) (Cubas et al., 1991; Culi and Modolell, 1998). Although the notion of pre-patterning provides a satisfying explanation for the regularly spaced arrangement of SOPs, the mechanisms that bias the SOP for selection remain unclear. Once a bias is determined, maintainance of the selection becomes important. For example, the simple model whereby cells that surround the single SOP are inhibited from adopting a neural fate goes hand-in-hand with a model whereby the SOP is affirmed, or ‘wins’, its cell fate. It is observed that once pre-patterned, the proneural AS-C genes play an active role in selecting a single precursor. That is, the proneural genes expressed in the precursor provide positive feedback to regulate themselves in cis and to upregulate other proneural target genes in trans to overcome the repression of those genes repressed by the E(spl)-C (Culi and Modolell, 1998; Gibert and Simpson, 2003; Nakao and Campos-Ortega, 1996; Skeath and Carroll, 1994). The choice of binary cell fate decisions during asymmetric cell divisions may also be regulated by the Notch signaling pathway (Fig. 1-9B). This special directional inhibitory signaling mechanism differs from lateral inhibition. In these specialized cell divisions, a sensory organ precursor (SOP) cell divides to produce two daughter cells. These daughter cells are each 52 capable of sending and receiving Notch signaling, however, rather than being equipotent, these cells have bias and directionality in their signaling interactions as influenced by determinants that were asymmetrically allocated between the daughter cells during cell division. This leads to adoption of differential sibling fates such that one cell expresses signal whereas the other receives signal from the Notch receptor (Fig. 1-9B) (Bray, 1998; Lai, 2004). This type of Notch inhibitory signaling occurs in the course of binary cell fate decisions following the SOP choice during bristle cell development (Roegiers and Jan, 2004). The SOP undergoes an asymmetric cell division such that Numb and Neur are segregated to the pIIb cell but not the pIIa cell (Fig. 1-9B) (Guo et al., 1996; Rhyu et al., 1994; Spana and Doe, 1996; Uemura et al., 1989). Neur promotes Dl activation while Numb antagonizes Notch activation in pIIb (Lai and Rubin, 2001; Le Borgne and Schweisguth, 2003b). This leads to an asymmetry of ligand and receptor: pIIb expresses ligand and pIIa expresses a competent receptor (Fig. 19B). The asymmetry results in activation of the Notch pathway in pIIa, activation of E(spl)-C, and inhibition of proneural genes (Gigliani et al., 1996; Heitzler et al., 1996). The pIIb divides asymmetrically to give forth sheath and neuronal cells (along with a glial precursor cell in the gliogenic SOP lineage); the pIIa divides asymmetrically to yield hair shaft cell and socket cell progeny (Fig. 1-9B) (Jan and Jan, 1995; Van De Bor and Giangrande, 2001). These cells together form a sensory unit in the adult fly. The bristle sensory unit is, therefore, a product of lateral inhibition generated by the SOP that in turn yields two different lineages of cells as a result of binary cell fate decisions mediated by the Notch pathway. In contrast to inhibitory Notch signaling, inductive Notch signaling promotes rather than represses a given cell fate in the signal-receiving cells by activating transcriptional targets (Bray, 1998; Lai, 2004). Like the binary cell fate decision, inductive Notch signaling proceeds amongst cells that have a biased, rather than an equivalent, potential (Lai, 2004). The inductive process is well-characterized in the development of the dorsal-ventral boundary in the Drosophila wing (Fig. 1-9C) (Rauskolb et al., 1999). In the developing wing tissue, Fng and Ser are expressed in the dorsal compartment whereas Dl is expressed in the 53 ventral part (Fig. 1-9C) (Doherty et al., 1996; Rauskolb et al., 1999). Since Fng inhibits Ser’s ability to interact with Notch, dorsal cells may only signal to dorsoventral boundary cells. These boundary cells receive signal from ventrally-expressed Dl as well as dorsally-expressed Ser. Thus, the dorsoventral boundary cells have a sharp activation of Notch (Fig. 1-9C). This leads to induction of the cell fate for the wing boundary through the induction, rather than the inhibition, of the genes wingless (wg), vestigial (vg), and others (Fleming et al., 1997; Lai, 2004; Panin et al., 1997; Rauskolb et al., 1999; Wu and Rao, 1999). In conclusion, the three general mechanisms for Notch signaling—lateral inhibition, binary cell fate decision and boundary formation—outlined above will be an important reference for our discussion of DER and Notch signaling pathway interactions below. Interplay of the DER and Notch signaling pathways Relationships between DER and Notch signaling during development Three main themes will run through our discussion of the mechanisms by which DER and Notch relate: consequence, time, and space. First, the outcome, or consequence, of DER and Notch signaling is highly context specific. Mutual antagonism is perhaps the most common relationship in EGFR-Notch interactions, although in some situations the pathways cooperate to potentiate each other’s signaling activities. Second, understanding the temporal integration of DER and Notch signaling, that is whether the two pathways operate sequentially or simultaneously and how that affects overall output, is a critical element that we are only beginning to have the tools to address at single cell resolution in an in vivo system. Third, the spatial coordination of DER and Notch signal transduction across a field of developing cells provides another key determinant of output specificity. For example, activation of DER and Notch may occur within the same cell, or in adjacent cells, and depending on context, these differences lead to profound distinctions in how the information is processed. In the remainder 54 of this review we will compare and contrast the various signaling interactions between DER and Notch during the temporal course of Drosophila eye development. The Drosophila eye Drosophila provides a superb genetically tractable system to study developmental signal transduction using genetic, biochemical, cellular, genomic, and molecular approaches. In particular, the evolutionarily conserved signaling circuitries that produce the precise stereotyped patterning of the Drosophila compound eye have proven an excellent backdrop for studying cellcell communication, especially with respect to signaling integration and cross-talk (Voas and Rebay, 2004). Importantly, many of the discoveries emanating from investigations of signaling mechanisms in the fly eye have proven to be broadly conserved. The Drosophila compound eye is composed of about 750 repetitive subunits, termed ommatidia, which are arranged symmetrically with respect to the horizontal equator line to form a precise neurocrystalline lattice (Fig. 1-10B). The cellular architecture of an ommatidium enables it to focus light from its individual visual field and to transmit the information to the brain via the axonal projections of the photoreceptors (Ready et al., 1976; Tomlinson and Ready, 1987). Each mature ommatidium contains eight neuronal photoreceptors (R1-8) arranged in a stereotypical trapezoidal pattern (Fig. 1-10C). The non-neuronal cells consist of four lenssecreting cone cells, two primary (1º) pigment cells, and a hexagonal lattice of secondary/tertiary (2º/3º) pigment cells with interspersed sensory bristle cells (Ready et al., 1976; Tomlinson and Ready, 1987). Cells within the lattice are specified in a regulated and reproducible manner (Ready et al., 1976; Tomlinson and Ready, 1987). Deviations from the precise development of the eye by loss or gain of cell types, e.g., through genetic mutation or transgene expression, lead to obvious defects. Characterization of the molecular basis behind these defects has resulted in elucidation of the genes and pathways required for proper development of the eye. In the ensuing discussion, we have selected several cell fate specification and morphogenetic 55 56 Fig. 1-10. The Drosophila eye. (A) Scanning electron micrograph of the adult compound eye. (B) Tangential section and cartoon of the adult eye showing the hexagonal ommatidia and regular pigment cell lattice. The dark-stained light-sensing photoreceptor organelles, or rhabdomeres, are arranged in a trapezoidal pattern. R1-7 are shown in this section; R8 is located at lower section levels. (B, C) Ommatidia are arranged symmetrically with respect to the equator, such that R3 points up in the dorsal half, whereas R4 points down in the ventral half. events as the backdrop for considering the various modes of DER-Notch pathway crosstalk and integration. Our goal is not to provide a comprehensive view of eye development per se, but rather to discuss the conserved signaling paradigms that have emerged from these studies. DER and Notch regulate the Retinal Determination Gene Network The eye and antenna arise from a columnar epithelial primordium, known as the eye- antennal imaginal disc (Fig. 1-11A). Set aside as small clusters of cells in the early embryo, imaginal disc epithelia undergo extensive proliferation and patterning during larval and pupal stages, and eventually give rise to the adult structures during metamorphosis (Cohen, 1993). During early larval stages, the eye-antennal disc grows through unpatterned, asynchronous cell division (Wolff and Ready, 1993). Within this epithelium, specification of the eye field occurs through the actions of the retinal determination gene network (RDGN) (Fig. 1-11B) (Pappu and Mardon, 2004; Silver and Rebay, 2005), a group of transcription factors that function as master regulators or eye selector factors. Proteins within the RDGN are expressed in a hierarchial manner whereby the Pax6 transcription factor Twin of eyeless (Toy) activates another Pax6 member Eyeless (Ey). Ey potentiates expression of Sine oculis (So), Eyes absent (Eya), and, subsequently, Dachshund (Dac). As the name suggests, these regulators function as a network: the hierarchy does not solely run in a linear fashion, but, rather, factors downstream initiate feedback loops to activate other members within the group. In addition, other factors such as Optix and the Pax6 member Eyegone (Eyg) play independent but necessary roles in specifying the eye field (Fig. 1-11A) (Dominguez and Casares, 2005; Pappu and Mardon, 2004; Silver and Rebay, 2005). The DER and Notch pathways both play critical roles during early eye development. Early reports postulated that these two pathways antagonize each other in the specification of the eye field within the eye-antennal disc anlagen in the latter half of the second larval instar (Kumar and Moses, 2001a; Kurata et al., 2000). However, a more recent report supports a model whereby Notch and DER are not involved in specification, but, rather, specification occurs earlier 57 during the first half of the second instar larval stage, independent of Notch and DER (Kenyon et al., 2003). At this time, the eye selector genes Ey and Toy retract from their ubiquitous expression in the eye-antennal disc to a region that includes the posterior two-thirds of the eyeantennal disc, where they inhibit the expression of the antennal marker distalless (dll) (Kenyon et al., 2003). The signaling events that initiate the re-distribution of Ey and Toy expression remain to be elucidated. The Notch signaling pathway promotes proliferation within the eye disc over a period of several days that spans from late first to the early third larval instar (Dominguez and Casares, 2005; Dominguez et al., 2004; Kenyon et al., 2003). Although expressed throughout the eye disc, Notch activation is restricted to the dorsoventral equator line through Notch’s inductive boundary formation activities that are regulated by Fng in the ventral half of the eye disc and the Fng-inhibiting Iroquois protein complex in the dorsal compartment (Fig. 1-11B) (Cavodeassi et al., 1999; Cho and Choi, 1998; Dominguez and Casares, 2005; Dominguez and de Celis, 1998; Papayannopoulos et al., 1998; Yang et al., 1999). From here, the Notch pathway induces proliferation of the eye disc by establishing an organizing center required for eye growth (Cavodeassi et al., 2001; Dominguez and de Celis, 1998; Kenyon et al., 2003; ReynoldsKenneally and Mlodzik, 2005). This organizing center requires Notch as well as Eyg to promote growth of the eye disc. The Jak/STAT pathway ligand Unpaired (Upd) is a downstream component of the organizing center that acts over long distance to promote cell proliferation over the entire eye field (Fig. 1-11B) (Chao et al., 2004; Reynolds-Kenneally and Mlodzik, 2005). It remains to be elucidated whether upd is a direct transcriptional target of the Notch signaling pathway during the proliferative process. In addition, Notch has been reported to mediate proliferation by promoting the G1/S transition by activating genes such as dE2F1, which encodes the Drosophila homolog of the vertebrate cell-cycle regulatory transcription factor E2F (Baonza and Freeman, 2005). Whether this mechanism contributes to early proliferation in the eye disc remains an open question. 58 The Notch pathway also plays a positive role with respect to early eye development Fig. 1-11. Regulation of early eye development by Notch and DER. (A) Eye growth and specification are controlled by the Retinal Determination Gene Network. Dark solids lines indicate transcriptional regulation. Notch indirectly activates Eya and may mutually antagonize DER (dotted lines). DER activates Eya through phosphorylation via MAPK (orange line). (B) The Iroquois Complex (Iro-C) inhibits Fng in the dorsal half of the early eye imaginal disc. Differential ligand expression of Dl (dorsal) and Ser (ventral) along with asymmetric Fng expression (ventral) mediates dorsal-ventral boundary formation where activated Notch (NotchICD) is expressed, resulting in disc growth via Eyg and Upd. Adapted from Cavodeassi et al., 2001. 59 through the regulation of the RDGN gene eyg. Because eyg is dispensible for the specification of the eye field, in this context, Notch again positively regulates eye growth, not specification (Fig. 1-11B) (Dominguez et al., 2004). It is not yet understood if the regulation of eyg occurs through direct transcriptional activation by Notch/Su(H), but analysis in conditional Notch mutants along with mosaic mutant analysis of Su(H) and Ser loss-of-function alleles implicate this to be the case. In support of this hypothesis, Eyg protein was found to be coincidently expressed with the E(spl)-C proteins (Dominguez et al., 2004). The observation that eye specification occurs earlier than had been previously reported and the insight that Notch regulates eye disc proliferation does not address the observation that the Notch and DER pathways antagonize each other during the second larval instar (Kenyon et al., 2003; Kumar and Moses, 2001a; Kurata et al., 2000). Does DER signaling still play a role at this stage? These observations implicate DER signaling in antagonizing the proliferation mediated by Notch. This would be surprising because the DER pathway is considered to be a key positive regulator of eye development with respect to proliferation and differentiation (Freeman, 1996). Furthermore, the DER pathway positively regulates the activity of the RDGN through MAPK phosphorylation of Eya (Fig. 1-11A) (Hsiao et al., 2001). Reconciliation of these observations can be achieved by proposing that the DER pathway functions in temporally distinct manners during the development of the eye: it may antagonize early Notch-mediated proliferation, but later promote proliferation and other aspects of eye development. If such a model is true, then it will be important to determine how regulation of DER specificity is achieved. DER and Notch at the morphogenetic furrow Following eye field specification, determination and differentiation occurs in a dramatic regulated fashion across the eye imaginal disc. During the late larval and early pupal stages, an indentation in the eye disc, known as the morphogenetic furrow (MF), is initiated from the posterior and progresses anteriorly in a wave that leaves in its wake differentiating cells (Fig. 60 1-12) (Wolff and Ready, 1991). Overtly undifferentiated cells randomly proliferate anterior to the MF, whereas posterior to the MF, ommatidial clusters start to form as cells that make up each ommatidium are recruited (Wolff and Ready, 1993). Five domains of gene expression have been defined in the developing eye disc to categorize the complex series of events that underly ommatidial development (Fig. 1-12). The MF lies within Zone III. Ahead of the MF are two domains of transcription factor expression, Zones I and II, where the early RDGN proteins—Toy, Ey, Eyg, Optix are expressed. Other RDGN proteins, such as Eya and So, are expressed in Zones II-V (Bessa et al., 2002; Pappu and Mardon, 2004). The MF is initiated by the coordination of several signaling pathways (Fig. 1-12A) (Treisman and Heberlein, 1998). Beginning at the posterior of the eye disc, the Hedgehog (Hh) signal is expressed in differentiating photoreceptor neurons. Hh is necessary and sufficient for MF formation (Heberlein and Moses, 1995; Ma et al., 1993). Hh induces both long-range signaling by the secreted protein Decapentaplegic (Dpp), the Drosophila Transforming Growth Factor-β (TGF-β) ortholog, and the pre-patterned expression of the basic helix-loop-helix (bHLH) transcription factor Atonal (Ato) (Baker et al., 1996; Greenwood and Struhl, 1999; Jarman et al., 1994; Jarman et al., 1995). Since Hh drives the progressive movement of the MF wave, cells that receive Hh will begin differentiation, and then express and send Hh to anterior cells (Fig. 1-12A) (Heberlein and Moses, 1995; Ma et al., 1993; Rogers et al., 2005). Dpp signals to cells anterior to the MF, in Zone II, to upregulate both Ato and the repressors Hairy (h) and Extramacrochaetae (Emc), which are needed to establish a pre-proneural (PPN) state (Fig. 1-12) (Brown et al., 1995; Fu and Baker, 2003; Greenwood and Struhl, 1999). Hairy regulates bHLH proteins through transcriptional repression; Emc represses bHLH proteins through protein-protein interaction (Ellis et al., 1990; Garrell and Modolell, 1990; Van Doren et al., 1992). Hairy and Emc provide negative feedback with respect to the Hh pathway to repress Ato transcription (Brown et al., 1995). Additionally, the Wg signaling pathway provides negative regulation of the MF, preventing MF initiation at the lateral margins of the disc (Fig. 112A) (Ma et al., 1993; Treisman et al., 1997). 61 The DER and Notch signaling pathways are both required for MF initiation upstream of the Hh and Dpp pathways, but downstream of Wg (Kumar and Moses, 2001b). In contrast to their antagonistic role in early eye development, described above, DER and Notch function cooperatively to promote the formation of the MF. Through the use of conditional mutants and tissue- and temporal-specific overexpression of activated DER pathway or dominant-negative Notch pathway components, the requirement of both of these pathways was identified. DER is required at two times for proper MF establishment: (1) before initiation of the MF (birth) and (2) for propagation of new ommatidial columns along the disc’s lateral margins (reincarnation) (Kumar and Moses, 2001b). Notch is required for the reincarnation step of MF generation, where dpp function is required to promote differentiation along the margins to permit the MF to move anteriorly (Chanut and Heberlein, 1997; Treisman and Heberlein, 1998; Wiersdorff et al., 1996). Epistasis experiments suggested that DER acts upstream of Notch (Kumar and Moses, 2001b) and upstream of Hh signaling during MF birth. DER and Notch also function upstream of Dpp during MF reincarnation (Kumar and Moses, 2001b). It remains to be elucidated how the DER and Notch pathways relate to each other at a molecular level. Do the two pathways interact directly or do they simply converge to promote MF initiation? Maintenance of the wave-like progression of the MF, where cells at or behind the MF signal Hh to cells ahead of the furrow in a cyclical manner, requires a mechanism that propagates the Hh signal forward. The observation that Hh protein expression is spatially coincident with where the DER pathway is activated suggests that DER signaling might provide this function (Fig. 1-12A) (Kumar et al., 2003; Rogers et al., 2005). The transcriptional activator Pointed (Pnt), a key effector of the DER pathway, binds to a set of four Ets-binding sites located in an enhancer element that drives hh expression in the developing photoreceptors behind the MF (Rogers et al., 2005). In vivo experiments show that loss of Pnt leads to downregulation of hh-enhancer reporter expression, strongly suggesting that Hh is activated by the DER pathway (Rogers et al., 2005). Thus, DER-activated expression of Hh induces progressive differentiation of photoreceptors, in turn leading to additional DER signaling and providing a positive feedback 62 loop for continued forward propagation of the MF (Rogers et al., 2005). The Notch pathway acts downstream of Hh signaling to promote proneural activity via inductive regulation of the bHLH transcription factor Atonal (Ato) (Fig. 1-12A) (Baker, 2000; Baker and Yu, 1997; Frankfort and Mardon, 2002; Hsiung and Moses, 2002; Jarman et al., 1994; Jarman et al., 1995). As aforementioned, Hairy and Emc negatively regulate ato expression in the PPN state (Brown et al., 1995) but are themselves negatively regulated by Notch signaling (Fig. 1-12) (Baonza et al., 2001). During the PPN period, ato is also directly repressed by the Notch pathway effector Su(H) (Li and Baker, 2001). Dl-mediated activation of the Notch pathway converts Su(H) to an activator with respect to ato transcription and also potentiates expression of the E(spl)-C, which downregulates the Hairy and Emc mediated repression of ato. The ensuing upregulation and accumulation of Ato immediately ahead of the MF, a process termed ‘proneural enhancement’, promotes subsequent neural differentiation (Fig. 1-12A,13A) (Baonza et al., 2001; Li and Baker, 2001) (Baker, 2000; Baker and Yu, 1997; Frankfort and Mardon, 2002). Because of its positive function in regulating proneural activity, Notch signaling is considered to function in an inductive signaling context during these events (Baonza and Freeman, 2001; Nagel and Preiss, 1999). Although both DER and Notch act positively in the MF, the two pathways may not directly cooperate, but rather may function in a precise temporal sequence to promote the MF. That is, Notch signaling promotes proneural enhancement ahead of the MF leading to neural differentiation and DER signaling in presumptive photoreceptors, which upregulates the Hh signal to initiate a new round of proneural development (Baonza et al., 2001; Rogers et al., 2005). Therefore, Notch and DER activity in the MF, like Hh, is cyclical, allowing neural differentiation to sweep across the eye field. Notch and DER in R8 founder cell determination The proneural protein Ato is essential for ommatidial formation (Jarman et al., 1995). Ato forms heteromultimers with another bHLH protein Daughterless (Da) (Jarman et al., 1993). 63 Fig. 1-12. Morphogenetic Furrow (MF) progression. (A) DER from differentiating cells activates the Hh pathway to initiate and re-initiate the MF. Hh, Dpp, and Notch synergize to activate Ato expression in region III. Expression domains of regulators from of the eye disc are indicated in (B). Adapted from Silver and Rebay, 2005. 64 These Ato-Da complexes are proposed to act as sequence-specific DNA-binding complexes to mediate the transcription of target genes. Ato is initially expressed broadly at the MF, but posteriorly, becomes restricted to a group of 20 cells known as intermediate groups or rosettes (Baker and Yu, 1997; Brown et al., 1996; Jarman et al., 1995). Prior to intermediate group restriction, ato expression is under the transcriptional activation of Hh and Notch (Fig. 1-12A,13) (Baonza et al., 2001). Through differential regulation of transcriptional enhancer sites, ato switches to a mode of autoregulation to promote restriction of cells into intermediate groups (Fig. 1-13A,C) (Baker et al., 1996; Dokucu et al., 1996; Sun et al., 1998). This signifies a critical switch whereby Ato is no longer activated by Notch, but instead represses Notch activity in the context of lateral inhibition. Thus, the Notch pathway transitions from a mode of inductive proneural enhancement to that of lateral inhibition (Fig. 1-13) (Baker and Yu, 1997). Ato expression, which correlates with competence to become founder cells, is refined in two steps: first, Ato is restricted to two to three cells termed the R8 equivalence group, and second, Ato expression is finally focused to a single R8 photoreceptor founder cell (Fig. 1-13A) (Baker and Yu, 1997; Dokucu et al., 1996; Jarman et al., 1994; Jarman et al., 1995). The process whereby Ato is restricted to a single founder cell within the intermediate group requires lateral inhibition through Notch signaling. Just as we described earlier for embryonic neurogenesis, strong Dl activity becomes restricted first to the equivalence group and then to the presumptive R8 cell and activates the Notch/Su(H)/E(spl)-C cascade which represses ato in surrounding cells (Fig. 1-13C) (Lee et al., 1996; Li and Baker, 2001). Successful lateral inhibition is important for producing the regular, evenly-spaced, alternating groups of cells that comprise the ommatidial units (Artavanis-Tsakonas et al., 1999; Baker, 2000; Frankfort and Mardon, 2002; Hsiung and Moses, 2002). Lateral inhibition is also facilitated by the fibrinogen-related dimeric secreted glycoprotein Scabrous (Sca) (Fig. 1-13C) (Baker et al., 1990; Baker and Zitron, 1995; Lee et al., 1996; Mlodzik et al., 1990a). In contrast to localized lateral inhibition, inhibitory events mediated by Sca occur over greater distances. Importantly, as R8 cells are spaced further apart 65 than cells seen in SOP formation, Sca function may provide a mechanism by which cell spacing over greater distances is achieved. Sca is activated downstream of Ato in the equivalence groups and then in the presumptive R8 from where it is secreted to the surrounding cells over multiple cell diameters. There, Sca interacts with Notch to single-out the R8 cell by facilitating the activation of the Notch pathway several cell-diameters away from the Sca-secreting cell (Baker and Zitron, 1995; Lee et al., 1996). This effectively broadens the inhibitory strength of the Ato-producing R8 cell to inhibit ato expression in its neighbors. Secreted Sca is received by surrounding cells and endocytosed along with inactive Notch receptors within these cells (Chou and Chien, 2002; Li et al., 2003a). Within the endosomes Sca somehow activates Notch through direct interaction, perhaps by protecting Notch from endocytic degradation (Li et al., 2003a; Powell et al., 2001). It would be interesting to determine whether endocytic targeting is affected and/or whether interactions with the E3 ubiquitin ligase Deltex are relevant to Notch endocytic sorting and activation in this context. Initially, DER was proposed to play a role in the specification of R8, because the expression of the hypermorphic DER allele Ellipse (Elp) leads to changes in ommatidial precluster formation and loss of R8 differentiation (Baker and Rubin, 1989). In order to determine DER’s requirement in R8 specification, mitotic recombination was induced to generate and compare DER null mutant clones with wildtype control clones (Dominguez et al., 1998; Xu and Rubin, 1993). Since DERnull homozygosity only yielded very small clones, likely due to a requirement for DER in cell proliferation and/or survival, the clones were induced in a Minute mutant background (Baonza et al., 2001; Dominguez et al., 1998; Lesokhin et al., 1999; Yang and Baker, 2001). In the Minute system, all cells, except those that are homozygous for DERnull, are heterozygous for a dominant growth-impeding mutation (Ferrus, 1975; Morata and Ripoll, 1975). This provides cells in the DERnull clone a competitive growth advantage. In this Minute background, R8s still develop in DERnull clones, suggesting that DER does not have a role in R8 specification (Baonza et al., 2001; Dominguez et al., 1998; Lesokhin et al., 1999; Yang and Baker, 2001). This conclusion was further confirmed when R8 specification was demonstrated in 66 Fig. 1-13. Founder cell specification. (A) Ato expression (blue) is upregulated in cells anterior to the MF via proneural enhancement (B). Lateral inhibition (C) refines Ato expression to a single Ato expressing cell that will be specified to be the R8 founder cell (A). Adapted from Frankfort and Mardon, 2002. eye discs from flies carrying the temperature-sensitive DERtsla allele and reared at the restrictive temperature. Loss-of-function clones of Ras similarly underscored the dispensability of DER pathway signaling for R8 specification (Spencer et al., 1998; Yang and Baker, 2001). As we have mentioned above, ommatidial spacing and the focusing of Ato to a single R8 is essential for the regular cell complement and pattern in the adult eye. While the Notch 67 pathway and Sca function are clearly needed, there is much debate concerning the requirement for DER in R8 spacing. Arguing for a role, DERnull and Ras mutant clones generated in a Minute background have aberrant and perturbed R8 spacing as immunohistochemically measured with Ato expression and R8-specific markers (Baonza et al., 2001; Dominguez et al., 1998; Lesokhin et al., 1999; Yang and Baker, 2001). However a subsequent study raised into question the interpretation of these experiments and argued against a role for DER in R8 spacing, based on the finding that Minute+ clones could produce a non-cell autonomous effect on Ato expression and R8 spacing, even in control clones (Rodrigues et al., 2005). In contrast, the investigators who conducted the DERnull studies (Baonza et al., 2001) have not observed such non-autonomous Minute-induced spacing defects in mutant clones of numerous other genes, but have observed spacing defects in non-Minute clones of various components of the DER pathway (A. Baonza and M. Freeman, pers. comm.). This suggests that the surprising finding by Rodrigues et al. (2005) likely reflects a problem with the particular genetic background rather than a general problem with the commonly-utilized Minute technique (M. Freeman, pers. comm.). Slightly more difficult to interpret, and further fomenting the debate, is the finding that at the restrictive temperature, DERtsla mutant clones initially maintain proper R8 spacing, but later develop R8 spacing defects, a result that may indicate a role in maintenance rather than initiation of R8 spacing (Kumar et al., 2003; Kumar et al., 1998; Rodrigues et al., 2005). Relevant to the interpretation of this experiment is the nature of the DERtsla allele, in particular whether or not it represents a true genetic null. The DERtsla allele carries an apparently critical substitution in the ligand-binding domain and has been reported to behave as a null at the restrictive temperature based on reduced protein stability, loss of dpMAPK, loss of phosphotyrosine signal, and loss of cell surface localization in Drosophila cultured cells (Kumar et al., 1998; Rodrigues et al., 2005). However, it remains formally possible that in vivo, DERtsla retains sufficient function, perhaps via binding to the alternate ligand Keren, to promote R8 spacing. Although definitive resolution of the debate will require further experimentation, it seems clear that if DER signaling contributes 68 to initial R8 spacing, then significantly lower levels of pathway activation are required in this context than in many others during eye development. An alternate model regarding DER’s role in R8 spacing is that fine-tuning through reduction or repression in level of DER signaling is important for proper R8 spacing. The prediction in this case is that even if DER signaling is not required for R8 spacing, overexpression of activated DER components should disrupt R8 spacing, as has in fact been seen with ectopic RasV12 expression (Spencer et al., 1998). Insight into the possible mechanism of DER repression derives from mutants for echinoid (ed), a gene that encodes a cell adhesion molecule. Ed is a component of adherens junctions and cooperates with DE-cadherin to mediate cell adhesion (Wei et al., 2005). Loss of ed leads to twinning of R8s at the MF. DER signaling activities, including dpMAPK levels, are upregulated in ed mutants suggesting that Ed functions to repress DER signaling (Fig. 1-13C) (Bai et al., 2001; Islam et al., 2003; Rawlins et al., 2003b; Spencer and Cagan, 2003). Interestingly, Ed cooperates with the Notch signaling pathway. ed mutants have similar phenotypes to that of Notch signaling mutants and synergize with mutants of this pathway. Together, Ed and Notch were reported to regulate the efficient development of SOPs and to antagonize DER signaling during embryonic and wing neurogenesis (Ahmed et al., 2003; Chandra et al., 2003; Escudero et al., 2003; Islam et al., 2003; Rawlins et al., 2003a). One model suggests that Ed potentiates Dl-Notch signaling by promoting Dl endocytosis in signal-producing cells (see earlier discussion concerning Notch ligand endocytosis). Although it has not yet been observed during eye development, perhaps Notch and Ed cooperate to antagonize DER signaling during R8 cell spacing (Fig. 1-13C). While the role of DER in R8 founder cell formation remains uncertain, it is clear that the Ras/MAPK pathway is active at some level in the intermediate groups and in the cells immediately posterior. For example, in wildtype eye discs, dpMAPK accumulation in the intermediate groups is dependent upon signaling through DER (Baonza et al., 2001; Kumar et al., 69 2003; Kumar et al., 1998; Lesokhin et al., 1999; Rodrigues et al., 2005; Yang and Baker, 2001). DERnull clones and DERtsla clones lack dpMAPK in these clusters. In DERtsla clones, dpMAPK re-appears two hours later, whereas in DERnull clones, there is no re-appearance (Baonza et al., 2001; Kumar et al., 1998; Lesokhin et al., 1999; Rodrigues et al., 2005; Yang and Baker, 2001). However the functional significance of elevated dpMAPK expression in the intermediate groups has been a topic of debate. Those arguing for a role for DER in the spacing of R8, have posited that the DER pathway plays an active role in lateral inhibition alongside the Notch pathway and Sca, based on the reduction of Ato expression and subsequent loss of photoreceptors observed in either the gain-of-function DERElp allele or upon ectopic expression of Rhomboid (Baker and Rubin, 1989; Baonza et al., 2001; Chen and Chien, 1999; Dokucu et al., 1996; Dominguez et al., 1998; Lesokhin et al., 1999; Spencer et al., 1998; Yang and Baker, 2001; Zak and Shilo, 1992). The DER pathway is hypothesized to inhibit Ato and affect R8 spacing in two ways: via activation of the secreted DER antagonist Argos (Aos) and/or via the activation of the homeodomain transcription factor Rough (Ro) (Saint et al., 1988). Initially, it was shown that the DERElp phenotype was dependent on functional aos to exert its effect (Lesokhin et al., 1999). Aos ectopic overexpression was also shown to lead to the loss of Ato and aberrant R8 spacing, suggesting Aos might be a lateral inhibition signal (Spencer et al., 1998). In large aosnull mutant clones, R8 spacing is perturbed non-cell autonomously, confirming Aos’ spacing function (Yang and Baker, 2001). The DER pathway may also regulate spacing—at least indirectly— through Ro, a homeodomain protein that represses Ato expression and R8-identity in the R2/5 equivalence group (Dokucu et al., 1996). Supporting this idea, Ro expression is abolished in DERnull mutant clones (Dominguez et al., 1998). Together these data suggest that DER may play a parallel role with Notch lateral inhibition to focus R8 to a single cell within the proneural clusters. In contrast, the competing model argues that the DER pathway does not regulate R8 spacing (Kumar et al., 2003; Kumar et al., 1998; Rodrigues et al., 2005). The major tenet of this 70 argument is that the high expression of dpMAPK within the intermediate group does not lead to expression of downstream target genes because dpMAPK in these cells is prevented from entering the nucleus to phosphorylate target transcriptional regulators. Rather, dpMAPK is in a state of ‘cytoplasmic hold’ within the intermediate groups, as evidenced by immunohistological studies that localize dpMAPK primarily to the cytoplasm of these cells (Kumar et al., 2003; Kumar et al., 1998; Rodrigues et al., 2005). A system of tagged MAPK proteins to monitor MAPK translocation into the nucleus suggested that cytoplasmic sequestration of dpMAPK in the intermediate group is relevant biologically (Kumar et al., 2003). That is, if MAPK is forced into the nucleus in these cells, patterning is perturbed (Kumar et al., 2003). Therefore, dpMAPK localization appears to be developmentally regulated, begging the question of the mechanism behind the cytoplasmic hold. A prime candidate is the importin Msk which is required for import of dpMAPK into the nucleus and whose expression levels appear to be dynamically regulated in the intermediate group region (Lorenzen et al., 2001; Marenda et al., 2006; Vrailas et al., 2006; Vrailas and Moses, 2006). In the intermediate group, Msk is present at such low abundance that signaling from the DER/Ras pathway is blocked. Posterior to these groups, where Msk is no longer limiting, dpMAPK is escorted efficiently into the nucleus where it phosphorylates key transcriptional effectors (Vrailas et al., 2006; Vrailas and Moses, 2006). One caveat to the above model is that although dpMAPK is not permitted to enter the nucleus in the cluster, it is possible that cytoplasmic dpMAPK could phosphorylate substrates that would then transduce the signal into the nucleus. The identification of dpMAPK cytoplasmic substrates through biochemical analyses of cytplasmic extracts may offer insight into this question. Nevertheless, the debate concerning DERnull Minute experiments has left open the question whether DER functions during R8 spacing. In order to capitalize on DERnull allele analysis, it would be interesting to determine if a variation of the Minute technique that uses eye-specific overexpression of the pro-apoptotic gene head involution defective (hid) may overcome the problems outlined above. In this system, cells that carry the wild-type DER allele 71 undergo programmed cell death, consequently generating a tissue that is completely null for DER (Stowers and Schwarz, 1999). DER and Notch are required for cell cycle regulation during eye formation Cell proliferation is tightly regulated during the development of the eye, particularly in the region of the MF where cell division transitions from asynchrony to synchronous G1 arrest (Ready et al., 1976; Thomas et al., 1994; Wolff and Ready, 1991). Ahead of the MF in Zone I, cells divide asynchronously. In Zone II, adjacent to the MF, cells become increasingly synchronous with an increase in mitosis, leading to G1 arrest in the MF/Zone III (Heberlein and Moses, 1995; Thomas et al., 1994). Photoreceptor specification initiates with recruitment of the R8 founder cell followed by pair-wise recruitment of R2/5 and R3/4 (Fig. 1-14) (Ready et al., 1976). In Zone IV, undifferentiated cells that lie between these five cell clusters return to the cell cycle for a synchronous and terminal cell division leading into Zone V (Thomas et al., 1994). This return to mitosis is referred to as the ‘second mitotic wave’ (SMW; Fig. 1-14) (Thomas et al., 1994). The Notch and DER pathways cooperate to positively regulate progression of the SMW, although they influence distinct phases of the cell cycle (Fig. 1-14) (Baker and Yu, 2001; Baonza and Freeman, 2005; Firth et al., 2000; Thomas, 2005). Clonal analysis of Notch mutant tissue indicates that mitosis and DNA synthesis is reduced, as measured by phosphohistone H3 (PH3) staining and BrdU incorporation, respectively (Baonza and Freeman, 2005). These observations suggest that Notch is required for the entry into S phase. This requirement is dependent on signaling from N’s ligand Dl and also requires Su(H) (Baonza and Freeman, 2005; Firth et al., 2000). Mechanistically, the Notch pathway regulates the G1/S transition by activating transcription of dE2F1 and Cyclin A (CycA). E2Fs comprise a family of transcriptional activators important for S phase entry, while CycA, although typically involved in M phase, here plays a special role in G1/S transition (Fig. 1-14B) (Baonza and Freeman, 2005). The other cell division checkpoint for the progression of the SMW occurs at the G2/ 72 Fig. 1-14. The Second Mitotic Wave. Following recruitment of the first five photoreceptors, unspecified cells return to the cell cycle via Notch-mediated G1/S and DER-mediated G2/M transitions. Cells recruited into the ommatidium during this process are indicated within the arrows (PC = pigment cells). Adapted from Baonza and Freeman, 2005. M transition. Genetic analyses have revealed that G2/M progression is dependent on DER activation with a requirement for the transforming growth factor-α (TGF-α) ligand Spitz (Spi) (Baker and Yu, 2001). The cdc25 phosphatase String (Stg) is downstream of DER signaling and is required for entry into mitosis (Fig. 1-14) (Baker and Yu, 2001). Spi and DER mutant clone tissue is defective in Zone IV mitoses, measured by phosphohistone H3 staining and accumulation of the G2-cyclin cycB (Baker, 2001). Conversely, overexpression of activated DER or Ras is sufficient to promote mitoses in Zone IV (Baker and Yu, 2001). How are the developing photoreceptors in the five cell cluster prevented from responding to the signaling events that drive the SMW? R8, the ‘oldest” neuron in the cluster, is maintained in G1, independent of Notch and DER signaling (Baker, 2001; Yang and Baker, 2003), 73 presumably because it has achieved a level of differentiation that prevents it from responding to or requiring such signals. Notch is known to be active in R2/5/3/4, but the differentiating cells do not return to the cell cycle (Baker et al., 1996; Dokucu et al., 1996). Therefore, there must be a mechanism that prevents R2/5/3/4 from entering S phase again. One possibility is that the DER pathway provides this function. Supporting this idea, in the absence of the DER pathway member Son of sevenless (Sos), R2/5/3/4 re-enter the cell cycle. This effect is suppressed in cells mutant for both Sos and Su(H) (Firth et al., 2000). The molecular basis for DER’s antagonism of Notch in maintaining G1 arrest in R2/5/3/4 remains to be elucidated. It has been recently noted that although R2/5 and R3/4 are specified soon after R8, they do not begin to differentiate, as measured by expression of neuronal markers, until several hours later (Yang and Baker, 2006). Interestingly, the timing of this short delay, or hold on differentiation, corresponds exactly to when the unspecified cells progress through the SMW. These observations suggest that the differentiative block might ‘protect’ the specified photoreceptors from the pro-mitotic signaling events driving the SMW. This inhibition of differentiation is independent of cell cycle progression, but rather is dependent on the relationship of the Notch and DER pathways (Yang and Baker, 2006). DER signaling is required not only for the specification of R2/5 and R3/4 but also for their differentiation (Freeman, 1996; Yang and Baker, 2006). However, Notch signaling inhibits the differentiation potentiated by DER in this region as evidenced by the premature differentiation of R2/5 and R3/4 that occurs in Nts mutants (Yang and Baker, 2006). In summary, the Notch and DER pathways have distinct roles in the progression of the SMW. In the undifferentiated cells, the two pathways cooperate to promote the SMW, although they act sequentially at distinct points in the cell cycle, whereas in differentiating neurons, they act antagonistically to maintain G1 arrest in the photoreceptors and to block inappropriate differentiation in the cycling cells. Exactly how such context specificity is achieved remains an important area of future investigation. 74 Combinatorial factors in the determination and differentiation of ommatidial cell types The DER pathway is required for the sequential recruitment and patterning of most cell types within an ommatidium, except for the R8 founder cell and interommatidial bristle cells (Dominguez et al., 1998; Frankfort et al., 2004; Freeman, 1996; Tomlinson and Ready, 1987; Wolff and Ready, 1993). Following R8 specification, DER signals repeatedly to recruit the remaining ommatidiatial cells in a stereotyped sequence: first, the remaining seven photoreceptors in a pair-wise manner—R2 and 5, R3 and R4, R1 and R6, and finally R7; second, the four cone cells; third, the primary (1°) pigment cells; and, lastly, the secondary (2°) and tertiary (3°) pigment cells (Freeman, 1997). Each round of specification relies on secretion of Spitz ligand from the differentiating cells to activate DER signaling in adjacent uncommitted cells, thereby recruiting the next cell type into the ommatidium. The reiterative use of DER signaling to determine distinct cell fates in an ommatidium has provided an ideal context in which to study the question of specificity, namely how repeated use of the same signaling machinery can elicit distinct outcomes over time and space. For example, with respect to photoreceptors, the distinction between cell types is important for the expression of rhodopsin (rh) subclasses, which in turn is important for the visual discrimination capability of the animal. The outer photoreceptors (R2/5, R3/4, and R1/6) express rhodopsin rh1, whereas the inner photoreceptors express specific combinations of rh3, rh4, rh5, and rh6 (Cook and Desplan, 2001; Cook et al., 2003). In order to meet this need for specificity, the DER pathway works in-concert with other signaling pathways, especially the Notch pathway, along with a suite of pre-patterned transcriptional regulators to mediate specification of cell types in the developing disc (Freeman, 1997; Mollereau et al., 2001; Sundaram, 2005; Voas and Rebay, 2004). Therefore, a combinatorial code integrates DER-mediated inputs, pre-patterned determinants, and cell-type specific differentiation factors to regulate the identity of ommatidial cell types (Freeman, 1997; Hayashi and Saigo, 2001; Kumar and Moses, 1997). The ensuing discussion will focus on the contributions of DER/Notch-specific patterning events to specific aspects of this combinatorial code. A complete review of combinatorial signaling in the eye, 75 with consideration of the recruitment of each cell type and related patterning determinants, can be found in the following reference (Voas and Rebay, 2004). Planar polarity regulates R3 versus R4 identity Following recruitment of photoreceptors R8, R2, and R5, secretion of the DER ligand Spi from these newly specified cells recruits the presumptive R3/4 cells into a five-cell rosette cluster (Freeman, 1994b; Freeman, 1996; Tomlinson and Ready, 1987; Wolff and Ready, 1993). Expression of the homeotic gene Spalt (Sal), the orphan nuclear-receptor Seven-up (Svp) and the helix-turn-helix (HTH) transcription factor Pipsqueak (Psq) further defines the R3/R4 pair combinatorially with the DER pathway (Begemann et al., 1995; Domingos et al., 2004; Kramer et al., 1995; Mlodzik et al., 1990b). Below we discuss how refinement of R3/4 identity occurs through Notch-mediated mechanisms that regulate planar cell polarity (PCP) across the eye disc. The adult photoreceptors lie in a trapezoidal configuration arranged in two chiral forms to generate mirror-image symmetry across the DV boundary, or equator (Fig. 1-10B,C) (Tomlinson and Ready, 1987). Initially, R3 and R4 are positioned in parallel to the anterior/posterior axis of the eye disc, i.e. orthogonal to the equator, with R3 positioned closer to the equator, and R4 closer to the margins of the eye disc (Fig. 1-15A). Later during morphogenesis, the ommatidia rotate 90° to assume the trapezoidal orientation seen in adult ommatidia, such that the R3 points ‘up’ in the dorsal half or ‘down’ in the ventral half of the eye (Fig. 1-15A) (Cooper and Bray, 1999; Tomlinson and Ready, 1987). The Notch and DER pathways, along with the Frizzled (Fz) pathway, play distinct roles in coordinating the polarity and orientation of the ommatidia (Dominguez and de Celis, 1998; Irvine, 1999; McNeill, 2002; Wolff, 2003; Wu and Rao, 1999). Fz, a serpentine receptor-like transmembrane protein best known for its role in the canonical Wg/Armadillo (Arm)/β-Catenin (β-Cat) signaling pathway, also controls planar organization via an alternate pathway transduced through the cytoplasmic protein Dishevelled (Dsh) (Adler and Lee, 2001; McNeill, 2002; Mlodzik, 1999; Strutt, 2003; Vinson et al., 1989; Zheng et al., 1995). In the eye disc prior 76 77 Fig. 1-15. Planar cell polarity and ommatidial rotation. (A) At the five-cell cluster stage, R3 and R4 begin to have distinct cell fates. DER signaling mediates ommatidial rotation with respect to the equator. (B) R3 receives signal from the equator bound by Fz. Dsh, bound in a complex with the Ank-repeat protein Diego (Dgo) and the atypical cadherin Flamingo (Fmi), transduces the signal to activate Dl expression. Dl signals from R3 to R4 to activate the Notch pathway, R4 fate, and E(spl)-C repression of Dl. R4 Fz-Dsh signal transduction is inhibited by a complex that includes the four-pass transmembrane protein Strabismus (Stbm) and the LIMinteraction domain protein Prickle (Pk). Adapted from Strutt and Strutt, 2003. to ommatidial rotation, the presumptive R3 cell, distinguished from R4 by its more proximal position relative to the equator, receives a signal from the equator that activates the Fz pathway (Fig. 1-15B) (Fanto et al., 1998; Fanto and Mlodzik, 1999; Zheng et al., 1995). Transduction of the polarizing signal by Fz leads to activation of Dl in R3, which in turn signals to the presumptive R4. Notch activation, indicated by induction of E(spl)-mδ transcription, promotes commitment to the R4 fate (Cooper and Bray, 1999; Fanto and Mlodzik, 1999). In the absence of Notch signaling, such as in Notch or Dl mosaic mutant clones, R4 adopts the R3 fate, leading to ommatidial polarity defects. Once Fz and Notch signaling have distinguished the R3 and R4 fates, the DER pathway facilitates the subsequent rotation that orients the ommatidia to their correct position (Fig. 115A). Mutations in the DER pathway, including those in pnt, Ras, the DER ligand spi, and the signaling inhibitor aos, exhibit PCP defects (Brown and Freeman, 2003; Gaengel and Mlodzik, 2003; Strutt and Strutt, 2003; Wolff, 2003). Careful analysis of these mutants suggested that the DER pathway may be involved in two processes: during initial events that establish PCP and/or at latter stages of ommatidial rotation. One model suggests that DER pathway mutations might interfere with Fz rotational cues due to the presence of too few or many photoreceptors with decreased or increased DER signaling, respectively (Strutt, 2003). Alternatively, or in addition to involvement during the initial events that establish PCP, the DER pathway may work at later steps to regulate DE-cadherin and myosin II during the rotation process (Brown and Freeman, 2003; Gaengel and Mlodzik, 2003). Thus, the Notch and DER pathways act sequentially in a tightly orchestrated temporal sequence of events to establish ommatidial PCP whereby Notch regulates R3/4 cell fate decisions and DER regulates ommatidial rotation. R7 develops through orchestrated signaling by DER, Sev, and Notch The determination of the final photoreceptor recruited to the ommatidium, R7, has been well-characterized and provides one of the best understood examples of how integration of signals from multiple pathways results in specific cellular outcomes (Nagaraj and Banerjee, 78 2004). Historically, many of the initial discoveries that defined the components of the RTK/Ras/ MAPK pathway, were identified through characterization of the signaling mechanism initiated by interaction of the membrane-bound ligand Bride-of-sevenless (Boss), expressed on the surface of R8, with the RTK Sevenless (Sev), found on R7 (Fig. 1-16) (Dickson and Hafen, 1993). The subject of this discussion will be the pioneering insights derived from investigations into how Sev signaling is integrated with both the DER and Notch pathways to specify a unique R7 photoreceptor neuron. As we have emphasized throughout this review, an important question in developmental signal transduction is how specific cellular outcomes are achieved in response to reiterative use of common signaling pathways. Study of R7 development in the fly eye has provided one of the first definitive answers and the paradigms defined in this system are likely to be broadly relevant to other developmental contexts. The sev mutation was first identified as a defect in which the R7 photoreceptor was missing and its precursor had adopted a non-neuronal cone cell fate (Tomlinson and Ready, 1986). sev was cloned and identified to encode an RTK that is expressed in R3/4, R1/6, and R7, but is only required for the development of R7 (Banerjee et al., 1987a; Banerjee et al., 1987b; Basler and Hafen, 1988; Hafen et al., 1987; Simon et al., 1989; Tomlinson and Ready, 1987). The restricted requirement for Sev in R7 was revealed when the Sev ligand Boss was shown to regulate R7 development (Reinke and Zipursky, 1988). Boss is a seven-pass transmembrane ligand with a large extracellular domain that is localized specifically in R8 to restrict Boss-Sev signaling to induce R7 (Hart et al., 1990; Reinke and Zipursky, 1988; Van Vactor et al., 1991). Boss binds to Sev and is trans-endocytosed by Sev into the R7 cell (Cagan et al., 1992; Kramer et al., 1991). Like DER, ligand-receptor interaction leads to RTK activation via transand auto-phosphorylation on tyrosine residues (Fig. 1-16) (Fortini et al., 1992; Hart et al., 1990). The signaling machinery downstream of Sev is identical to that recruited by DER activation, and, in fact, the Ras/MAPK signaling cassette was initially characterized in Drosophila through genetic investigations of Sev-mediated control of R7 development (Biggs et al., 1994; Brunner 79 Fig. 1-16. Combinatorial signaling in R7. The Boss/Sev, Spi/DER and Dl/Notch signaling pathways promote R7 cell fate. et al., 1994a; Dickson et al., 1992; Fortini et al., 1992; Gebelein et al., 2004; Olivier et al., 1993; Rebay and Rubin, 1995; Simon et al., 1991; Simon et al., 1993). If Sev and DER both feed into the same signaling pathway and regulate the same target genes, then why are both required for the proper determination of R7? (Brunner et al., 1994b; Diaz-Benjumea and Hafen, 1994; Freeman, 1996). The answer appears to be that a significantly higher level of Ras/MAPK activation is required to overcome repressive mechanisms specific to R7 (Fig. 1-16). Supporting this idea, overexpression of one RTK has been shown to compensate for loss of the other. For example, if DER is overexpressed in sev mutants, then proper R7 specification is restored (Freeman, 1996). 80 Elevated levels of RTK signaling appear necessary to overcome two parallel mechanisms that negatively regulate R7 determination. The first, mediated by the Ets family repressor Yan, is not unique to R7 (Fig. 1-16). Although different cell fate specification events in the eye all require Ras/MAPK signaling to dismantle Yan-mediated repression of differentiation, the specific target genes involved are likely to be cell-type specific. Interestingly, regulation of yan expression appears to be an important point of cross-talk between the DER and Notch signaling pathways in this context (Rohrbaugh et al., 2002). Genetic interactions between yan and Su(H) or Notch provided the impetus for in vitro and in vivo studies that demonstrated a direct role for Su(H) in activating yan transcription in response to Notch activation (Rohrbaugh et al., 2002). Therefore, Notch promotes transcription of yan and thereby helps to establish a specific threshold for DER responsiveness during the precisely choreographed events of ommatidial assembly (Fig. 1-16). To allow each cell fate to be correctly specified, Ras/MAPK signaling must exceed this threshold. The second repressive mechanism involves the Zn-finger transcriptional repressor Tramtrack88 (Ttk88) (Lai et al., 1997; Lai et al., 1996; Xiong and Montell, 1993). Ttk88 blocks expression of specific neural fate promoting genes in the presumptive R7, most notably the homeodomain transcription factor prospero (pros) (Fig. 1-16) (Kauffmann et al., 1996; Xu et al., 2000). Expression of pros and other Ttk88 targets is derepressed by RTK pathway activation through expression of its target phyllopod (phyl) (Chang et al., 1995; Dickson et al., 1995; Lai et al., 1997; Lai et al., 1996; Xu et al., 2000). Phyl mediates degradation of Ttk88 through interaction with the R7 specific factor Seven in absentia (Sina)—a protein that interacts with a ubiquitin-conjugating enzyme for proteasome degradation (Fig. 1-16) (Li et al., 1997; Tang et al., 1997). Interestingly, and relevant to the discussion below, pros is also directly repressed by Yan, suggesting that coordinated remodeling of transcriptional complexes in response to upstream signaling events is central to proper R7 specification (Fig. 1-16). The mechanisms by which the DER, Sev, and Notch pathways function in the development of R7 is one of the best understood examples of signal integration and 81 combinatorial cell fate specification during development. The DER and Sev pathways play both direct and indirect roles in the transcriptional activation of pros in R7. Indirectly, as mentioned above, signaling through these RTK pathways dismantles both Yan and Ttk88mediated repression of pros. Acting more directly, the RTK pathway effector Pnt activates pros expression (Fig. 1-16) (Xu et al., 2000). Furthermore, adjacent to the four Ets-binding sites, the Runt/AML1/CBFA1-like transcription factor Lozenge (Lz), binds to two Lz-binding sites in the same pros enhancer (Fig. 1-16) (Daga et al., 1996; Flores et al., 1998; Xu et al., 2000). Thus, a combination of transcription factors is required for the activation of pros. The Notch pathway has a critical role in the development of R7 (Fig. 1-16). Activation of Notch signaling in R7 can be followed molecularly by onset of expression of a transgene containing an enhancer-reporter transcriptional fusion element for the Notch/Su(H)-target E(spl)mδ (Cooper et al., 2000). The genetic requirement for Notch was demonstrated when it was observed that a hypomorphic allele Nspl or temperature-sensitive Notch (Nts) mutants exhibited transformation of R7 cells to outer photoreceptor cells when placed at the restrictive temperature (Cooper et al., 2000; Van Vactor et al., 1991). Expression of a dominant repressor of Notch signal transduction [sev-Su(H)enR] in R1/6-7 similarly resulted in R7 to outer photoreceptor transformation (Tomlinson and Struhl, 2001). Conversely, targeted overexpression of activated Notch (Nact) in cells that include the outer photoreceptors, instructed these cells to adopt R7 characteristics (Cooper et al., 2000). Notch’s ability to inductively potentiate R7 development is dependent on expression of the Notch ligand Dl in R1/6 (Fig. 1-16) (Cooper and Bray, 2000; Tomlinson and Struhl, 2001). Therefore, the DER, Sev, and Notch pathways appear to cooperate in specifying the correct cell fate of the R7 photoreceptor. Moreover, the combinatorial control by DER and Notch suggests that target genes may have enhancer/promoter elements that may be bound by transcriptional effectors of both pathways. Such a model has been formally proven in the context of cone cell development, as described below. 82 The DER and Notch pathways determine cone cell fate The lens for each ommatidium is secreted by four cone cells that rely on DER signaling from the neighboring R cells for the determination of their fate (Freeman, 1996; Ready et al., 1976; Tomlinson and Ready, 1987). Analogous to the case of prospero expression in the R7 neuron, expression of dPax2, the Drosophila homolog of the vertebrate Pax2 gene and a marker of cone cell fate, is regulated through combinatorial transcriptional control of the lzspa minimal enhancer (SME) by Lozenge (Lz), the DER effectors Yan and PntP2, and the Notch pathway transcription factor Su(H) (Fig. 1-17) (Daga et al., 1996; Flores et al., 2000). Thus, insufficient DER or Notch signaling results in loss of dPax2 expression and elimination of cone cells (Flores et al., 2000; Fu et al., 1998; Fu and Noll, 1997). Cone cell development is dependent on inductive signaling through both the DER and Notch pathways. Spi and Dl are released from the photoreceptor (R) cells to activate signaling in the presumptive cone cells (Tsuda et al., 2006; Tsuda et al., 2002). Inductive Notch signaling is mediated by Ebi and the novel protein Strawberry notch (Sno) (Fig. 1-17B,D) (Tsuda et al., 2006; Tsuda et al., 2002). Ebi is the homolog of transducin β-like 1 (TBL1), a conserved F-box/ WD-40-repeat-containing protein that functions in co-repressor complexes but has recently been identified to convert transcriptional cofactors between repressive and stimulative states (Dong et al., 1999; Guenther et al., 2000; Yoon et al., 2003). Ebi and Sno influence the Notch pathway by regulating in R cells the transcription of mRNA encoding the Notch ligand Dl that is received by the cone cells (Fig. 1-17B,D) (Tsuda et al., 2006; Tsuda et al., 2002). A model has evolved for Dl regulation in R cells based on the following set of observations (Fig. 1-17): first, Ebi, Sno, and the DER pathway positively regulate Dl expression in R cells (Tsuda et al., 2006; Tsuda et al., 2002); second, the bifunctional transcription factor Su(H) switches between its role as a repressor and an activator (Tsuda et al., 2006; Tsuda et al., 2002); and third, Ebi and Sno interact with a repressor complex that contains Su(H), the Drosophila SMRT homolog Smrter (Smr), and HDACs (see discussion on Notch transcriptional repression) (Tsuda et al., 2006; Tsuda et al., 2002). 83 Fig. 1-17. DER and Notch inductive signaling in cone cells. In an initial model, (A) in unstimulated R cells, transcriptional repression of Dl through a Su(H)/Smr complex occurs in presumptive R cells. (B) DER activation in the R cell activates Ebi/Sno inhibition of the Su(H) repressor complex, permitting inductive signaling by Dl/Notch and Spi/DER to cone cells to activate dPax2. Adapted from (Lai, 2002a). The model was then revised to propose that (C) in unstimulated R cells, repression of Dl occurs through Chn. (D) chn expression is repressed by a Sno/Ebi/Su(H)/Smr complex. Adapted from Tsuda et al., 2006. 84 The model initially suggested that Dl transcription was inhibited by a Su(H)/Smr complex in R cells prior to inductive signaling (Fig. 1-17A,B) (Lai, 2002a; Tsuda et al., 2002). In support of this argument, Dl protein is greatly overexpressed in Su(H) mutant eye clones (Tsuda et al., 2002). Overexpression studies identified the DER pathway as acting in concert with Sno and Ebi to antagonize the Su(H)/Smr complex. The outcome of these interactions was proposed to derepress Su(H)/Smr mediated transcriptional inhibition of Dl (Tsuda et al., 2002). The mechanism by which the DER pathway cooperates with Sno and Ebi—transcriptional, post-transcriptional, direct, or parallel—is unclear, but the relationship was supported further by loss-of-function mutant analysis (Tsuda et al., 2002). On the other hand, the disruption of the Su(H)/Smr repressor complex was inferred from protein localization studies. In wild-type eye discs, Smr is localized predominantly to the cytoplasm, whereas in ebi or sno mutant eye discs, Smr is mostly nuclear. Thus, Ebi and Sno were proposed to promote Smr nuclear export and downregulation of Smr by the proteasome (Tsuda et al., 2002). Further, the ability of investigators to co-IP Sno, Ebi, Su(H), and Smr indicated the formation of a stable disruption complex. The above model concluded that Dl is regulated downstream of Su(H). However, it was unclear whether Su(H) represses Dl transcription directly through DNA-binding or indirectly through an intermediary. In seeking to identify how Dl was regulated, the model described above was revised. Specifically, new insight into the mechanistic context by which Su(H) functions to repress Dl led to the suggestion that in R cells prior to Dl expression, Su(H) functions as an activator of transcription in a complex with activated NotchICD (Fig. 1-17C) (Tsuda et al., 2006). However NotchICD/Su(H) does not regulate Dl directly, but rather potentiates the expression of Charlatan (Chn), a homolog of the vertebrate Neuronal Restricted Silencing transcription Factor (NRSF/REST), a well-characterized transcriptional repressor (Jones and Meech, 1999; Kania et al., 1995). Supporting this, in Su(H) mutant eye discs, chn expression was downregulated. Further, Su(H) was biochemically shown to bind to the Chn promoter by EMSA at degenerate Su(H)-binding sites (Tsuda et al., 2006). Activation by NotchICD/Su(H) leads to 85 Chn’s repression of Dl transcription through interaction with its cofactor dCoREST. Thus Dl expression is elevated in chn mutants and the Dl promoter contains two Chn-Binding Elements (CBEs) which suggests direct transcriptional regulation (Tsuda et al., 2006). This revised model of Dl repression is still consistent with a role of Su(H) to antagonize Dl, as described in the earlier model. In order for Dl expression to be de-repressed, NICD/Su(H)-mediated activation of chn must be countered (Fig. 1-17D). DER pathway activation leads Ebi (and presumably Sno) to form a complex with Su(H) and Smr to generate a repressor complex that inhibits chn expression through binding of the chn promoter. The NICD complex is then swapped for the Ebi/Smr complex, resulting in repression of chn transcription. Binding of each of the Su(H)containing complexes to the DNA is mutually exclusive as measured through ChIP (Tsuda et al., 2006). These data are consistent with the formation of a Sno/Ebi/Su(H)/Smr complex that acts to transcriptionally repress chn, subsequently leading to Dl derepression. Interestingly, this contrasts with the earlier hypothesis that this complex was disrupted in transcriptional repression capability. That is, Su(H) is now thought to switch from a transcriptional activator to a repressor—a contrast from the previous model where Su(H) repressor activity is disrupted by Ebi/Sno. This new model, on the other hand, does not address the earlier result that Sno/Ebi disrupts Smr localization—identified initially as a marker for the interruption of the Su(H)/Smr complex. Furthermore, the investigators did not relate this earlier result to propose a biological function of Smr protein re-localization seen in wildtype compared to ebi or sno mutant eye discs (Tsuda et al., 2002). A study of the genetic interactions between smr and ebi/sno may clarify these questions. In the above evolving model, the ensuing production of Dl from the R cells activates inductive Notch signaling in the presumptive cone cells. Due to reiterative DER signaling, Spi is also secreted from the R cells to activate the DER signaling pathway in the cone cells. This potentiates PntP2 transcriptional activity and occludes Yan transcriptional repression. Activated PntP2, NICD/Su(H), and Lz all converge to induce dPax2 expression via the SME enhancer 86 element, providing a clear mechanistic explanation for how DER and Notch signaling are integrated to regulate cone cell development (Fig. 1-17B,D) (Flores et al., 2000; Lai, 2002a; Tsuda et al., 2006; Tsuda et al., 2002). Moreover, the complex nature between DER and Notch signaling is underscored through their opposition in R cells and their cooperation in the adjacent cone cells. DER, Notch, Programmed Cell Death and the development of pigment cells Following cone cell recruitment, the DER and Notch signaling pathways are integral in the development of the pigment cells. Each ommatidium contains two 1° pigment cells surrounded by a hexagonal lattice of 2° and 3° pigment cells and bristle cells that are shared between ommatidia (Higashijima et al., 1992; Ready et al., 1976; Wolff and Ready, 1993). Thus, each ommatidium effectively has six 2° cells, three 3° cells, and three bristles (Miller and Cagan, 1998). Following R and cone cell determination, the pigment cells and the sensory bristle cells are determined from the undifferentiated pool of cells remaining in the eye discs. As previously discussed in the case of the photoreceptors, determination of accessory cell fates also relies on sequential rounds of DER signaling, with earlier specified cell types providing the secreted Spitz ligand. Here, we will discuss a different role for DER and Notch signaling, namely how they regulate survival versus apoptosis decisions during the final stages of ommatidial patterning. Cell death is essential during development to pattern and specify mature structures (Brachmann and Cagan, 2003; Hay et al., 2004; Rusconi et al., 2000; Steller and Grether, 1994; Twomey and McCarthy, 2005). Late in pupal development, after the final pigment cells have been recruited to the ommatidia, the remaining undifferentiated interommatidial cells undergo regulated programmed cell death (PCD) (Hay et al., 1994; Ready et al., 1976; Wolff and Ready, 1991). Failure to eliminate these surplus cells perturbs the ommatidial lattice structure and leads to a malformed adult eye. Again, the Notch and DER pathways play pivotal and spatially restricted roles in this final event (Baker, 2001; Rusconi et al., 2000). 87 The role of the Notch pathway in promoting cell death in this context was first suggested based on the finding that temperature-sensitive Nts mutants shifted to the restrictive temperature at late stages of eye development have supernumerary 2° pigment cells, a defect resulting from loss of PCD (Cagan and Ready, 1989). In a subsequent experiment, laser ablation of 1° cells caused the opposite phenotype, namely an increase in PCD and loss of 2° and 3° cells (Miller and Cagan, 1998). These results suggested that the 1° cells provide a signal that represses PCD and/or promotes survival of cells within the 2°/3° equivalence group, or interommatidial precursor cells (IPCs) (Miller and Cagan, 1998). The DER pathway feeds directly into this process of cell survival/death of IPCs by antagonizing PCD (Miller and Cagan, 1998). The DER pathway has been characterized as important for cell survival in many contexts throughout Drosophila development (Baker, 2001; Hay et al., 1995; Hay et al., 1994; Kurada and White, 1998; Sawamoto et al., 1998; Yu et al., 2002). Mechanistically, DER activation leads to MAPK-mediated phosphorylation and downregulation of the pro-apoptotic protein Hid, thereby preventing PCD (Bergmann et al., 1998; Kurada and White, 1998). Supporting such a role late in eye development, expression of constitutively-activated RasV12 at the time of pigment cell formation abrogates cell death. Similarly, loss of DER in the IPCs leads to loss of 2°/3° pigment cells and increased death (Freeman, 1996; Miller and Cagan, 1998). Signaling by the ligand Spi is thought to be derived from the 1° and cone cells, but this hypothesis is still in need of testing (Miller and Cagan, 1998). However, genetic epistasis analysis suggests that although Spi plays a role in survival, it may not provide the specific survival signal that effectively antagonizes Notch-mediated PCD of IPCs (Yu et al., 2002). Rather, Notch indirectly promotes PCD by overriding and antagonizing the downstream DER survival pathway (Yu et al., 2002). If Notch promotes PCD and DER inhibits it, how is the ‘winner’ determined? More specifically, how does Notch signaling override the DER survival signal in the undifferentiated IPCs but not in the specified pigment cells? The answers to such questions remain unclear. One possible answer might involve Aos, the inhibitory DER ligand (Sawamoto et al., 1994; 88 Sawamoto et al., 1998). If Notch signaling activated Aos expression, this would be sufficient to dampen DER output, resulting in death. Additional mechanistic insights may derive from analysis of mutations in other genes that alter the death/survival balance in the pupal eye, particularly with respect to how they interface with the DER and Notch pathways. One such example is klumpfuss (klu), a gene encoding a zinc finger transcription factor, whose loss leads to supernumerary 2˚ pigment cells (Rusconi et al., 2004). Klu, normally expressed in the IPCs, is necessary and sufficient for PCD of these cells (Rusconi et al., 2004). Dominant genetic interactions between klu mutations and hypermorphic DERelp or RasV12 suggest Klu opposes the DER pathway to promote PCD in the retina. (Rusconi et al., 2004). Furthermore, loss of klu in the retina induces greater activation of dpMAPK in the DER pathway, demonstrating molecularly the dampening of the signal (Rusconi et al., 2004). It will be interesting to determine how Klu functionally inhibits the DER pathway and whether this activity is regulated by the Notch pathway. Concluding Remarks We have only touched upon the signaling responsibilities for DER and Notch pathway interactions during metazoan development, but it is apparent from this discussion that these two signaling networks play an integral role in modulating each other’s activities. Although we have not discussed the interactions between DER and Notch outside of the eye, the common themes regarding how the two pathways are coordinated in time and space appear to be broadly conserved and directly applicable to many other developmental contexts (Carmena et al., 2002; Chandra et al., 2003; Diaz-Benjumea and Garcia-Bellido, 1990; Diaz-Benjumea and Hafen, 1994; Escudero et al., 2003; Islam et al., 2003; Jordan et al., 2000; Kojima, 2004; Schober et al., 2005; Walters et al., 2005; zur Lage and Jarman, 1999; zur Lage et al., 2004). Furthermore, the high degree of evolutionary conservation of the EGFR and Notch signaling pathways means that the highly regulated and choreographed interactions we have described in the fly eye, are relevant to understanding how antagonism, cooperation, and coordination of EGFR and Notch 89 contribute to vertebrate development, and by extension, how improper integration contributes to the development of human disease (Sundaram, 2005). Future progress in elucidating further the role of these two pathways will derive both from continuing the types of experiments that have led us to our current level of understanding and from implementing novel methodologies now available in the post-genomic era. For example, genetic screens have been pivotal in identifying novel factors, co-factors, or effectors that modulate the response of DER and Notch signaling in specific developmental contexts (Diaz-Benjumea and Garcia-Bellido, 1990; Price et al., 1997). These forward genetic techniques are not yet completely saturated, and are likely to continue to yield important discoveries well into the future. Continued emphasis on elucidating the biochemical and molecular nature of the DER/Notch interaction mechanisms suggested by the genetics will also remain a top priority. Simultaneously, incorporating functional genomic approaches such as global expression profiling and high-throughput RNAi screens to identify additional nodes of pathway cross-talk promises to shed new light into the intricacies underlying these two signaling pathways (DasGupta et al., 2005; Jordan et al., 2005; Lum et al., 2003; Nybakken et al., 2005; Wheeler et al., 2004; Wheeler et al., 2005). Acknowledgements We apologize to those whose work was not cited. We thank Jennifer C. Jemc and Matthew Freeman for helpful comments on the manuscript and to all members of the Rebay Lab for stimulating discussions. D.B.D. was supported by a National Science Foundation Graduate Research Fellowship. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. References Adler, P. N., and Lee, H. (2001). Frizzled signaling and cell-cell interactions in planar polarity. Curr Opin Cell Biol 13, 635-640. Ahmed, A., Chandra, S., Magarinos, M., and Vaessin, H. (2003). Echinoid mutants exhibit 90 neurogenic phenotypes and show synergistic interactions with the Notch signaling pathway. Development 130, 6295-6304. Aroian, R. V., Koga, M., Mendel, J. E., Ohshima, Y., and Sternberg, P. W. (1990). The let-23 gene necessary for Caenorhabditis elegans vulval induction encodes a tyrosine kinase of the EGF receptor subfamily. Nature 348, 693-699. Artavanis-Tsakonas, S., Rand, M. D., and Lake, R. J. (1999). Notch signaling: cell fate control and signal integration in development. Science 284, 770-776. Aster, J. C., Robertson, E. S., Hasserjian, R. P., Turner, J. R., Kieff, E., and Sklar, J. (1997). Oncogenic forms of NOTCH1 lacking either the primary binding site for RBP-Jkappa or nuclear localization sequences retain the ability to associate with RBP-Jkappa and activate transcription. J Biol Chem 272, 11336-11343. Avruch, J., Khokhlatchev, A., Kyriakis, J. M., Luo, Z., Tzivion, G., Vavvas, D., and Zhang, X. F. (2001). Ras activation of the Raf kinase: tyrosine kinase recruitment of the MAP kinase cascade. Recent Prog Horm Res 56, 127-155. Bai, J., Chiu, W., Wang, J., Tzeng, T., Perrimon, N., and Hsu, J. (2001). The cell adhesion molecule Echinoid defines a new pathway that antagonizes the Drosophila EGF receptor signaling pathway. Development 128, 591-601. Bailey, A. M., and Posakony, J. W. (1995). Suppressor of hairless directly activates transcription of enhancer of split complex genes in response to Notch receptor activity. Genes Dev 9, 2609-2622. Baker, D. A., Mille-Baker, B., Wainwright, S. M., Ish-Horowicz, D., and Dibb, N. J. (2001). Mae mediates MAP kinase phosphorylation of Ets transcription factors in Drosophila. Nature 411, 330-334. Baker, N. E. (2000). Notch signaling in the nervous system. Pieces still missing from the puzzle. Bioessays 22, 264-273. Baker, N. E. (2001). Cell proliferation, survival, and death in the Drosophila eye. Semin Cell Dev Biol 12, 499-507. Baker, N. E., Mlodzik, M., and Rubin, G. M. (1990). Spacing differentiation in the developing Drosophila eye: a fibrinogen-related lateral inhibitor encoded by scabrous. Science 250, 1370-1377. Baker, N. E., and Rubin, G. M. (1989). Effect on eye development of dominant mutations in Drosophila homologue of the EGF receptor. Nature 340, 150-153. Baker, N. E., Yu, S., and Han, D. (1996). Evolution of proneural atonal expression during distinct regulatory phases in the developing Drosophila eye. Curr Biol 6, 1290-1301. Baker, N. E., and Yu, S. Y. (1997). Proneural function of neurogenic genes in the developing Drosophila eye. Curr Biol 7, 122-132. Baker, N. E., and Yu, S. Y. (2001). The EGF receptor defines domains of cell cycle progression and survival to regulate cell number in the developing Drosophila eye. Cell 104, 699-708. Baker, N. E., and Zitron, A. E. (1995). Drosophila eye development: Notch and Delta amplify a neurogenic pattern conferred on the morphogenetic furrow by scabrous. Mech Dev 49, 173189. Baker, S. E., Lorenzen, J. A., Miller, S. W., Bunch, T. A., Jannuzi, A. L., Ginsberg, M. H., Perkins, L. A., and Brower, D. L. (2002). Genetic interaction between integrins and moleskin, a gene encoding a Drosophila homolog of importin-7. Genetics 162, 285-296. 91 Banerjee, U., Renfranz, P. J., Hinton, D. R., Rabin, B. A., and Benzer, S. (1987a). The sevenless+ protein is expressed apically in cell membranes of developing Drosophila retina; it is not restricted to cell R7. Cell 51, 151-158. Banerjee, U., Renfranz, P. J., Pollock, J. A., and Benzer, S. (1987b). Molecular characterization and expression of sevenless, a gene involved in neuronal pattern formation in the Drosophila eye. Cell 49, 281-291. Bang, A. G., Bailey, A. M., and Posakony, J. W. (1995). Hairless promotes stable commitment to the sensory organ precursor cell fate by negatively regulating the activity of the Notch signaling pathway. Dev Biol 172, 479-494. Bang, A. G., and Kintner, C. (2000). Rhomboid and Star facilitate presentation and processing of the Drosophila TGF-alpha homolog Spitz. Genes Dev 14, 177-186. Baonza, A., Casci, T., and Freeman, M. (2001). A primary role for the epidermal growth factor receptor in ommatidial spacing in the Drosophila eye. Curr Biol 11, 396-404. Baonza, A., and Freeman, M. (2001). Notch signalling and the initiation of neural development in the Drosophila eye. Development 128, 3889-3898. Baonza, A., and Freeman, M. (2005). Control of cell proliferation in the Drosophila eye by Notch signaling. Dev Cell 8, 529-539. Barbacid, M. (1987). ras genes. Annu Rev Biochem 56, 779-827. Bardin, A. J., and Schweisguth, F. (2006). Bearded family members inhibit neuralized-mediated endocytosis and signaling activity of delta in Drosophila. Dev Cell 10, 245-255. Barolo, S., Stone, T., Bang, A. G., and Posakony, J. W. (2002). Default repression and Notch signaling: Hairless acts as an adaptor to recruit the corepressors Groucho and dCtBP to Suppressor of Hairless. Genes Dev 16, 1964-1976. Barolo, S., Walker, R. G., Polyanovsky, A. D., Freschi, G., Keil, T., and Posakony, J. W. (2000). A notch-independent activity of suppressor of hairless is required for normal mechanoreceptor physiology. Cell 103, 957-969. Baron, M. (2003). An overview of the Notch signalling pathway. Semin Cell Dev Biol 14, 113119. Barrick, D., and Kopan, R. (2006). The notch transcription activation complex makes its move. Cell 124, 883-885. Basler, K., and Hafen, E. (1988). Control of photoreceptor cell fate by the sevenless protein requires a functional tyrosine kinase domain. Cell 54, 299-311. Bazley, L. A., and Gullick, W. J. (2005). The epidermal growth factor receptor family. Endocr Relat Cancer 12 Suppl 1, S17-27. Beatus, P., and Lendahl, U. (1998). Notch and neurogenesis. J Neurosci Res 54, 125-136. Begemann, G., Michon, A. M., vd Voorn, L., Wepf, R., and Mlodzik, M. (1995). The Drosophila orphan nuclear receptor seven-up requires the Ras pathway for its function in photoreceptor determination. Development 121, 225-235. Berdnik, D., Torok, T., Gonzalez-Gaitan, M., and Knoblich, J. A. (2002). The endocytic protein alpha-Adaptin is required for numb-mediated asymmetric cell division in Drosophila. Dev Cell 3, 221-231. Bergmann, A., Agapite, J., McCall, K., and Steller, H. (1998). The Drosophila gene hid is a direct molecular target of Ras-dependent survival signaling. Cell 95, 331-341. Bessa, J., Gebelein, B., Pichaud, F., Casares, F., and Mann, R. S. (2002). Combinatorial control of Drosophila eye development by Eyeless, Homothorax, and Teashirt. Genes Dev 16, 241592 2427. Bier, E., Jan, L. Y., and Jan, Y. N. (1990). rhomboid, a gene required for dorsoventral axis establishment and peripheral nervous system development in Drosophila melanogaster. Genes Dev 4, 190-203. Biggs, W. H., 3rd, Zavitz, K. H., Dickson, B., van der Straten, A., Brunner, D., Hafen, E., and Zipursky, S. L. (1994). The Drosophila rolled locus encodes a MAP kinase required in the sevenless signal transduction pathway. Embo J 13, 1628-1635. Black, R. A. (2002). Tumor necrosis factor-alpha converting enzyme. Int J Biochem Cell Biol 34, 1-5. Bland, C. E., Kimberly, P., and Rand, M. D. (2003). Notch-induced proteolysis and nuclear localization of the Delta ligand. J Biol Chem 278, 13607-13610. Boriack-Sjodin, P. A., Margarit, S. M., Bar-Sagi, D., and Kuriyan, J. (1998). The structural basis of the activation of Ras by Sos. Nature 394, 337-343. Bork, P. (1993). Hundreds of ankyrin-like repeats in functionally diverse proteins: mobile modules that cross phyla horizontally? Proteins 17, 363-374. Brachmann, C. B., and Cagan, R. L. (2003). Patterning the fly eye: the role of apoptosis. Trends Genet 19, 91-96. Bray, S. (1998). Notch signalling in Drosophila: three ways to use a pathway. Semin Cell Dev Biol 9, 591-597. Breeden, L., and Nasmyth, K. (1987). Similarity between cell-cycle genes of budding yeast and fission yeast and the Notch gene of Drosophila. Nature 329, 651-654. Brou, C., Logeat, F., Gupta, N., Bessia, C., LeBail, O., Doedens, J. R., Cumano, A., Roux, P., Black, R. A., and Israel, A. (2000). A novel proteolytic cleavage involved in Notch signaling: the role of the disintegrin-metalloprotease TACE. Mol Cell 5, 207-216. Brown, K. E., and Freeman, M. (2003). Egfr signalling defines a protective function for ommatidial orientation in the Drosophila eye. Development 130, 5401-5412. Brown, N. L., Paddock, S. W., Sattler, C. A., Cronmiller, C., Thomas, B. J., and Carroll, S. B. (1996). daughterless is required for Drosophila photoreceptor cell determination, eye morphogenesis, and cell cycle progression. Dev Biol 179, 65-78. Brown, N. L., Sattler, C. A., Paddock, S. W., and Carroll, S. B. (1995). Hairy and emc negatively regulate morphogenetic furrow progression in the Drosophila eye. Cell 80, 879-887. Bruckner, K., Perez, L., Clausen, H., and Cohen, S. (2000). Glycosyltransferase activity of Fringe modulates Notch-Delta interactions. Nature 406, 411-415. Brunner, D., Ducker, K., Oellers, N., Hafen, E., Scholz, H., and Klambt, C. (1994a). The ETS domain protein pointed-P2 is a target of MAP kinase in the sevenless signal transduction pathway. Nature 370, 386-389. Brunner, D., Oellers, N., Szabad, J., Biggs, W. H., 3rd, Zipursky, S. L., and Hafen, E. (1994b). A gain-of-function mutation in Drosophila MAP kinase activates multiple receptor tyrosine kinase signaling pathways. Cell 76, 875-888. Buday, L., and Downward, J. (1993). Epidermal growth factor regulates p21ras through the formation of a complex of receptor, Grb2 adapter protein, and Sos nucleotide exchange factor. Cell 73, 611-620. Cagan, R. L., Kramer, H., Hart, A. C., and Zipursky, S. L. (1992). The bride of sevenless and sevenless interaction: internalization of a transmembrane ligand. Cell 69, 393-399. Cagan, R. L., and Ready, D. F. (1989). Notch is required for successive cell decisions in the 93 developing Drosophila retina. Genes Dev 3, 1099-1112. Carmena, A., Buff, E., Halfon, M. S., Gisselbrecht, S., Jimenez, F., Baylies, M. K., and Michelson, A. M. (2002). Reciprocal regulatory interactions between the Notch and Ras signaling pathways in the Drosophila embryonic mesoderm. Dev Biol 244, 226-242. Castro, B., Barolo, S., Bailey, A. M., and Posakony, J. W. (2005). Lateral inhibition in proneural clusters: cis-regulatory logic and default repression by Suppressor of Hairless. Development 132, 3333-3344. Cavodeassi, F., Diez Del Corral, R., Campuzano, S., and Dominguez, M. (1999). Compartments and organising boundaries in the Drosophila eye: the role of the homeodomain Iroquois proteins. Development 126, 4933-4942. Cavodeassi, F., Modolell, J., and Gomez-Skarmeta, J. L. (2001). The Iroquois family of genes: from body building to neural patterning. Development 128, 2847-2855. Cervantes, S., Saura, C. A., Pomares, E., Gonzalez-Duarte, R., and Marfany, G. (2004). Functional implications of the presenilin dimerization: reconstitution of gamma-secretase activity by assembly of a catalytic site at the dimer interface of two catalytically inactive presenilins. J Biol Chem 279, 36519-36529. Chandra, S., Ahmed, A., and Vaessin, H. (2003). The Drosophila IgC2 domain protein Friend-ofEchinoid, a paralogue of Echinoid, limits the number of sensory organ precursors in the wing disc and interacts with the Notch signaling pathway. Dev Biol 256, 302-316. Chandu, D., Huppert, S. S., and Kopan, R. (2006). Analysis of transmembrane domain mutants is consistent with sequential cleavage of Notch by gamma-secretase. J Neurochem 96, 228-235. Chang, H. C., Solomon, N. M., Wassarman, D. A., Karim, F. D., Therrien, M., Rubin, G. M., and Wolff, T. (1995). phyllopod functions in the fate determination of a subset of photoreceptors in Drosophila. Cell 80, 463-472. Chanut, F., and Heberlein, U. (1997). Role of decapentaplegic in initiation and progression of the morphogenetic furrow in the developing Drosophila retina. Development 124, 559-567. Chao, J. L., Tsai, Y. C., Chiu, S. J., and Sun, Y. H. (2004). Localized Notch signal acts through eyg and upd to promote global growth in Drosophila eye. Development 131, 3839-3847. Chen, C. K., and Chien, C. T. (1999). Negative regulation of atonal in proneural cluster formation of Drosophila R8 photoreceptors. Proc Natl Acad Sci U S A 96, 5055-5060. Chen, X., Zhang, B., and Fischer, J. A. (2002). A specific protein substrate for a deubiquitinating enzyme: Liquid facets is the substrate of Fat facets. Genes Dev 16, 289-294. Chitnis, A. (2006). Why is delta endocytosis required for effective activation of notch? Dev Dyn 235, 886-894. Cho, K. O., and Choi, K. W. (1998). Fringe is essential for mirror symmetry and morphogenesis in the Drosophila eye. Nature 396, 272-276. Chong, H., Vikis, H. G., and Guan, K. L. (2003). Mechanisms of regulating the Raf kinase family. Cell Signal 15, 463-469. Chou, Y. H., and Chien, C. T. (2002). Scabrous controls ommatidial rotation in the Drosophila compound eye. Dev Cell 3, 839-850. Chung, H. M., and Struhl, G. (2001). Nicastrin is required for Presenilin-mediated transmembrane cleavage in Drosophila. Nat Cell Biol 3, 1129-1132. Cohen, S. M. (1993). Imaginal disc development. In The Development of Drosophila melanogaster, M. Bate, and A. Martinez Arias, eds. (Cold Spring Harbor Laboratory Press), pp. 747-841. 94 Cook, T., and Desplan, C. (2001). Photoreceptor subtype specification: from flies to humans. Semin Cell Dev Biol 12, 509-518. Cook, T., Pichaud, F., Sonneville, R., Papatsenko, D., and Desplan, C. (2003). Distinction between Color Photoreceptor Cell Fates Is Controlled by Prospero in Drosophila. Dev Cell 4, 853-864. Cooper, M. T., and Bray, S. J. (1999). Frizzled regulation of Notch signalling polarizes cell fate in the Drosophila eye. Nature 397, 526-530. Cooper, M. T., and Bray, S. J. (2000). R7 photoreceptor specification requires Notch activity. Curr Biol 10, 1507-1510. Cooper, M. T., Tyler, D. M., Furriols, M., Chalkiadaki, A., Delidakis, C., and Bray, S. (2000). Spatially restricted factors cooperate with notch in the regulation of Enhancer of split genes. Dev Biol 221, 390-403. Cornell, M., Evans, D. A., Mann, R., Fostier, M., Flasza, M., Monthatong, M., ArtavanisTsakonas, S., and Baron, M. (1999). The Drosophila melanogaster Suppressor of deltex gene, a regulator of the Notch receptor signaling pathway, is an E3 class ubiquitin ligase. Genetics 152, 567-576. Crews, C. M., Alessandrini, A., and Erikson, R. L. (1992). The primary structure of MEK, a protein kinase that phosphorylates the ERK gene product. Science 258, 478-480. Crews, C. M., and Erikson, R. L. (1992). Purification of a murine protein-tyrosine/threonine kinase that phosphorylates and activates the Erk-1 gene product: relationship to the fission yeast byr1 gene product. Proc Natl Acad Sci U S A 89, 8205-8209. Cubas, P., de Celis, J. F., Campuzano, S., and Modolell, J. (1991). Proneural clusters of achaetescute expression and the generation of sensory organs in the Drosophila imaginal wing disc. Genes Dev 5, 996-1008. Culi, J., and Modolell, J. (1998). Proneural gene self-stimulation in neural precursors: an essential mechanism for sense organ development that is regulated by Notch signaling. Genes Dev 12, 2036-2047. Cyert, M. S. (2001). Regulation of nuclear localization during signaling. J Biol Chem 276, 20805-20808. Daga, A., Karlovich, C. A., Dumstrei, K., and Banerjee, U. (1996). Patterning of cells in the Drosophila eye by Lozenge, which shares homologous domains with AML1. Genes Dev 10, 1194-1205. DasGupta, R., Kaykas, A., Moon, R. T., and Perrimon, N. (2005). Functional genomic analysis of the Wnt-wingless signaling pathway. Science 308, 826-833. De Strooper, B. (2003). Aph-1, Pen-2, and Nicastrin with Presenilin generate an active gammaSecretase complex. Neuron 38, 9-12. De Strooper, B., Annaert, W., Cupers, P., Saftig, P., Craessaerts, K., Mumm, J. S., Schroeter, E. H., Schrijvers, V., Wolfe, M. S., Ray, W. J., et al. (1999). A presenilin-1-dependent gammasecretase-like protease mediates release of Notch intracellular domain. Nature 398, 518-522. Deblandre, G. A., Lai, E. C., and Kintner, C. (2001). Xenopus neuralized is a ubiquitin ligase that interacts with XDelta1 and regulates Notch signaling. Dev Cell 1, 795-806. Delidakis, C., and Artavanis-Tsakonas, S. (1992). The Enhancer of split [E(spl)] locus of Drosophila encodes seven independent helix-loop-helix proteins. Proc Natl Acad Sci U S A 89, 8731-8735. Dhillon, A. S., Pollock, C., Steen, H., Shaw, P. E., Mischak, H., and Kolch, W. (2002). Cyclic 95 AMP-dependent kinase regulates Raf-1 kinase mainly by phosphorylation of serine 259. Mol Cell Biol 22, 3237-3246. Diaz-Benjumea, F. J., and Garcia-Bellido, A. (1990). Behaviour of cells mutant for an EGF receptor homologue of Drosophila in genetic mosaics. Proc Biol Sci 242, 36-44. Diaz-Benjumea, F. J., and Hafen, E. (1994). The sevenless signalling cassette mediates Drosophila EGF receptor function during epidermal development. Development 120, 569578. Dickson, B., and Hafen, E. (1993). Genetic dissection of eye development in Drosophila. In The Development of Drosophila melanogaster, M. Bate, and A. Martinez Arias, eds. (Cold Spring Harbor Laboratory Press), pp. 1327-1362. Dickson, B., Sprenger, F., Morrison, D., and Hafen, E. (1992). Raf functions downstream of Ras1 in the Sevenless signal transduction pathway. Nature 360, 600-603. Dickson, B. J., Dominguez, M., van der Straten, A., and Hafen, E. (1995). Control of Drosophila photoreceptor cell fates by phyllopod, a novel nuclear protein acting downstream of the Raf kinase. Cell 80, 453-462. Doe, C. Q., and Goodman, C. S. (1985). Early events in insect neurogenesis. II. The role of cell interactions and cell lineage in the determination of neuronal precursor cells. Dev Biol 111, 206-219. Doherty, D., Feger, G., Younger-Shepherd, S., Jan, L. Y., and Jan, Y. N. (1996). Delta is a ventral to dorsal signal complementary to Serrate, another Notch ligand, in Drosophila wing formation. Genes Dev 10, 421-434. Dokucu, M. E., Zipursky, S. L., and Cagan, R. L. (1996). Atonal, rough and the resolution of proneural clusters in the developing Drosophila retina. Development 122, 4139-4147. Domingos, P. M., Mlodzik, M., Mendes, C. S., Brown, S., Steller, H., and Mollereau, B. (2004). Spalt transcription factors are required for R3/R4 specification and establishment of planar cell polarity in the Drosophila eye. Development 131, 5695-5702. Dominguez, M., and Casares, F. (2005). Organ specification-growth control connection: new insights from the Drosophila eye-antennal disc. Dev Dyn 232, 673-684. Dominguez, M., and de Celis, J. F. (1998). A dorsal/ventral boundary established by Notch controls growth and polarity in the Drosophila eye. Nature 396, 276-278. Dominguez, M., Ferres-Marco, D., Gutierrez-Avino, F. J., Speicher, S. A., and Beneyto, M. (2004). Growth and specification of the eye are controlled independently by Eyegone and Eyeless in Drosophila melanogaster. Nat Genet 36, 31-39. Dominguez, M., Wasserman, J. D., and Freeman, M. (1998). Multiple functions of the EGF receptor in Drosophila eye development. Curr Biol 8, 1039-1048. Dong, X., Tsuda, L., Zavitz, K. H., Lin, M., Li, S., Carthew, R. W., and Zipursky, S. L. (1999). ebi regulates epidermal growth factor receptor signaling pathways in Drosophila. Genes Dev 13, 954-965. Dou, S., Zeng, X., Cortes, P., Erdjument-Bromage, H., Tempst, P., Honjo, T., and Vales, L. D. (1994). The recombination signal sequence-binding protein RBP-2N functions as a transcriptional repressor. Mol Cell Biol 14, 3310-3319. Downward, J., Yarden, Y., Mayes, E., Scrace, G., Totty, N., Stockwell, P., Ullrich, A., Schlessinger, J., and Waterfield, M. D. (1984). Close similarity of epidermal growth factor receptor and v-erb-B oncogene protein sequences. Nature 307, 521-527. Ellis, H. M., Spann, D. R., and Posakony, J. W. (1990). extramacrochaetae, a negative regulator 96 of sensory organ development in Drosophila, defines a new class of helix-loop-helix proteins. Cell 61, 27-38. Escudero, L. M., Wei, S. Y., Chiu, W. H., Modolell, J., and Hsu, J. C. (2003). Echinoid synergizes with the Notch signaling pathway in Drosophila mesothorax bristle patterning. Development 130, 6305-6316. Fanto, M., Mayes, C. A., and Mlodzik, M. (1998). Linking cell-fate specification to planar polarity: determination of the R3/R4 photoreceptors is a prerequisite for the interpretation of the Frizzled mediated polarity signal. Mech Dev 74, 51-58. Fanto, M., and Mlodzik, M. (1999). Asymmetric Notch activation specifies photoreceptors R3 and R4 and planar polarity in the Drosophila eye. Nature 397, 523-526. Fehon, R. G., Kooh, P. J., Rebay, I., Regan, C. L., Xu, T., Muskavitch, M. A., and ArtavanisTsakonas, S. (1990). Molecular interactions between the protein products of the neurogenic loci Notch and Delta, two EGF-homologous genes in Drosophila. Cell 61, 523-534. Ferrus, A. (1975). Parameters of mitotic recombination in minute mutants of Drosophila melanogaster. Genetics 79, 589-599. Firth, L., Manchester, J., Lorenzen, J. A., Baron, M., and Perkins, L. A. (2000). Identification of genomic regions that interact with a viable allele of the Drosophila protein tyrosine phosphatase corkscrew. Genetics 156, 733-748. Firth, L. C., Li, W., Zhang, H., and Baker, N. E. (2005). Analyses of RAS Regulation of Eye Development in Drosophila melanogaster. Methods Enzymol 407, 711-721. Fleming, R. J., Gu, Y., and Hukriede, N. A. (1997). Serrate-mediated activation of Notch is specifically blocked by the product of the gene fringe in the dorsal compartment of the Drosophila wing imaginal disc. Development 124, 2973-2981. Fleming, R. J., Scottgale, T. N., Diederich, R. J., and Artavanis-Tsakonas, S. (1990). The gene Serrate encodes a putative EGF-like transmembrane protein essential for proper ectodermal development in Drosophila melanogaster. Genes Dev 4, 2188-2201. Flores, G. V., Daga, A., Kalhor, H. R., and Banerjee, U. (1998). Lozenge is expressed in pluripotent precursor cells and patterns multiple cell types in the Drosophila eye through the control of cell-specific transcription factors. Development 125, 3681-3687. Flores, G. V., Duan, H., Yan, H., Nagaraj, R., Fu, W., Zou, Y., Noll, M., and Banerjee, U. (2000). Combinatorial signaling in the specification of unique cell fates. Cell 103, 75-85. Fortini, M. E., and Artavanis-Tsakonas, S. (1994). The suppressor of hairless protein participates in notch receptor signaling. Cell 79, 273-282. Fortini, M. E., Simon, M. A., and Rubin, G. M. (1992). Signalling by the sevenless protein tyrosine kinase is mimicked by Ras1 activation. Nature 355, 559-561. Francis, R., McGrath, G., Zhang, J., Ruddy, D. A., Sym, M., Apfeld, J., Nicoll, M., Maxwell, M., Hai, B., Ellis, M. C., et al. (2002). aph-1 and pen-2 are required for Notch pathway signaling, gamma-secretase cleavage of betaAPP, and presenilin protein accumulation. Dev Cell 3, 8597. Frankfort, B. J., and Mardon, G. (2002). R8 development in the Drosophila eye: a paradigm for neural selection and differentiation. Development 129, 1295-1306. Frankfort, B. J., Pepple, K. L., Mamlouk, M., Rose, M. F., and Mardon, G. (2004). Senseless is required for pupal retinal development in Drosophila. Genesis 38, 182-194. Freeman, M. (1994a). Misexpression of the Drosophila argos gene, a secreted regulator of cell determination. Development 120, 2297-2304. 97 Freeman, M. (1994b). The spitz gene is required for photoreceptor determination in the Drosophila eye where it interacts with the EGF receptor. Mech Dev 48, 25-33. Freeman, M. (1996). Reiterative use of the EGF receptor triggers differentiation of all cell types in the Drosophila eye. Cell 87, 651-660. Freeman, M. (1997). Cell determination strategies in the Drosophila eye. Development 124, 261270. Freeman, M., Klambt, C., Goodman, C. S., and Rubin, G. M. (1992). The argos gene encodes a diffusible factor that regulates cell fate decisions in the Drosophila eye. Cell 69, 963-975. Fryer, C. J., Lamar, E., Turbachova, I., Kintner, C., and Jones, K. A. (2002). Mastermind mediates chromatin-specific transcription and turnover of the Notch enhancer complex. Genes Dev 16, 1397-1411. Fu, W., and Baker, N. E. (2003). Deciphering synergistic and redundant roles of Hedgehog, Decapentaplegic and Delta that drive the wave of differentiation in Drosophila eye development. Development 130, 5229-5239. Fu, W., Duan, H., Frei, E., and Noll, M. (1998). shaven and sparkling are mutations in separate enhancers of the Drosophila Pax2 homolog. Development 125, 2943-2950. Fu, W., and Noll, M. (1997). The Pax2 homolog sparkling is required for development of cone and pigment cells in the Drosophila eye. Genes Dev 11, 2066-2078. Furriols, M., and Bray, S. (2000). Dissecting the mechanisms of suppressor of hairless function. Dev Biol 227, 520-532. Furukawa, T., Kobayakawa, Y., Tamura, K., Kimura, K., Kawaichi, M., Tanimura, T., and Honjo, T. (1995). Suppressor of hairless, the Drosophila homologue of RBP-J kappa, transactivates the neurogenic gene E(spl)m8. Jpn J Genet 70, 505-524. Gaengel, K., and Mlodzik, M. (2003). Egfr signaling regulates ommatidial rotation and cell motility in the Drosophila eye via MAPK/Pnt signaling and the Ras effector Canoe/AF6. Development 130, 5413-5423. Garratt, A. N., Britsch, S., and Birchmeier, C. (2000). Neuregulin, a factor with many functions in the life of a schwann cell. Bioessays 22, 987-996. Garrell, J., and Modolell, J. (1990). The Drosophila extramacrochaetae locus, an antagonist of proneural genes that, like these genes, encodes a helix-loop-helix protein. Cell 61, 39-48. Gebelein, B., McKay, D. J., and Mann, R. S. (2004). Direct integration of Hox and segmentation gene inputs during Drosophila development. Nature 431, 653-659. Gibert, J. M., and Simpson, P. (2003). Evolution of cis-regulation of the proneural genes. Int J Dev Biol 47, 643-651. Gigliani, F., Longo, F., Gaddini, L., and Battaglia, P. A. (1996). Interactions among the bHLH domains of the proteins encoded by the Enhancer of split and achaete-scute gene complexes of Drosophila. Mol Gen Genet 251, 628-634. Golembo, M., Schweitzer, R., Freeman, M., and Shilo, B. Z. (1996). Argos transcription is induced by the Drosophila EGF receptor pathway to form an inhibitory feedback loop. Development 122, 223-230. Golub, T. R., Barker, G. F., Lovett, M., and Gilliland, D. G. (1994). Fusion of PDGF receptor beta to a novel ets-like gene, tel, in chronic myelomonocytic leukemia with t(5;12) chromosomal translocation. Cell 77, 307-316. Goutte, C., Hepler, W., Mickey, K. M., and Priess, J. R. (2000). aph-2 encodes a novel extracellular protein required for GLP-1-mediated signaling. Development 127, 2481-2492. 98 Goutte, C., Tsunozaki, M., Hale, V. A., and Priess, J. R. (2002). APH-1 is a multipass membrane protein essential for the Notch signaling pathway in Caenorhabditis elegans embryos. Proc Natl Acad Sci U S A 99, 775-779. Greenwald, I. (1994). Structure/function studies of lin-12/Notch proteins. Curr Opin Genet Dev 4, 556-562. Greenwood, S., and Struhl, G. (1999). Progression of the morphogenetic furrow in the Drosophila eye: the roles of Hedgehog, Decapentaplegic and the Raf pathway. Development 126, 5795-5808. Guenther, M. G., Lane, W. S., Fischle, W., Verdin, E., Lazar, M. A., and Shiekhattar, R. (2000). A core SMRT corepressor complex containing HDAC3 and TBL1, a WD40-repeat protein linked to deafness. Genes Dev 14, 1048-1057. Guo, M., Jan, L. Y., and Jan, Y. N. (1996). Control of daughter cell fates during asymmetric division: interaction of Numb and Notch. Neuron 17, 27-41. Gupta-Rossi, N., Six, E., LeBail, O., Logeat, F., Chastagner, P., Olry, A., Israel, A., and Brou, C. (2004). Monoubiquitination and endocytosis direct gamma-secretase cleavage of activated Notch receptor. J Cell Biol 166, 73-83. Haddon, C., Jiang, Y. J., Smithers, L., and Lewis, J. (1998). Delta-Notch signalling and the patterning of sensory cell differentiation in the zebrafish ear: evidence from the mind bomb mutant. Development 125, 4637-4644. Hafen, E., Basler, K., Edstroem, J. E., and Rubin, G. M. (1987). Sevenless, a cell-specific homeotic gene of Drosophila, encodes a putative transmembrane receptor with a tyrosine kinase domain. Science 236, 55-63. Haglund, K., Di Fiore, P. P., and Dikic, I. (2003). Distinct monoubiquitin signals in receptor endocytosis. Trends Biochem Sci 28, 598-603. Haines, N., and Irvine, K. D. (2003). Glycosylation regulates Notch signalling. Nat Rev Mol Cell Biol 4, 786-797. Hall, B. E., Yang, S. S., Boriack-Sjodin, P. A., Kuriyan, J., and Bar-Sagi, D. (2001). Structurebased mutagenesis reveals distinct functions for Ras switch 1 and switch 2 in Sos-catalyzed guanine nucleotide exchange. J Biol Chem 276, 27629-27637. Haltiwanger, R. S. (2002). Regulation of signal transduction pathways in development by glycosylation. Curr Opin Struct Biol 12, 593-598. Haltiwanger, R. S., and Stanley, P. (2002). Modulation of receptor signaling by glycosylation: fringe is an O-fucose-beta1,3-N-acetylglucosaminyltransferase. Biochim Biophys Acta 1573, 328-335. Hart, A. C., Kramer, H., Van Vactor, D. L., Jr., Paidhungat, M., and Zipursky, S. L. (1990). Induction of cell fate in the Drosophila retina: the bride of sevenless protein is predicted to contain a large extracellular domain and seven transmembrane segments. Genes Dev 4, 18351847. Hay, B. A., Huh, J. R., and Guo, M. (2004). The genetics of cell death: approaches, insights and opportunities in Drosophila. Nat Rev Genet 5, 911-922. Hay, B. A., Wassarman, D. A., and Rubin, G. M. (1995). Drosophila homologs of baculovirus inhibitor of apoptosis proteins function to block cell death. Cell 83, 1253-1262. Hay, B. A., Wolff, T., and Rubin, G. M. (1994). Expression of baculovirus P35 prevents cell death in Drosophila. Development 120, 2121-2129. Hayashi, T., and Saigo, K. (2001). Diversification of cell types in the Drosophila eye by 99 differential expression of prepattern genes. Mech Dev 108, 13-27. Heberlein, U., and Moses, K. (1995). Mechanisms of Drosophila retinal morphogenesis: the virtues of being progressive. Cell 81, 987-990. Heitzler, P., Bourouis, M., Ruel, L., Carteret, C., and Simpson, P. (1996). Genes of the Enhancer of split and achaete-scute complexes are required for a regulatory loop between Notch and Delta during lateral signalling in Drosophila. Development 122, 161-171. Higashijima, S., Kojima, T., Michiue, T., Ishimaru, S., Emori, Y., and Saigo, K. (1992). Dual Bar homeo box genes of Drosophila required in two photoreceptor cells, R1 and R6, and primary pigment cells for normal eye development. Genes Dev 6, 50-60. Hinkle, C. L., Mohan, M. J., Lin, P., Yeung, N., Rasmussen, F., Milla, M. E., and Moss, M. L. (2003). Multiple metalloproteinases process protransforming growth factor-alpha (proTGFalpha). Biochemistry 42, 2127-2136. Hinkle, C. L., Sunnarborg, S. W., Loiselle, D., Parker, C. E., Stevenson, M., Russell, W. E., and Lee, D. C. (2004). Selective roles for tumor necrosis factor alpha-converting enzyme/ ADAM17 in the shedding of the epidermal growth factor receptor ligand family: the juxtamembrane stalk determines cleavage efficiency. J Biol Chem 279, 24179-24188. Hori, K., Fostier, M., Ito, M., Fuwa, T. J., Go, M. J., Okano, H., Baron, M., and Matsuno, K. (2004). Drosophila deltex mediates suppressor of Hairless-independent and late-endosomal activation of Notch signaling. Development 131, 5527-5537. Howes, R., Wasserman, J. D., and Freeman, M. (1998). In vivo analysis of Argos structurefunction. Sequence requirements for inhibition of the Drosophila epidermal growth factor receptor. J Biol Chem 273, 4275-4281. Hsiao, F. C., Williams, A., Davies, E. L., and Rebay, I. (2001). Eyes absent mediates cross-talk between retinal determination genes and the receptor tyrosine kinase signaling pathway. Dev Cell 1, 51-61. Hsieh, J. J., Henkel, T., Salmon, P., Robey, E., Peterson, M. G., and Hayward, S. D. (1996). Truncated mammalian Notch1 activates CBF1/RBPJk-repressed genes by a mechanism resembling that of Epstein-Barr virus EBNA2. Mol Cell Biol 16, 952-959. Hsieh, J. J., Zhou, S., Chen, L., Young, D. B., and Hayward, S. D. (1999). CIR, a corepressor linking the DNA binding factor CBF1 to the histone deacetylase complex. Proc Natl Acad Sci U S A 96, 23-28. Hsiung, F., and Moses, K. (2002). Retinal development in Drosophila: specifying the first neuron. Hum Mol Genet 11, 1207-1214. Hu, Y., Ye, Y., and Fortini, M. E. (2002). Nicastrin is required for gamma-secretase cleavage of the Drosophila Notch receptor. Dev Cell 2, 69-78. Huang, W., Alessandrini, A., Crews, C. M., and Erikson, R. L. (1993). Raf-1 forms a stable complex with Mek1 and activates Mek1 by serine phosphorylation. Proc Natl Acad Sci U S A 90, 10947-10951. Hunter, T. (2000). Signaling--2000 and beyond. Cell 100, 113-127. Irvine, K. D. (1999). Fringe, Notch, and making developmental boundaries. Curr Opin Genet Dev 9, 434-441. Irvine, K. D., and Rauskolb, C. (2001). Boundaries in development: formation and function. Annu Rev Cell Dev Biol 17, 189-214. Irvine, K. D., and Vogt, T. F. (1997). Dorsal-ventral signaling in limb development. Curr Opin Cell Biol 9, 867-876. 100 Irvine, K. D., and Wieschaus, E. (1994). fringe, a Boundary-specific signaling molecule, mediates interactions between dorsal and ventral cells during Drosophila wing development. Cell 79, 595-606. Islam, R., Wei, S. Y., Chiu, W. H., Hortsch, M., and Hsu, J. C. (2003). Neuroglian activates Echinoid to antagonize the Drosophila EGF receptor signaling pathway. Development 130, 2051-2059. Itoh, M., Kim, C. H., Palardy, G., Oda, T., Jiang, Y. J., Maust, D., Yeo, S. Y., Lorick, K., Wright, G. J., Ariza-McNaughton, L., et al. (2003). Mind bomb is a ubiquitin ligase that is essential for efficient activation of Notch signaling by Delta. Dev Cell 4, 67-82. Iwatsubo, T. (2004). The gamma-secretase complex: machinery for intramembrane proteolysis. Curr Opin Neurobiol 14, 379-383. Jan, Y. N., and Jan, L. Y. (1995). Maggot’s hair and bug’s eye: role of cell interactions and intrinsic factors in cell fate specification. Neuron 14, 1-5. Jarman, A. P., Grau, Y., Jan, L. Y., and Jan, Y. N. (1993). atonal is a proneural gene that directs chordotonal organ formation in the Drosophila peripheral nervous system. Cell 73, 13071321. Jarman, A. P., Grell, E. H., Ackerman, L., Jan, L. Y., and Jan, Y. N. (1994). Atonal is the proneural gene for Drosophila photoreceptors. Nature 369, 398-400. Jarman, A. P., Sun, Y., Jan, L. Y., and Jan, Y. N. (1995). Role of the proneural gene, atonal, in formation of Drosophila chordotonal organs and photoreceptors. Development 121, 20192030. Jarriault, S., Brou, C., Logeat, F., Schroeter, E. H., Kopan, R., and Israel, A. (1995). Signalling downstream of activated mammalian Notch. Nature 377, 355-358. Jarriault, S., and Greenwald, I. (2005). Evidence for functional redundancy between C. elegans ADAM proteins SUP-17/Kuzbanian and ADM-4/TACE. Dev Biol 287, 1-10. Jaumot, M., and Hancock, J. F. (2001). Protein phosphatases 1 and 2A promote Raf-1 activation by regulating 14-3-3 interactions. Oncogene 20, 3949-3958. Jennings, B., Preiss, A., Delidakis, C., and Bray, S. (1994). The Notch signalling pathway is required for Enhancer of split bHLH protein expression during neurogenesis in the Drosophila embryo. Development 120, 3537-3548. Jin, M. H., Sawamoto, K., Ito, M., and Okano, H. (2000). The interaction between the Drosophila secreted protein argos and the epidermal growth factor receptor inhibits dimerization of the receptor and binding of secreted spitz to the receptor. Mol Cell Biol 20, 2098-2107. Jones, F. S., and Meech, R. (1999). Knockout of REST/NRSF shows that the protein is a potent repressor of neuronally expressed genes in non-neural tissues. Bioessays 21, 372-376. Jordan, K. C., Clegg, N. J., Blasi, J. A., Morimoto, A. M., Sen, J., Stein, D., McNeill, H., Deng, W. M., Tworoger, M., and Ruohola-Baker, H. (2000). The homeobox gene mirror links EGF signalling to embryonic dorso-ventral axis formation through notch activation. Nat Genet 24, 429-433. Jordan, K. C., Hatfield, S. D., Tworoger, M., Ward, E. J., Fischer, K. A., Bowers, S., and Ruohola-Baker, H. (2005). Genome wide analysis of transcript levels after perturbation of the EGFR pathway in the Drosophila ovary. Dev Dyn 232, 709-724. Juanes, P. P., Ferreira, L., Montero, J. C., Arribas, J., and Pandiella, A. (2005). N-terminal cleavage of proTGFalpha occurs at the cell surface by a TACE-independent activity. Biochem J 389, 161-172. 101 Kania, A., Salzberg, A., Bhat, M., D’Evelyn, D., He, Y., Kiss, I., and Bellen, H. J. (1995). P-element mutations affecting embryonic peripheral nervous system development in Drosophila melanogaster. Genetics 139, 1663-1678. Kao, H. Y., Ordentlich, P., Koyano-Nakagawa, N., Tang, Z., Downes, M., Kintner, C. R., Evans, R. M., and Kadesch, T. (1998). A histone deacetylase corepressor complex regulates the Notch signal transduction pathway. Genes Dev 12, 2269-2277. Kato, H., Taniguchi, Y., Kurooka, H., Minoguchi, S., Sakai, T., Nomura-Okazaki, S., Tamura, K., and Honjo, T. (1997). Involvement of RBP-J in biological functions of mouse Notch1 and its derivatives. Development 124, 4133-4141. Kauffmann, R. C., Li, S., Gallagher, P. A., Zhang, J., and Carthew, R. W. (1996). Ras1 signaling and transcriptional competence in the R7 cell of Drosophila. Genes Dev 10, 2167-2178. Kenyon, K. L., Ranade, S. S., Curtiss, J., Mlodzik, M., and Pignoni, F. (2003). Coordinating proliferation and tissue specification to promote regional identity in the Drosophila head. Dev Cell 5, 403-414. Khokhlatchev, A. V., Canagarajah, B., Wilsbacher, J., Robinson, M., Atkinson, M., Goldsmith, E., and Cobb, M. H. (1998). Phosphorylation of the MAP kinase ERK2 promotes its homodimerization and nuclear translocation. Cell 93, 605-615. Kidd, S., Kelley, M. R., and Young, M. W. (1986). Sequence of the notch locus of Drosophila melanogaster: relationship of the encoded protein to mammalian clotting and growth factors. Mol Cell Biol 6, 3094-3108. Kidd, S., and Lieber, T. (2002). Furin cleavage is not a requirement for Drosophila Notch function. Mech Dev 115, 41-51. Kidd, S., Lieber, T., and Young, M. W. (1998). Ligand-induced cleavage and regulation of nuclear entry of Notch in Drosophila melanogaster embryos. Genes Dev 12, 3728-3740. Killar, L., White, J., Black, R., and Peschon, J. (1999). Adamalysins. A family of metzincins including TNF-alpha converting enzyme (TACE). Ann N Y Acad Sci 878, 442-452. Kimble, J., Henderson, S., and Crittenden, S. (1998). Notch/LIN-12 signaling: transduction by regulated protein slicing. Trends Biochem Sci 23, 353-357. Klambt, C. (2000). EGF receptor signalling: the importance of presentation. Curr Biol 10, R388391. Klambt, C. (2002). EGF receptor signalling: roles of star and rhomboid revealed. Curr Biol 12, R21-23. Klambt, C., Knust, E., Tietze, K., and Campos-Ortega, J. A. (1989). Closely related transcripts encoded by the neurogenic gene complex enhancer of split of Drosophila melanogaster. Embo J 8, 203-210. Klapper, L. N., Kirschbaum, M. H., Sela, M., and Yarden, Y. (2000). Biochemical and clinical implications of the ErbB/HER signaling network of growth factor receptors. Adv Cancer Res 77, 25-79. Klein, D. E., Nappi, V. M., Reeves, G. T., Shvartsman, S. Y., and Lemmon, M. A. (2004). Argos inhibits epidermal growth factor receptor signalling by ligand sequestration. Nature 430, 1040-1044. Klein, T. (2002). kuzbanian is required cell autonomously during Notch signalling in the Drosophila wing. Dev Genes Evol 212, 251-255. Klein, T., Seugnet, L., Haenlin, M., and Martinez Arias, A. (2000). Two different activities of Suppressor of Hairless during wing development in Drosophila. Development 127, 3553102 3566. Knust, E., Bremer, K. A., Vassin, H., Ziemer, A., Tepass, U., and Campos-Ortega, J. A. (1987a). The enhancer of split locus and neurogenesis in Drosophila melanogaster. Dev Biol 122, 262-273. Knust, E., Schrons, H., Grawe, F., and Campos-Ortega, J. A. (1992). Seven genes of the Enhancer of split complex of Drosophila melanogaster encode helix-loop-helix proteins. Genetics 132, 505-518. Knust, E., Tietze, K., and Campos-Ortega, J. A. (1987b). Molecular analysis of the neurogenic locus Enhancer of split of Drosophila melanogaster. Embo J 6, 4113-4123. Kojima, T. (2004). The mechanism of Drosophila leg development along the proximodistal axis. Dev Growth Differ 46, 115-129. Kolch, W. (2000). Meaningful relationships: the regulation of the Ras/Raf/MEK/ERK pathway by protein interactions. Biochem J 351 Pt 2, 289-305. Kolch, W. (2005). Coordinating ERK/MAPK signalling through scaffolds and inhibitors. Nat Rev Mol Cell Biol 6, 827-837. Kolodkin, A. L., Pickup, A. T., Lin, D. M., Goodman, C. S., and Banerjee, U. (1994). Characterization of Star and its interactions with sevenless and EGF receptor during photoreceptor cell development in Drosophila. Development 120, 1731-1745. Kondoh, K., Torii, S., and Nishida, E. (2005). Control of MAP kinase signaling to the nucleus. Chromosoma 114, 86-91. Kopan, R. (2002). Notch: a membrane-bound transcription factor. J Cell Sci 115, 1095-1097. Kopczynski, C. C., Alton, A. K., Fechtel, K., Kooh, P. J., and Muskavitch, M. A. (1988). Delta, a Drosophila neurogenic gene, is transcriptionally complex and encodes a protein related to blood coagulation factors and epidermal growth factor of vertebrates. Genes Dev 2, 17231735. Kramer, H., Cagan, R. L., and Zipursky, S. L. (1991). Interaction of bride of sevenless membrane-bound ligand and the sevenless tyrosine-kinase receptor. Nature 352, 207-212. Kramer, S., West, S. R., and Hiromi, Y. (1995). Cell fate control in the Drosophila retina by the orphan receptor seven- up: its role in the decisions mediated by the ras signaling pathway. Development 121, 1361-1372. Kumar, J., and Moses, K. (1997). Transcription factors in eye development: a gorgeous mosaic? Genes Dev 11, 2023-2028. Kumar, J. P., Hsiung, F., Powers, M. A., and Moses, K. (2003). Nuclear translocation of activated MAP kinase is developmentally regulated in the developing Drosophila eye. Development 130, 3703-3714. Kumar, J. P., and Moses, K. (2001a). EGF receptor and Notch signaling act upstream of Eyeless/ Pax6 to control eye specification. Cell 104, 687-697. Kumar, J. P., and Moses, K. (2001b). The EGF receptor and notch signaling pathways control the initiation of the morphogenetic furrow during Drosophila eye development. Development 128, 2689-2697. Kumar, J. P., Tio, M., Hsiung, F., Akopyan, S., Gabay, L., Seger, R., Shilo, B. Z., and Moses, K. (1998). Dissecting the roles of the Drosophila EGF receptor in eye development and MAP kinase activation. Development 125, 3875-3885. Kumar, V., Bustin, S. A., and McKay, I. A. (1995). Transforming growth factor alpha. Cell Biol Int 19, 373-388. 103 Kurada, P., and White, K. (1998). Ras promotes cell survival in Drosophila by downregulating hid expression. Cell 95, 319-329. Kurata, S., Go, M. J., Artavanis-Tsakonas, S., and Gehring, W. J. (2000). Notch signaling and the determination of appendage identity. Proc Natl Acad Sci U S A 97, 2117-2122. Kuroda, K., Han, H., Tani, S., Tanigaki, K., Tun, T., Furukawa, T., Taniguchi, Y., Kurooka, H., Hamada, Y., Toyokuni, S., and Honjo, T. (2003). Regulation of Marginal Zone B Cell Development by MINT, a Suppressor of Notch/RBP-J Signaling Pathway. Immunity 18, 301312. Kurooka, H., and Honjo, T. (2000). Functional interaction between the mouse notch1 intracellular region and histone acetyltransferases PCAF and GCN5. J Biol Chem 275, 17211-17220. Kurooka, H., Kuroda, K., and Honjo, T. (1998). Roles of the ankyrin repeats and C-terminal region of the mouse notch1 intracellular region. Nucleic Acids Res 26, 5448-5455. Lai, E. C. (2002a). Developmental signaling: shrimp and strawberries help flies make cones. Curr Biol 12, R722-724. Lai, E. C. (2002b). Keeping a good pathway down: transcriptional repression of Notch pathway target genes by CSL proteins. EMBO Rep 3, 840-845. Lai, E. C. (2004). Notch signaling: control of cell communication and cell fate. Development 131, 965-973. Lai, E. C., Bodner, R., and Posakony, J. W. (2000). The enhancer of split complex of Drosophila includes four Notch-regulated members of the bearded gene family. Development 127, 34413455. Lai, E. C., Deblandre, G. A., Kintner, C., and Rubin, G. M. (2001). Drosophila neuralized is a ubiquitin ligase that promotes the internalization and degradation of delta. Dev Cell 1, 783794. Lai, E. C., Roegiers, F., Qin, X., Jan, Y. N., and Rubin, G. M. (2005). The ubiquitin ligase Drosophila Mind bomb promotes Notch signaling by regulating the localization and activity of Serrate and Delta. Development 132, 2319-2332. Lai, E. C., and Rubin, G. M. (2001). neuralized functions cell-autonomously to regulate a subset of notch-dependent processes during adult Drosophila development. Dev Biol 231, 217-233. Lai, Z. C., Fetchko, M., and Li, Y. (1997). Repression of Drosophila photoreceptor cell fate through cooperative action of two transcriptional repressors Yan and Tramtrack. Genetics 147, 1131-1137. Lai, Z. C., Harrison, S. D., Karim, F., Li, Y., and Rubin, G. M. (1996). Loss of tramtrack gene activity results in ectopic R7 cell formation, even in a sina mutant background. Proc Natl Acad Sci U S A 93, 5025-5030. Laudon, H., Hansson, E. M., Melen, K., Bergman, A., Farmery, M. R., Winblad, B., Lendahl, U., von Heijne, G., and Naslund, J. (2005). A nine-transmembrane domain topology for presenilin 1. J Biol Chem 280, 35352-35360. LaVoie, M. J., and Selkoe, D. J. (2003). The Notch ligands, Jagged and Delta, are sequentially processed by alpha-secretase and presenilin/gamma-secretase and release signaling fragments. J Biol Chem 278, 34427-34437. Le Borgne, R. (2006). Regulation of Notch signalling by endocytosis and endosomal sorting. Curr Opin Cell Biol. Le Borgne, R., Bardin, A., and Schweisguth, F. (2005). The roles of receptor and ligand 104 endocytosis in regulating Notch signaling. Development 132, 1751-1762. Le Borgne, R., and Schweisguth, F. (2003a). Notch signaling: endocytosis makes delta signal better. Curr Biol 13, R273-275. Le Borgne, R., and Schweisguth, F. (2003b). Unequal segregation of Neuralized biases Notch activation during asymmetric cell division. Dev Cell 5, 139-148. Lecourtois, M., and Schweisguth, F. (1995). The neurogenic suppressor of hairless DNAbinding protein mediates the transcriptional activation of the enhancer of split complex genes triggered by Notch signaling. Genes Dev 9, 2598-2608. Lee, D. C., Sunnarborg, S. W., Hinkle, C. L., Myers, T. J., Stevenson, M. Y., Russell, W. E., Castner, B. J., Gerhart, M. J., Paxton, R. J., Black, R. A., et al. (2003). TACE/ADAM17 processing of EGFR ligands indicates a role as a physiological convertase. Ann N Y Acad Sci 995, 22-38. Lee, E. C., Hu, X., Yu, S. Y., and Baker, N. E. (1996). The scabrous gene encodes a secreted glycoprotein dimer and regulates proneural development in Drosophila eyes. Mol Cell Biol 16, 1179-1188. Lee, J. R., Urban, S., Garvey, C. F., and Freeman, M. (2001). Regulated intracellular ligand transport and proteolysis control EGF signal activation in Drosophila. Cell 107, 161-171. Lei, L., Xu, A., Panin, V. M., and Irvine, K. D. (2003). An O-fucose site in the ligand binding domain inhibits Notch activation. Development 130, 6411-6421. Lemmon, M. A., Bu, Z., Ladbury, J. E., Zhou, M., Pinchasi, D., Lax, I., Engelman, D. M., and Schlessinger, J. (1997). Two EGF molecules contribute additively to stabilization of the EGFR dimer. Embo J 16, 281-294. Leong, G. M., Subramaniam, N., Issa, L. L., Barry, J. B., Kino, T., Driggers, P. H., Hayman, M. J., Eisman, J. A., and Gardiner, E. M. (2004). Ski-interacting protein, a bifunctional nuclear receptor coregulator that interacts with N-CoR/SMRT and p300. Biochem Biophys Res Commun 315, 1070-1076. Lesokhin, A. M., Yu, S. Y., Katz, J., and Baker, N. E. (1999). Several levels of EGF receptor signaling during photoreceptor specification in wild-type, Ellipse, and null mutant Drosophila. Dev Biol 205, 129-144. Li, S., Li, Y., Carthew, R. W., and Lai, Z. C. (1997). Photoreceptor cell differentiation requires regulated proteolysis of the transcriptional repressor Tramtrack. Cell 90, 469-478. Li, Y., and Baker, N. E. (2001). Proneural enhancement by Notch overcomes Suppressor-ofHairless repressor function in the developing Drosophila eye. Curr Biol 11, 330-338. Li, Y., Fetchko, M., Lai, Z. C., and Baker, N. E. (2003a). Scabrous and Gp150 are endosomal proteins that regulate Notch activity. Development 130, 2819-2827. Li, Y., Lei, L., Irvine, K. D., and Baker, N. E. (2003b). Notch activity in neural cells triggered by a mutant allele with altered glycosylation. Development 130, 2829-2840. Lieber, T., Kidd, S., Alcamo, E., Corbin, V., and Young, M. W. (1993). Antineurogenic phenotypes induced by truncated Notch proteins indicate a role in signal transduction and may point to a novel function for Notch in nuclei. Genes Dev 7, 1949-1965. Lieber, T., Kidd, S., and Young, M. W. (2002). kuzbanian-mediated cleavage of Drosophila Notch. Genes Dev 16, 209-221. Livneh, E., Glazer, L., Segal, D., Schlessinger, J., and Shilo, B. Z. (1985). The Drosophila EGF receptor gene homolog: conservation of both hormone binding and kinase domains. Cell 40, 599-607. 105 Logeat, F., Bessia, C., Brou, C., LeBail, O., Jarriault, S., Seidah, N. G., and Israel, A. (1998). The Notch1 receptor is cleaved constitutively by a furin-like convertase. Proc Natl Acad Sci U S A 95, 8108-8112. Lopez-Schier, H., and St Johnston, D. (2002). Drosophila nicastrin is essential for the intramembranous cleavage of notch. Dev Cell 2, 79-89. Lorenzen, J. A., Baker, S. E., Denhez, F., Melnick, M. B., Brower, D. L., and Perkins, L. A. (2001). Nuclear import of activated D-ERK by DIM-7, an importin family member encoded by the gene moleskin. Development 128, 1403-1414. Louvi, A., and Artavanis-Tsakonas, S. (2006). Notch signalling in vertebrate neural development. Nat Rev Neurosci 7, 93-102. Lubman, O. Y., Korolev, S. V., and Kopan, R. (2004). Anchoring notch genetics and biochemistry; structural analysis of the ankyrin domain sheds light on existing data. Mol Cell 13, 619-626. Lum, L., Yao, S., Mozer, B., Rovescalli, A., Von Kessler, D., Nirenberg, M., and Beachy, P. A. (2003). Identification of Hedgehog pathway components by RNAi in Drosophila cultured cells. Science 299, 2039-2045. Ma, C., Zhou, Y., Beachy, P. A., and Moses, K. (1993). The segment polarity gene hedgehog is required for progression of the morphogenetic furrow in the developing Drosophila eye. Cell 75, 927-938. Macdonald, S. G., Crews, C. M., Wu, L., Driller, J., Clark, R., Erikson, R. L., and McCormick, F. (1993). Reconstitution of the Raf-1-MEK-ERK signal transduction pathway in vitro. Mol Cell Biol 13, 6615-6620. Magee, T., and Marshall, C. (1999). New insights into the interaction of Ras with the plasma membrane. Cell 98, 9-12. Marenda, D. R., Vrailas, A. D., Rodrigues, A. B., Cook, S., Powers, M. A., Lorenzen, J. A., Perkins, L. A., and Moses, K. (2006). MAP kinase subcellular localization controls both pattern and proliferation in the developing Drosophila wing. Development 133, 43-51. Margarit, S. M., Sondermann, H., Hall, B. E., Nagar, B., Hoelz, A., Pirruccello, M., Bar-Sagi, D., and Kuriyan, J. (2003). Structural evidence for feedback activation by Ras.GTP of the Rasspecific nucleotide exchange factor SOS. Cell 112, 685-695. Mazaleyrat, S. L., Fostier, M., Wilkin, M. B., Aslam, H., Evans, D. A., Cornell, M., and Baron, M. (2003). Down-regulation of Notch target gene expression by Suppressor of deltex. Dev Biol 255, 363-372. McCormick, F. (1994). Activators and effectors of ras p21 proteins. Curr Opin Genet Dev 4, 7176. McGill, M. A., and McGlade, C. J. (2003). Mammalian numb proteins promote Notch1 receptor ubiquitination and degradation of the Notch1 intracellular domain. J Biol Chem 278, 2319623203. McNeill, H. (2002). Planar polarity: location, location, location. Curr Biol 12, R449-451. Merlos-Suarez, A., Ruiz-Paz, S., Baselga, J., and Arribas, J. (2001). Metalloprotease-dependent protransforming growth factor-alpha ectodomain shedding in the absence of tumor necrosis factor-alpha-converting enzyme. J Biol Chem 276, 48510-48517. Miller, D. T., and Cagan, R. L. (1998). Local induction of patterning and programmed cell death in the developing Drosophila retina. Development 125, 2327-2335. Mishra-Gorur, K., Rand, M. D., Perez-Villamil, B., and Artavanis-Tsakonas, S. (2002). Down106 regulation of Delta by proteolytic processing. J Cell Biol 159, 313-324. Mlodzik, M. (1999). Planar polarity in the Drosophila eye: a multifaceted view of signaling specificity and cross-talk. Embo J 18, 6873-6879. Mlodzik, M., Baker, N. E., and Rubin, G. M. (1990a). Isolation and expression of scabrous, a gene regulating neurogenesis in Drosophila. Genes Dev 4, 1848-1861. Mlodzik, M., Hiromi, Y., Weber, U., Goodman, C. S., and Rubin, G. M. (1990b). The Drosophila seven-up gene, a member of the steroid receptor gene superfamily, controls photoreceptor cell fates. Cell 60, 211-224. Moehlmann, T., Winkler, E., Xia, X., Edbauer, D., Murrell, J., Capell, A., Kaether, C., Zheng, H., Ghetti, B., Haass, C., and Steiner, H. (2002). Presenilin-1 mutations of leucine 166 equally affect the generation of the Notch and APP intracellular domains independent of their effect on Abeta 42 production. Proc Natl Acad Sci U S A 99, 8025-8030. Moghal, N., and Sternberg, P. W. (1999). Multiple positive and negative regulators of signaling by the EGF- receptor. Curr Opin Cell Biol 11, 190-196. Mohammadi, M., Honegger, A., Sorokin, A., Ullrich, A., Schlessinger, J., and Hurwitz, D. R. (1993). Aggregation-induced activation of the epidermal growth factor receptor protein tyrosine kinase. Biochemistry 32, 8742-8748. Mollereau, B., Dominguez, M., Webel, R., Colley, N. J., Keung, B., de Celis, J. F., and Desplan, C. (2001). Two-step process for photoreceptor formation in Drosophila. Nature 412, 911-913. Moloney, D. J., Shair, L. H., Lu, F. M., Xia, J., Locke, R., Matta, K. L., and Haltiwanger, R. S. (2000). Mammalian Notch1 is modified with two unusual forms of O-linked glycosylation found on epidermal growth factor-like modules. J Biol Chem 275, 9604-9611. Morata, G., and Ripoll, P. (1975). Minutes: mutants of drosophila autonomously affecting cell division rate. Dev Biol 42, 211-221. Morel, V., Lecourtois, M., Massiani, O., Maier, D., Preiss, A., and Schweisguth, F. (2001). Transcriptional repression by suppressor of hairless involves the binding of a hairless-dCtBP complex in Drosophila. Curr Biol 11, 789-792. Morel, V., and Schweisguth, F. (2000). Repression by suppressor of hairless and activation by Notch are required to define a single row of single-minded expressing cells in the Drosophila embryo. Genes Dev 14, 377-388. Morrison, D. K., and Davis, R. J. (2003). Regulation of MAP kinase signaling modules by scaffold proteins in mammals. Annu Rev Cell Dev Biol 19, 91-118. Moscoso del Prado, J., and Garcia-Bellido, A. (1984). Cell interactions in the generation of chaete pattern in Drosophila. Roux’s Archives Dev Biol 193, 246-251. Mumm, J. S., and Kopan, R. (2000). Notch signaling: from the outside in. Dev Biol 228, 151165. Mumm, J. S., Schroeter, E. H., Saxena, M. T., Griesemer, A., Tian, X., Pan, D. J., Ray, W. J., and Kopan, R. (2000). A ligand-induced extracellular cleavage regulates gamma-secretase-like proteolytic activation of Notch1. Mol Cell 5, 197-206. Nagaraj, R., and Banerjee, U. (2004). The little R cell that could. Int J Dev Biol 48, 755-760. Nagel, A. C., Krejci, A., Tenin, G., Bravo-Patino, A., Bray, S., Maier, D., and Preiss, A. (2005). Hairless-mediated repression of notch target genes requires the combined activity of Groucho and CtBP corepressors. Mol Cell Biol 25, 10433-10441. Nagel, A. C., and Preiss, A. (1999). Notchspl is deficient for inductive processes in the eye, and 107 E(spl)D enhances split by interfering with proneural activity. Dev Biol 208, 406-415. Nakagawa, T., Guichard, A., Castro, C. P., Xiao, Y., Rizen, M., Zhang, H. Z., Hu, D., Bang, A., Helms, J., Bier, E., and Derynck, R. (2005). Characterization of a human rhomboid homolog, p100hRho/RHBDF1, which interacts with TGF-alpha family ligands. Dev Dyn 233, 13151331. Nakao, K., and Campos-Ortega, J. A. (1996). Persistent expression of genes of the enhancer of split complex suppresses neural development in Drosophila. Neuron 16, 275-286. Nam, Y., Sliz, P., Song, L., Aster, J. C., and Blacklow, S. C. (2006). Structural Basis for Cooperativity in Recruitment of MAML Coactivators to Notch Transcription Complexes. Cell 124, 973-983. Neuman-Silberberg, F. S., Schejter, E., Hoffmann, F. M., and Shilo, B. Z. (1984). The Drosophila ras oncogenes: structure and nucleotide sequence. Cell 37, 1027-1033. Neuman-Silberberg, F. S., and Schupbach, T. (1993). The Drosophila dorsoventral patterning gene gurken produces a dorsally localized RNA and encodes a TGF alpha-like protein. Cell 75, 165-174. Nybakken, K., Vokes, S. A., Lin, T. Y., McMahon, A. P., and Perrimon, N. (2005). A genomewide RNA interference screen in Drosophila melanogaster cells for new components of the Hh signaling pathway. Nat Genet 37, 1323-1332. Oellers, N., Dehio, M., and Knust, E. (1994). bHLH proteins encoded by the Enhancer of split complex of Drosophila negatively interfere with transcriptional activation mediated by proneural genes. Mol Gen Genet 244, 465-473. Okajima, T., and Irvine, K. D. (2002). Regulation of notch signaling by o-linked fucose. Cell 111, 893-904. Okajima, T., Xu, A., and Irvine, K. D. (2003). Modulation of notch-ligand binding by protein Ofucosyltransferase 1 and fringe. J Biol Chem 278, 42340-42345. Okochi, M., Fukumori, A., Jiang, J., Itoh, N., Kimura, R., Steiner, H., Haass, C., Tagami, S., and Takeda, M. (2006). Secretion of the Notch-1 Abeta-like peptide during Notch signaling. J Biol Chem 281, 7890-7898. Okochi, M., Steiner, H., Fukumori, A., Tanii, H., Tomita, T., Tanaka, T., Iwatsubo, T., Kudo, T., Takeda, M., and Haass, C. (2002). Presenilins mediate a dual intramembranous gammasecretase cleavage of Notch-1. Embo J 21, 5408-5416. Olayioye, M. A., Neve, R. M., Lane, H. A., and Hynes, N. E. (2000). The ErbB signaling network: receptor heterodimerization in development and cancer. Embo J 19, 3159-3167. Olivier, J. P., Raabe, T., Henkemeyer, M., Dickson, B., Mbamalu, G., Margolis, B., Schlessinger, J., Hafen, E., and Pawson, T. (1993). A Drosophila SH2-SH3 adaptor protein implicated in coupling the sevenless tyrosine kinase to an activator of Ras guanine nucleotide exchange, Sos. Cell 73, 179-191. Olson, T. S., and Lane, M. D. (1989). A common mechanism for posttranslational activation of plasma membrane receptors? Faseb J 3, 1618-1624. O’Neill, E. M., Rebay, I., Tjian, R., and Rubin, G. M. (1994). The activities of two Ets-related transcription factors required for Drosophila eye development are modulated by the Ras/ MAPK pathway. Cell 78, 137-147. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W., et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J 21, 5417-5426. 108 Overstreet, E., Fitch, E., and Fischer, J. A. (2004). Fat facets and Liquid facets promote Delta endocytosis and Delta signaling in the signaling cells. Development 131, 5355-5366. Pan, D., and Rubin, G. M. (1997). Kuzbanian controls proteolytic processing of Notch and mediates lateral inhibition during Drosophila and vertebrate neurogenesis. Cell 90, 271-280. Panin, V. M., Papayannopoulos, V., Wilson, R., and Irvine, K. D. (1997). Fringe modulates Notch-ligand interactions. Nature 387, 908-912. Panin, V. M., Shao, L., Lei, L., Moloney, D. J., Irvine, K. D., and Haltiwanger, R. S. (2002). Notch ligands are substrates for protein O-fucosyltransferase-1 and Fringe. J Biol Chem 277, 29945-29952. Papayannopoulos, V., Tomlinson, A., Panin, V. M., Rauskolb, C., and Irvine, K. D. (1998). Dorsal-ventral signaling in the Drosophila eye. Science 281, 2031-2034. Pappu, K. S., and Mardon, G. (2004). Genetic control of retinal specification and determination in Drosophila. Int J Dev Biol 48, 913-924. Parks, A. L., Klueg, K. M., Stout, J. R., and Muskavitch, M. A. (2000). Ligand endocytosis drives receptor dissociation and activation in the Notch pathway. Development 127, 13731385. Paroush, Z., Finley, R. L., Jr., Kidd, T., Wainwright, S. M., Ingham, P. W., Brent, R., and IshHorowicz, D. (1994). Groucho is required for Drosophila neurogenesis, segmentation, and sex determination and interacts directly with hairy-related bHLH proteins. Cell 79, 805-815. Pasternak, S. H., Callahan, J. W., and Mahuran, D. J. (2004). The role of the endosomal/ lysosomal system in amyloid-beta production and the pathophysiology of Alzheimer’s disease: reexamining the spatial paradox from a lysosomal perspective. J Alzheimers Dis 6, 53-65. Pawson, T., Gish, G. D., and Nash, P. (2001). SH2 domains, interaction modules and cellular wiring. Trends Cell Biol 11, 504-511. Pawson, T., and Scott, J. D. (1997). Signaling through scaffold, anchoring, and adaptor proteins. Science 278, 2075-2080. Pearson, G., Robinson, F., Beers Gibson, T., Xu, B. E., Karandikar, M., Berman, K., and Cobb, M. H. (2001). Mitogen-activated protein (MAP) kinase pathways: regulation and physiological functions. Endocr Rev 22, 153-183. Periz, G., and Fortini, M. E. (2004). Functional reconstitution of gamma-secretase through coordinated expression of presenilin, nicastrin, Aph-1, and Pen-2. J Neurosci Res 77, 309322. Perrimon, N., and Perkins, L. A. (1997). There must be 50 ways to rule the signal: the case of the Drosophila EGF receptor. Cell 89, 13-16. Peschon, J. J., Slack, J. L., Reddy, P., Stocking, K. L., Sunnarborg, S. W., Lee, D. C., Russell, W. E., Castner, B. J., Johnson, R. S., Fitzner, J. N., et al. (1998). An essential role for ectodomain shedding in mammalian development. Science 282, 1281-1284. Petcherski, A. G., and Kimble, J. (2000a). LAG-3 is a putative transcriptional activator in the C. elegans Notch pathway. Nature 405, 364-368. Petcherski, A. G., and Kimble, J. (2000b). Mastermind is a putative activator for Notch. Curr Biol 10, R471-473. Pitsouli, C., and Delidakis, C. (2005). The interplay between DSL proteins and ubiquitin ligases in Notch signaling. Development 132, 4041-4050. Portin, P. (2002). General outlines of the molecular genetics of the Notch signalling pathway in 109 Drosophila melanogaster: a review. Hereditas 136, 89-96. Portin, P., and Rantanen, M. (1991). Interaction of master mind, big brain, neuralized and Notch genes of Drosophila melanogaster as expressed in adult morphology. Hereditas 114, 197200. Powell, P. A., Wesley, C., Spencer, S., and Cagan, R. L. (2001). Scabrous complexes with Notch to mediate boundary formation. Nature 409, 626-630. Preiss, A., Hartley, D. A., and Artavanis-Tsakonas, S. (1988). The molecular genetics of Enhancer of split, a gene required for embryonic neural development in Drosophila. Embo J 7, 3917-3927. Price, J. V., Savenye, E. D., Lum, D., and Breitkreutz, A. (1997). Dominant enhancers of Egfr in Drosophila melanogaster: genetic links between the Notch and Egfr signaling pathways. Genetics 147, 1139-1153. Qi, H., Rand, M. D., Wu, X., Sestan, N., Wang, W., Rakic, P., Xu, T., and Artavanis-Tsakonas, S. (1999). Processing of the notch ligand delta by the metalloprotease Kuzbanian. Science 283, 91-94. Qiao, F., Harada, B., Song, H., Whitelegge, J., Courey, A. J., and Bowie, J. U. (2006). Mae inhibits Pointed-P2 transcriptional activity by blocking its MAPK docking site. Embo J 25, 70-79. Qiao, F., Song, H., Kim, C. A., Sawaya, M. R., Hunter, J. B., Gingery, M., Rebay, I., Courey, A. J., and Bowie, J. U. (2004). Derepression by depolymerization; structural insights into the regulation of Yan by Mae. Cell 118, 163-173. Qin, H., Wang, J., Liang, Y., Taniguchi, Y., Tanigaki, K., and Han, H. (2004). RING1 inhibits transactivation of RBP-J by Notch through interaction with LIM protein KyoT2. Nucleic Acids Res 32, 1492-1501. Qiu, L., Joazeiro, C., Fang, N., Wang, H. Y., Elly, C., Altman, Y., Fang, D., Hunter, T., and Liu, Y. C. (2000). Recognition and ubiquitination of Notch by Itch, a hect-type E3 ubiquitin ligase. J Biol Chem 275, 35734-35737. Raabe, T., and Rapp, U. R. (2003). Ras signaling: PP2A puts Ksr and Raf in the right place. Curr Biol 13, R635-637. Radtke, F., and Raj, K. (2003). The role of Notch in tumorigenesis: oncogene or tumour suppressor? Nat Rev Cancer 3, 756-767. Radtke, F., Schweisguth, F., and Pear, W. (2005). The Notch ‘gospel’. EMBO Rep 6, 1120-1125. Rampal, R., Arboleda-Velasquez, J. F., Nita-Lazar, A., Kosik, K. S., and Haltiwanger, R. S. (2005). Highly conserved O-fucose sites have distinct effects on Notch1 function. J Biol Chem 280, 32133-32140. Rand, M. D., Grimm, L. M., Artavanis-Tsakonas, S., Patriub, V., Blacklow, S. C., Sklar, J., and Aster, J. C. (2000). Calcium depletion dissociates and activates heterodimeric notch receptors. Mol Cell Biol 20, 1825-1835. Rauskolb, C., Correia, T., and Irvine, K. D. (1999). Fringe-dependent separation of dorsal and ventral cells in the Drosophila wing. Nature 401, 476-480. Rawlins, E. L., Lovegrove, B., and Jarman, A. P. (2003a). Echinoid facilitates Notch pathway signalling during Drosophila neurogenesis through functional interaction with Delta. Development 130, 6475-6484. Rawlins, E. L., White, N. M., and Jarman, A. P. (2003b). Echinoid limits R8 photoreceptor specification by inhibiting inappropriate EGF receptor signalling within R8 equivalence 110 groups. Development 130, 3715-3724. Ray, W. J., Yao, M., Mumm, J., Schroeter, E. H., Saftig, P., Wolfe, M., Selkoe, D. J., Kopan, R., and Goate, A. M. (1999a). Cell surface presenilin-1 participates in the gamma-secretase-like proteolysis of Notch. J Biol Chem 274, 36801-36807. Ray, W. J., Yao, M., Nowotny, P., Mumm, J., Zhang, W., Wu, J. Y., Kopan, R., and Goate, A. M. (1999b). Evidence for a physical interaction between presenilin and Notch. Proc Natl Acad Sci U S A 96, 3263-3268. Ready, D. F., Hanson, T. E., and Benzer, S. (1976). Development of the Drosophila retina, a neurocrystalline lattice. Dev Biol 53, 217-240. Rebay, I., Fehon, R. G., and Artavanis-Tsakonas, S. (1993). Specific truncations of Drosophila Notch define dominant activated and dominant negative forms of the receptor. Cell 74, 319329. Rebay, I., Fleming, R. J., Fehon, R. G., Cherbas, L., Cherbas, P., and Artavanis-Tsakonas, S. (1991). Specific EGF repeats of Notch mediate interactions with Delta and Serrate: implications for Notch as a multifunctional receptor. Cell 67, 687-699. Rebay, I., and Rubin, G. M. (1995). Yan functions as a general inhibitor of differentiation and is negatively regulated by activation of the Ras1/MAPK pathway. Cell 81, 857-866. Reeves, G. T., Kalifa, R., Klein, D. E., Lemmon, M. A., and Shvartsman, S. Y. (2005). Computational analysis of EGFR inhibition by Argos. Dev Biol 284, 523-535. Reich, A., and Shilo, B. Z. (2002). Keren, a new ligand of the Drosophila epidermal growth factor receptor, undergoes two modes of cleavage. Embo J 21, 4287-4296. Reinke, R., and Zipursky, S. L. (1988). Cell-cell interaction in the Drosophila retina: the bride of sevenless gene is required in photoreceptor cell R8 for R7 cell development. Cell 55, 321330. Ren, R., Mayer, B. J., Cicchetti, P., and Baltimore, D. (1993). Identification of a ten-amino acid proline-rich SH3 binding site. Science 259, 1157-1161. Reynolds-Kenneally, J., and Mlodzik, M. (2005). Notch signaling controls proliferation through cell-autonomous and non-autonomous mechanisms in the Drosophila eye. Dev Biol 285, 3848. Rhyu, M. S., Jan, L. Y., and Jan, Y. N. (1994). Asymmetric distribution of numb protein during division of the sensory organ precursor cell confers distinct fates to daughter cells. Cell 76, 477-491. Riese, D. J., 2nd, and Stern, D. F. (1998). Specificity within the EGF family/ErbB receptor family signaling network. Bioessays 20, 41-48. Rodrigues, A. B., Werner, E., and Moses, K. (2005). Genetic and biochemical analysis of the role of Egfr in the morphogenetic furrow of the developing Drosophila eye. Development 132, 4697-4707. Roegiers, F., and Jan, Y. N. (2004). Asymmetric cell division. Curr Opin Cell Biol 16, 195-205. Roehl, H., and Kimble, J. (1993). Control of cell fate in C. elegans by a GLP-1 peptide consisting primarily of ankyrin repeats. Nature 364, 632-635. Rogers, E. M., Brennan, C. A., Mortimer, N. T., Cook, S., Morris, A. R., and Moses, K. (2005). Pointed regulates an eye-specific transcriptional enhancer in the Drosophila hedgehog gene, which is required for the movement of the morphogenetic furrow. Development 132, 48334843. Rohrbaugh, M., Ramos, E., Nguyen, D., Price, M., Wen, Y., and Lai, Z. C. (2002). Notch 111 activation of yan expression is antagonized by RTK/pointed signaling in the Drosophila eye. Curr Biol 12, 576-581. Roussel, M., Saule, S., Lagrou, C., Rommens, C., Beug, H., Graf, T., and Stehelin, D. (1979). Three new types of viral oncogene of cellular origin specific for haematopoietic cell transformation. Nature 281, 452-455. Rusconi, J. C., Fink, J. L., and Cagan, R. (2004). klumpfuss regulates cell death in the Drosophila retina. Mech Dev 121, 537-546. Rusconi, J. C., Hays, R., and Cagan, R. L. (2000). Programmed cell death and patterning in Drosophila. Cell Death Differ 7, 1063-1070. Rutledge, B. J., Zhang, K., Bier, E., Jan, Y. N., and Perrimon, N. (1992). The Drosophila spitz gene encodes a putative EGF-like growth factor involved in dorsal-ventral axis formation and neurogenesis. Genes Dev 6, 1503-1517. Saint, R., Kalionis, B., Lockett, T. J., and Elizur, A. (1988). Pattern formation in the developing eye of Drosophila melanogaster is regulated by the homoeo-box gene, rough. Nature 334, 151-154. Sakamoto, K., Ohara, O., Takagi, M., Takeda, S., and Katsube, K. (2002). Intracellular cellautonomous association of Notch and its ligands: a novel mechanism of Notch signal modification. Dev Biol 241, 313-326. Sakata, T., Sakaguchi, H., Tsuda, L., Higashitani, A., Aigaki, T., Matsuno, K., and Hayashi, S. (2004). Drosophila Nedd4 regulates endocytosis of notch and suppresses its ligandindependent activation. Curr Biol 14, 2228-2236. Sanchez-Irizarry, C., Carpenter, A. C., Weng, A. P., Pear, W. S., Aster, J. C., and Blacklow, S. C. (2004). Notch subunit heterodimerization and prevention of ligand-independent proteolytic activation depend, respectively, on a novel domain and the LNR repeats. Mol Cell Biol 24, 9265-9273. Sapir, A., Assa-Kunik, E., Tsruya, R., Schejter, E., and Shilo, B. Z. (2005). Unidirectional Notch signaling depends on continuous cleavage of Delta. Development 132, 123-132. Sawamoto, K., Okabe, M., Tanimura, T., Mikoshiba, K., Nishida, Y., and Okano, H. (1996). The Drosophila secreted protein Argos regulates signal transduction in the Ras/MAPK pathway. Dev Biol 178, 13-22. Sawamoto, K., Okano, H., Kobayakawa, Y., Hayashi, S., Mikoshiba, K., and Tanimura, T. (1994). The function of argos in regulating cell fate decisions during Drosophila eye and wing vein development. Dev Biol 164, 267-276. Sawamoto, K., Taguchi, A., Hirota, Y., Yamada, C., Jin, M., and Okano, H. (1998). Argos induces programmed cell death in the developing Drosophila eye by inhibition of the Ras pathway. Cell Death Differ 5, 548. Schaeffer, H. J., and Weber, M. J. (1999). Mitogen-activated protein kinases: specific messages from ubiquitous messengers. Mol Cell Biol 19, 2435-2444. Schejter, E. D., Segal, D., Glazer, L., and Shilo, B. Z. (1986). Alternative 5’ exons and tissuespecific expression of the Drosophila EGF receptor homolog transcripts. Cell 46, 1091-1101. Schlessinger, J. (1988). Signal transduction by allosteric receptor oligomerization. Trends Biochem Sci 13, 443-447. Schlessinger, J. (1993). How receptor tyrosine kinases activate Ras. Trends Biochem Sci 18, 273275. Schlessinger, J. (2000). Cell signaling by receptor tyrosine kinases. Cell 103, 211-225. 112 Schnepp, B., Donaldson, T., Grumbling, G., Ostrowski, S., Schweitzer, R., Shilo, B. Z., and Simcox, A. (1998). EGF domain swap converts a drosophila EGF receptor activator into an inhibitor. Genes Dev 12, 908-913. Schnepp, B., Grumbling, G., Donaldson, T., and Simcox, A. (1996). Vein is a novel component in the Drosophila epidermal growth factor receptor pathway with similarity to the neuregulins. Genes Dev 10, 2302-2313. Schober, M., Rebay, I., and Perrimon, N. (2005). Function of the ETS transcription factor Yan in border cell migration. Development 132, 3493-3504. Schweisguth, F. (2004). Notch signaling activity. Curr Biol 14, R129-138. Schweisguth, F., and Posakony, J. W. (1992). Suppressor of Hairless, the Drosophila homolog of the mouse recombination signal-binding protein gene, controls sensory organ cell fates. Cell 69, 1199-1212. Schweisguth, F., and Posakony, J. W. (1994). Antagonistic activities of Suppressor of Hairless and Hairless control alternative cell fates in the Drosophila adult epidermis. Development 120, 1433-1441. Schweitzer, R., Howes, R., Smith, R., Shilo, B. Z., and Freeman, M. (1995). Inhibition of Drosophila EGF receptor activation by the secreted protein Argos. Nature 376, 699-702. Seger, R., and Krebs, E. G. (1995). The MAPK signaling cascade. Faseb J 9, 726-735. Seidah, N. G., and Prat, A. (2002). Precursor convertases in the secretory pathway, cytosol and extracellular milieu. Essays Biochem 38, 79-94. Seidel, J. J., and Graves, B. J. (2002). An ERK2 docking site in the Pointed domain distinguishes a subset of ETS transcription factors. Genes Dev 16, 127-137. Selkoe, D., and Kopan, R. (2003). Notch and Presenilin: regulated intramembrane proteolysis links development and degeneration. Annu Rev Neurosci 26, 565-597. Shaye, D. D., and Greenwald, I. (2005). LIN-12/Notch trafficking and regulation of DSL ligand activity during vulval induction in Caenorhabditis elegans. Development 132, 5081-5092. Sherrill, J. M. (1997). Self-phosphorylation of epidermal growth factor receptor is an intermolecular reaction. Biochemistry 36, 12890-12896. Shilo, B. Z. (2003). Signaling by the Drosophila epidermal growth factor receptor pathway during development. Exp Cell Res 284, 140-149. Shilo, B. Z. (2005). Regulating the dynamics of EGF receptor signaling in space and time. Development 132, 4017-4027. Silver, S. J., and Rebay, I. (2005). Signaling circuitries in development: insights from the retinal determination gene network. Development 132, 3-13. Simon, M. A., Bowtell, D. D., Dodson, G. S., Laverty, T. R., and Rubin, G. M. (1991). Ras1 and a putative guanine nucleotide exchange factor perform crucial steps in signaling by the sevenless protein tyrosine kinase. Cell 67, 701-716. Simon, M. A., Bowtell, D. D., and Rubin, G. M. (1989). Structure and activity of the sevenless protein: a protein tyrosine kinase receptor required for photoreceptor development in Drosophila. Proc Natl Acad Sci U S A 86, 8333-8337. Simon, M. A., Dodson, G. S., and Rubin, G. M. (1993). An SH3-SH2-SH3 protein is required for p21Ras1 activation and binds to sevenless and Sos proteins in vitro. Cell 73, 169-177. Simpson, P. (1997). Notch signalling in development: on equivalence groups and asymmetric developmental potential. Curr Opin Genet Dev 7, 537-542. Singson, A., Leviten, M. W., Bang, A. G., Hua, X. H., and Posakony, J. W. (1994). Direct 113 downstream targets of proneural activators in the imaginal disc include genes involved in lateral inhibitory signaling. Genes Dev 8, 2058-2071. Six, E., Ndiaye, D., Laabi, Y., Brou, C., Gupta-Rossi, N., Israel, A., and Logeat, F. (2003). The Notch ligand Delta1 is sequentially cleaved by an ADAM protease and gamma-secretase. Proc Natl Acad Sci U S A 100, 7638-7643. Skeath, J. B., and Carroll, S. B. (1994). The achaete-scute complex: generation of cellular pattern and fate within the Drosophila nervous system. Faseb J 8, 714-721. Slieker, L. J., and Lane, M. D. (1985). Post-translational processing of the epidermal growth factor receptor. Glycosylation-dependent acquisition of ligand-binding capacity. J Biol Chem 260, 687-690. Soderquist, A. M., and Carpenter, G. (1984). Glycosylation of the epidermal growth factor receptor in A-431 cells. The contribution of carbohydrate to receptor function. J Biol Chem 259, 12586-12594. Song, H., Nie, M., Qiao, F., Bowie, J. U., and Courey, A. J. (2005). Antagonistic regulation of Yan nuclear export by Mae and Crm1 may increase the stringency of the Ras response. Genes Dev 19, 1767-1772. Sotillos, S., Roch, F., and Campuzano, S. (1997). The metalloprotease-disintegrin Kuzbanian participates in Notch activation during growth and patterning of Drosophila imaginal discs. Development 124, 4769-4779. Spana, E. P., and Doe, C. Q. (1996). Numb antagonizes Notch signaling to specify sibling neuron cell fates. Neuron 17, 21-26. Spencer, S. A., and Cagan, R. L. (2003). Echinoid is essential for regulation of Egfr signaling and R8 formation during Drosophila eye development. Development 130, 3725-3733. Spencer, S. A., Powell, P. A., Miller, D. T., and Cagan, R. L. (1998). Regulation of EGF receptor signaling establishes pattern across the developing Drosophila retina. Development 125, 4777-4790. Steller, H., and Grether, M. E. (1994). Programmed cell death in Drosophila. Neuron 13, 12691274. Stowers, R. S., and Schwarz, T. L. (1999). A genetic method for generating Drosophila eyes composed exclusively of mitotic clones of a single genotype. Genetics 152, 1631-1639. Struhl, G., and Adachi, A. (2000). Requirements for presenilin-dependent cleavage of notch and other transmembrane proteins. Mol Cell 6, 625-636. Struhl, G., Fitzgerald, K., and Greenwald, I. (1993). Intrinsic activity of the Lin-12 and Notch intracellular domains in vivo. Cell 74, 331-345. Struhl, G., and Greenwald, I. (1999). Presenilin is required for activity and nuclear access of Notch in Drosophila. Nature 398, 522-525. Strutt, D. (2003). Frizzled signalling and cell polarisation in Drosophila and vertebrates. Development 130, 4501-4513. Strutt, H., and Strutt, D. (2003). EGF signaling and ommatidial rotation in the Drosophila eye. Curr Biol 13, 1451-1457. Sturtevant, M. A., Roark, M., and Bier, E. (1993). The Drosophila rhomboid gene mediates the localized formation of wing veins and interacts genetically with components of the EGF-R signaling pathway. Genes Dev 7, 961-973. Sun, Y., Jan, L. Y., and Jan, Y. N. (1998). Transcriptional regulation of atonal during development of the Drosophila peripheral nervous system. Development 125, 3731-3740. 114 Sundaram, M., and Han, M. (1996). Control and integration of cell signaling pathways during C. elegans vulval development. Bioessays 18, 473-480. Sundaram, M. V. (2005). The love-hate relationship between Ras and Notch. Genes Dev 19, 1825-1839. Sunnarborg, S. W., Hinkle, C. L., Stevenson, M., Russell, W. E., Raska, C. S., Peschon, J. J., Castner, B. J., Gerhart, M. J., Paxton, R. J., Black, R. A., and Lee, D. C. (2002). Tumor necrosis factor-alpha converting enzyme (TACE) regulates epidermal growth factor receptor ligand availability. J Biol Chem 277, 12838-12845. Tamura, K., Taniguchi, Y., Minoguchi, S., Sakai, T., Tun, T., Furukawa, T., and Honjo, T. (1995). Physical interaction between a novel domain of the receptor Notch and the transcription factor RBP-J kappa/Su(H). Curr Biol 5, 1416-1423. Tan, P. B., and Kim, S. K. (1999). Signaling specificity: the RTK/RAS/MAP kinase pathway in metazoans. Trends Genet 15, 145-149. Tang, A. H., Neufeld, T. P., Kwan, E., and Rubin, G. M. (1997). PHYL acts to down-regulate TTK88, a transcriptional repressor of neuronal cell fates, by a SINA-dependent mechanism. Cell 90, 459-467. Taniguchi, Y., Furukawa, T., Tun, T., Han, H., and Honjo, T. (1998). LIM protein KyoT2 negatively regulates transcription by association with the RBP-J DNA-binding protein. Mol Cell Biol 18, 644-654. Tanoue, T., and Nishida, E. (2003). Molecular recognitions in the MAP kinase cascades. Cell Signal 15, 455-462. Therrien, M., Chang, H. C., Solomon, N. M., Karim, F. D., Wassarman, D. A., and Rubin, G. M. (1995). KSR, a novel protein kinase required for RAS signal transduction. Cell 83, 879-888. Therrien, M., Michaud, N. R., Rubin, G. M., and Morrison, D. K. (1996). KSR modulates signal propagation within the MAPK cascade. Genes Dev 10, 2684-2695. Thomas, B. J. (2005). Cell-cycle control during development: taking it up a notch. Dev Cell 8, 451-452. Thomas, B. J., Gunning, D. A., Cho, J., and Zipursky, L. (1994). Cell cycle progression in the developing Drosophila eye: roughex encodes a novel protein required for the establishment of G1. Cell 77, 1003-1014. Tomlinson, A., and Ready, D. F. (1986). sevenless, a cell specific homeotic mutation of the Drosophila eye. Science 231, 400-402. Tomlinson, A., and Ready, D. F. (1987). Neuronal differentiation in the Drosophila ommatidium. Developmental Biology 120, 366-376. Tomlinson, A., and Struhl, G. (2001). Delta/Notch and Boss/Sevenless signals act combinatorially to specify the Drosophila R7 photoreceptor. Mol Cell 7, 487-495. Tootle, T. L., Lee, P. S., and Rebay, I. (2003). CRM1-mediated nuclear export and regulated activity of the Receptor Tyrosine Kinase antagonist YAN require specific interactions with MAE. Development 130, 845-857. Treisman, J. E., and Heberlein, U. (1998). Eye development in Drosophila: formation of the eye field and control of differentiation. Curr Top Dev Biol 39, 119-158. Treisman, J. E., Luk, A., Rubin, G. M., and Heberlein, U. (1997). eyelid antagonizes wingless signaling during Drosophila development and has homology to the Bright family of DNAbinding proteins. Genes Dev 11, 1949-1962. Tsai, C. C., Kao, H. Y., Yao, T. P., McKeown, M., and Evans, R. M. (1999). SMRTER, a 115 Drosophila nuclear receptor coregulator, reveals that EcR- mediated repression is critical for development. Mol Cell 4, 175-186. Tsruya, R., Schlesinger, A., Reich, A., Gabay, L., Sapir, A., and Shilo, B. Z. (2002). Intracellular trafficking by Star regulates cleavage of the Drosophila EGF receptor ligand Spitz. Genes Dev 16, 222-234. Tsuda, L., Inoue, Y. H., Yoo, M. A., Mizuno, M., Hata, M., Lim, Y. M., Adachi-Yamada, T., Ryo, H., Masamune, Y., and Nishida, Y. (1993). A protein kinase similar to MAP kinase activator acts downstream of the raf kinase in Drosophila. Cell 72, 407-414. Tsuda, L., Kaido, M., Lim, Y. M., Kato, K., Aigaki, T., and Hayashi, S. (2006). An NRSF/RESTlike repressor downstream of Ebi/SMRTER/Su(H) regulates eye development in Drosophila. Embo J. Tsuda, L., Nagaraj, R., Zipursky, S., and Banerjee, U. (2002). An EGFR/Ebi/Sno Pathway Promotes Delta Expression by Inactivating Su(H)/SMRTER Repression during Inductive Notch Signaling. Cell 110, 625. Tsuda, T., Ikeda, Y., and Taniguchi, N. (2000). The Asn-420-linked sugar chain in human epidermal growth factor receptor suppresses ligand-independent spontaneous oligomerization. Possible role of a specific sugar chain in controllable receptor activation. J Biol Chem 275, 21988-21994. Twomey, C., and McCarthy, J. V. (2005). Pathways of apoptosis and importance in development. J Cell Mol Med 9, 345-359. Uemura, T., Shepherd, S., Ackerman, L., Jan, L. Y., and Jan, Y. N. (1989). numb, a gene required in determination of cell fate during sensory organ formation in Drosophila embryos. Cell 58, 349-360. Ullrich, A., and Schlessinger, J. (1990). Signal transduction by receptors with tyrosine kinase activity. Cell 61, 203-212. Urban, S., and Freeman, M. (2002). Intramembrane proteolysis controls diverse signalling pathways throughout evolution. Curr Opin Genet Dev 12, 512. Urban, S., and Freeman, M. (2003). Substrate specificity of rhomboid intramembrane proteases is governed by helix-breaking residues in the substrate transmembrane domain. Mol Cell 11, 1425-1434. Urban, S., Lee, J. R., and Freeman, M. (2001). Drosophila rhomboid-1 defines a family of putative intramembrane serine proteases. Cell 107, 173-182. Urban, S., Lee, J. R., and Freeman, M. (2002a). A family of Rhomboid intramembrane proteases activates all Drosophila membrane-tethered EGF ligands. Embo J 21, 4277-4286. Urban, S., Schlieper, D., and Freeman, M. (2002b). Conservation of intramembrane proteolytic activity and substrate specificity in prokaryotic and eukaryotic rhomboids. Curr Biol 12, 1507-1512. Van De Bor, V., and Giangrande, A. (2001). Notch signaling represses the glial fate in fly PNS. Development 128, 1381-1390. van der Geer, P., Hunter, T., and Lindberg, R. A. (1994). Receptor protein-tyrosine kinases and their signal transduction pathways. Annu Rev Cell Biol 10, 251-337. Van Doren, M., Powell, P. A., Pasternak, D., Singson, A., and Posakony, J. W. (1992). Spatial regulation of proneural gene activity: auto- and cross-activation of achaete is antagonized by extramacrochaetae. Genes Dev 6, 2592-2605. Van Vactor, D. L., Jr., Cagan, R. L., Kramer, H., and Zipursky, S. L. (1991). Induction in the 116 developing compound eye of Drosophila: multiple mechanisms restrict R7 induction to a single retinal precursor cell. Cell 67, 1145-1155. Vassin, H., Bremer, K. A., Knust, E., and Campos-Ortega, J. A. (1987). The neurogenic gene Delta of Drosophila melanogaster is expressed in neurogenic territories and encodes a putative transmembrane protein with EGF-like repeats. Embo J 6, 3431-3440. Vennstrom, B., and Bishop, J. M. (1982). Isolation and characterization of chicken DNA homologous to the two putative oncogenes of avian erythroblastosis virus. Cell 28, 135-143. Vetrivel, K. S., Cheng, H., Lin, W., Sakurai, T., Li, T., Nukina, N., Wong, P. C., Xu, H., and Thinakaran, G. (2004). Association of gamma-secretase with lipid rafts in post-Golgi and endosome membranes. J Biol Chem 279, 44945-44954. Vinos, J., and Freeman, M. (2000). Evidence that Argos is an antagonistic ligand of the EGF receptor. Oncogene 19, 3560-3562. Vinson, C. R., Conover, S., and Adler, P. N. (1989). A Drosophila tissue polarity locus encodes a protein containing seven potential transmembrane domains. Nature 338, 263-264. Vivekanand, P., Tootle, T. L., and Rebay, I. (2004). MAE, a dual regulator of the EGFR signaling pathway, is a target of the Ets transcription factors PNT and YAN. Mech Dev 121, 14691479. Voas, M. G., and Rebay, I. (2004). Signal integration during development: insights from the Drosophila eye. Dev Dyn 229, 162-175. Vrailas, A. D., Marenda, D. R., Cook, S. E., Powers, M. A., Lorenzen, J. A., Perkins, L. A., and Moses, K. (2006). smoothened and thickveins regulate Moleskin/Importin 7-mediated MAP kinase signaling in the developing Drosophila eye. Development 133, 1485-1494. Vrailas, A. D., and Moses, K. (2006). Smoothened, thickveins and the genetic control of cell cycle and cell fate in the developing Drosophila eye. Mech Dev 123, 151-165. Wallberg, A. E., Pedersen, K., Lendahl, U., and Roeder, R. G. (2002). p300 and PCAF act cooperatively to mediate transcriptional activation from chromatin templates by notch intracellular domains in vitro. Mol Cell Biol 22, 7812-7819. Walters, J. W., Munoz, C., Paaby, A. B., and Dinardo, S. (2005). Serrate-Notch signaling defines the scope of the initial denticle field by modulating EGFR activation. Dev Biol 286, 415-426. Waltzer, L., Bourillot, P. Y., Sergeant, A., and Manet, E. (1995). RBP-J kappa repression activity is mediated by a co-repressor and antagonized by the Epstein-Barr virus transcription factor EBNA2. Nucleic Acids Res 23, 4939-4945. Wang, W., and Struhl, G. (2004). Drosophila Epsin mediates a select endocytic pathway that DSL ligands must enter to activate Notch. Development 131, 5367-5380. Wang, W., and Struhl, G. (2005). Distinct roles for Mind bomb, Neuralized and Epsin in mediating DSL endocytosis and signaling in Drosophila. Development 132, 2883-2894. Wang, Y., Lee, G. F., Kelley, R. F., and Spellman, M. W. (1996). Identification of a GDP-Lfucose:polypeptide fucosyltransferase and enzymatic addition of O-linked fucose to EGF domains. Glycobiology 6, 837-842. Wang, Y., Shao, L., Shi, S., Harris, R. J., Spellman, M. W., Stanley, P., and Haltiwanger, R. S. (2001). Modification of epidermal growth factor-like repeats with O-fucose. Molecular cloning and expression of a novel GDP-fucose protein O-fucosyltransferase. J Biol Chem 276, 40338-40345. Wang, Y., and Spellman, M. W. (1998). Purification and characterization of a GDP-fucose: polypeptide fucosyltransferase from Chinese hamster ovary cells. J Biol Chem 273, 8112117 8118. Wassarman, D. A., Therrien, M., and Rubin, G. M. (1995). The Ras signaling pathway in Drosophila. Curr Opin Genet Dev 5, 44-50. Wasserman, J. D., and Freeman, M. (1998). An autoregulatory cascade of EGF receptor signaling patterns the Drosophila egg. Cell 95, 355-364. Watson, D. K., McWilliams-Smith, M. J., Nunn, M. F., Duesberg, P. H., O’Brien, S. J., and Papas, T. S. (1985). The ets sequence from the transforming gene of avian erythroblastosis virus, E26, has unique domains on human chromosomes 11 and 21: both loci are transcriptionally active. Proc Natl Acad Sci U S A 82, 7294-7298. Wei, S. Y., Escudero, L. M., Yu, F., Chang, L. H., Chen, L. Y., Ho, Y. H., Lin, C. M., Chou, C. S., Chia, W., Modolell, J., and Hsu, J. C. (2005). Echinoid is a component of adherens junctions that cooperates with DE-Cadherin to mediate cell adhesion. Dev Cell 8, 493-504. Wen, C., Metzstein, M. M., and Greenwald, I. (1997). SUP-17, a Caenorhabditis elegans ADAM protein related to Drosophila KUZBANIAN, and its role in LIN-12/NOTCH signalling. Development 124, 4759-4767. Wharton, K. A., Johansen, K. M., Xu, T., and Artavanis-Tsakonas, S. (1985a). Nucleotide sequence from the neurogenic locus notch implies a gene product that shares homology with proteins containing EGF-like repeats. Cell 43, 567-581. Wharton, K. A., Yedvobnick, B., Finnerty, V. G., and Artavanis-Tsakonas, S. (1985b). opa: a novel family of transcribed repeats shared by the Notch locus and other developmentally regulated loci in D. melanogaster. Cell 40, 55-62. Wheeler, D. B., Bailey, S. N., Guertin, D. A., Carpenter, A. E., Higgins, C. O., and Sabatini, D. M. (2004). RNAi living-cell microarrays for loss-of-function screens in Drosophila melanogaster cells. Nat Methods 1, 127-132. Wheeler, D. B., Carpenter, A. E., and Sabatini, D. M. (2005). Cell microarrays and RNA interference chip away at gene function. Nat Genet 37 Suppl, S25-30. Whitson, K. B., Whitson, S. R., Red-Brewer, M. L., McCoy, A. J., Vitali, A. A., Walker, F., Johns, T. G., Beth, A. H., and Staros, J. V. (2005). Functional effects of glycosylation at Asn579 of the epidermal growth factor receptor. Biochemistry 44, 14920-14931. Wiersdorff, V., Lecuit, T., Cohen, S. M., and Mlodzik, M. (1996). Mad acts downstream of Dpp receptors, revealing a differential requirement for dpp signaling in initiation and propagation of morphogenesis in the Drosophila eye. Development 122, 2153-2162. Wigglesworth, V. B. (1940). Local and general factors in the development of ‘pattern’ in Rhodnius prolixus (Hemiptera). J Exp Biol 17, 180-200. Wilkin, M. B., and Baron, M. (2005). Endocytic regulation of Notch activation and downregulation (review). Mol Membr Biol 22, 279-289. Wilkin, M. B., Carbery, A. M., Fostier, M., Aslam, H., Mazaleyrat, S. L., Higgs, J., Myat, A., Evans, D. A., Cornell, M., and Baron, M. (2004). Regulation of notch endosomal sorting and signaling by Drosophila Nedd4 family proteins. Curr Biol 14, 2237-2244. Wilson, J. J., and Kovall, R. A. (2006). Crystal Structure of the CSL-Notch-Mastermind Ternary Complex Bound to DNA. Cell 124, 985-996. Wolff, T. (2003). EGF receptor signaling: putting a new spin on eye development. Curr Biol 13, R813-814. Wolff, T., and Ready, D. F. (1991). The beginning of pattern formation in the Drosophila compound eye: the morphogenetic furrow and the second mitotic wave. Development 113, 118 841-850. Wolff, T., and Ready, D. F. (1993). Pattern formation in the Drosophila retina. In The Development of Drosophila melanogaster, M. Bate, and A. Martinez Arias, eds. (Cold Spring Harbor, N.Y., Cold Spring Harbor Laboratory Press), pp. 1277-1325. Wu, J. Y., and Rao, Y. (1999). Fringe: defining borders by regulating the notch pathway. Curr Opin Neurobiol 9, 537-543. Wu, L., Aster, J. C., Blacklow, S. C., Lake, R., Artavanis-Tsakonas, S., and Griffin, J. D. (2000). MAML1, a human homologue of Drosophila mastermind, is a transcriptional co-activator for NOTCH receptors. Nat Genet 26, 484-489. Wu, L., and Griffin, J. D. (2004). Modulation of Notch signaling by mastermind-like (MAML) transcriptional co-activators and their involvement in tumorigenesis. Semin Cancer Biol 14, 348-356. Wurmbach, E., Wech, I., and Preiss, A. (1999). The Enhancer of split complex of Drosophila melanogaster harbors three classes of Notch responsive genes. Mech Dev 80, 171-180. Xiong, W. C., and Montell, C. (1993). tramtrack is a transcriptional repressor required for cell fate determination in the Drosophila eye. Genes Dev 7, 1085-1096. Xu, C., Kauffmann, R. C., Zhang, J., Kladny, S., and Carthew, R. W. (2000). Overlapping activators and repressors delimit transcriptional response to receptor tyrosine kinase signals in the Drosophila eye. Cell 103, 87-97. Xu, T., and Rubin, G. M. (1993). Analysis of genetic mosaics in developing and adult Drosophila tissues. Development 117, 1223-1237. Yamada, T., Okabe, M., and Hiromi, Y. (2003). EDL/MAE regulates EGF-mediated induction by antagonizing Ets transcription factor Pointed. Development 130, 4085-4096. Yang, C. H., Simon, M. A., and McNeill, H. (1999). mirror controls planar polarity and equator formation through repression of fringe expression and through control of cell affinities. Development 126, 5857-5866. Yang, L., and Baker, N. E. (2001). Role of the EGFR/Ras/Raf pathway in specification of photoreceptor cells in the Drosophila retina. Development 128, 1183-1191. Yang, L., and Baker, N. E. (2003). Cell cycle withdrawal, progression, and cell survival regulation by EGFR and its effectors in the differentiating Drosophila eye. Dev Cell 4, 359369. Yang, L., and Baker, N. E. (2006). Notch activity opposes ras-induced differentiation during the second mitotic wave of the developing Drosophilaeye. BMC Dev Biol 6, 8. Ye, Y., Lukinova, N., and Fortini, M. E. (1999). Neurogenic phenotypes and altered Notch processing in Drosophila Presenilin mutants. Nature 398, 525-529. Yeh, E., Dermer, M., Commisso, C., Zhou, L., McGlade, C. J., and Boulianne, G. L. (2001). Neuralized functions as an E3 ubiquitin ligase during Drosophila development. Curr Biol 11, 1675-1679. Yoon, H. G., Chan, D. W., Huang, Z. Q., Li, J., Fondell, J. D., Qin, J., and Wong, J. (2003). Purification and functional characterization of the human N-CoR complex: the roles of HDAC3, TBL1 and TBLR1. Embo J 22, 1336-1346. Yu, G., Nishimura, M., Arawaka, S., Levitan, D., Zhang, L., Tandon, A., Song, Y. Q., Rogaeva, E., Chen, F., Kawarai, T., et al. (2000). Nicastrin modulates presenilin-mediated notch/glp-1 signal transduction and betaAPP processing. Nature 407, 48-54. Yu, S. Y., Yoo, S. J., Yang, L., Zapata, C., Srinivasan, A., Hay, B. A., and Baker, N. E. (2002). 119 A pathway of signals regulating effector and initiator caspases in the developing Drosophila eye. Development 129, 3269-3278. Yuzawa, S., Yokochi, M., Hatanaka, H., Ogura, K., Kataoka, M., Miura, K., Mandiyan, V., Schlessinger, J., and Inagaki, F. (2001). Solution structure of Grb2 reveals extensive flexibility necessary for target recognition. J Mol Biol 306, 527-537. Zak, N. B., and Shilo, B. Z. (1992). Localization of DER and the pattern of cell divisions in wildtype and Ellipse eye imaginal discs. Dev Biol 149, 448-456. Zhang, F., Strand, A., Robbins, D., Cobb, M. H., and Goldsmith, E. J. (1994). Atomic structure of the MAP kinase ERK2 at 2.3 A resolution. Nature 367, 704-711. Zhang, J., Zhang, F., Ebert, D., Cobb, M. H., and Goldsmith, E. J. (1995). Activity of the MAP kinase ERK2 is controlled by a flexible surface loop. Structure 3, 299-307. Zhang, X., Gureasko, J., Shen, K., Cole, P. A., and Kuriyan, J. (2006). An allosteric mechanism for activation of the kinase domain of epidermal growth factor receptor. Cell 125, 1137-1149. Zheng, L., Zhang, J., and Carthew, R. W. (1995). frizzled regulates mirror-symmetric pattern formation in the Drosophila eye. Development 121, 3045-3055. Zhou, S., Fujimuro, M., Hsieh, J. J., Chen, L., and Hayward, S. D. (2000a). A role for SKIP in EBNA2 activation of CBF1-repressed promoters. J Virol 74, 1939-1947. Zhou, S., Fujimuro, M., Hsieh, J. J., Chen, L., Miyamoto, A., Weinmaster, G., and Hayward, S. D. (2000b). SKIP, a CBF1-associated protein, interacts with the ankyrin repeat domain of NotchIC To facilitate NotchIC function. Mol Cell Biol 20, 2400-2410. zur Lage, P., and Jarman, A. P. (1999). Antagonism of EGFR and notch signalling in the reiterative recruitment of Drosophila adult chordotonal sense organ precursors. Development 126, 3149-3157. zur Lage, P. I., Powell, L. M., Prentice, D. R., McLaughlin, P., and Jarman, A. P. (2004). EGF receptor signaling triggers recruitment of Drosophila sense organ precursors by stimulating proneural gene autoregulation. Dev Cell 7, 687-696. Zweifel, M. E., and Barrick, D. (2001). Studies of the ankyrin repeats of the Drosophila melanogaster Notch receptor. 2. Solution stability and cooperativity of unfolding. Biochemistry 40, 14357-14367. 120 Chapter 2. Split ends: molecular function and role in development David B. Doroquez and Ilaria Rebay 121 Introduction In order to study the cross-talk and integration of Epidermal Growth Factor Receptor (EGFR) and Notch signaling during Drosophila eye development, we hypothesized that the protein Split ends (Spen) was a likely regulator of EGFR-Notch pathway interactions (see Chapter 3): first, spen was identified to genetically interact with components of the EGFR pathway (Chen and Rebay, 2000; Dickson et al., 1996; Firth et al., 2000; Rebay et al., 2000) and, second, mammalian orthologs of Spen were identified molecularly as regulators of Notch signaling (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005). Here, we set the stage for our analyses of Spen by describing Spen’s role in Drosophila development and then describe its structure and function as well as its relevance in vertebrate development and human disease. spen is linked to defects in Notch and EGFR signaling It was in the Drosophila embryo that the first spen mutation was isolated in a screen to identify genes required for regulating length, fasciculation, and orientation of axons in the peripheral nervous system (PNS) (Kolodziej et al., 1995). In wild-type, lateral chordotonal (lch) neurons—five sensory neurons located in each abdominal hemisegment—originate as a midlateral cluster and ventrally send axons bundled in fascicles to connect to targets in the central nervous system (CNS). Loss-of-function mutations in spen result in improper axon target innervation, reduced number of lch neurons, and defasciculation of the axons, hence the mutation was anointed ‘split ends’ (Kolodziej et al., 1995; Kuang et al., 2000). spen mutations may contribute to defects in Notch signaling during lch PNS neuron development (Kuang et al., 2000). In wild-type, lch neuron development occurs step-wise such that three lch precursors are initially selected from a field of proneural Atonal (Ato)-expressing cells through the process of Notch-mediated lateral inhibition (Jarman et al., 1993; reviewed in Artavanis-Tsakonas et al., 1999). The three lch precursors then recruit two neighboring ectodermal cells to the lch neuronal fate and these cells differentiate into five lch neurons (Okabe and Okano, 1997). In spen mutants, variable numbers of Ato-expressing lch precursors are 122 observed during early stages of proneural cluster refinement. Some hemisegments had more precursors while others had fewer than wild-type (Kuang et al., 2000). These data suggest that loss of spen leads to failure in lch precursor formation or Notch-mediated lateral inhibition. In support of the latter, Suppressor of Hairless [Su(H)], the transcription factor required for Notch signal transduction, was reduced (Kuang et al., 2000). Therefore, spen functions positively with respect to the Notch pathway by maintaining the levels of its transcriptional effector, Su(H). This conclusion is not consistent with the roles of mammalian Spen orthologs as antagonists of the Notch pathway and it is not consistent with our observations of spen’s antagonism of the Notch pathway during eye development (see Chapter 3). However, opposite effects on the Notch pathway could represent tissue-specific differences in spen function. Axons in the embryonic CNS also have severe defects in guidance and extension (Chen and Rebay, 2000; Kuang et al., 2000). In the embryonic midline, midline glial cells (MGCs) are required for CNS morphogenesis and midline axon guidance (reviewed in Granderath and Klambt, 1999; Jacobs, 2000). MGCs normally migrate in a tight configuration during CNS development in wild-type embryos. However, MGCs do not properly migrate and diffusely spread out in spen mutants. Later in embryonic development, the quantity of MGCs at the midline is reduced in spen mutants, resulting in axon guidance defects. The activity of EGFR signaling is critical at the midline where it is required for proper MGC migration and cell survival (Hummel et al., 1997; Klambt et al., 1991; Seeger et al., 1993; Stemerdink and Jacobs, 1997). The spen phenotype is reminiscent of phenotypes observed in mutants of the EGFR pathway (Klambt et al., 1991). Consistent with this observation, spen mutants synergize with mutants in pnt, a positive effector of the EGFR pathway (Chen and Rebay, 2000). These data are consistent with a model where spen normally functions in a positive or parallel manner with respect to EGFR signaling during CNS development. It should be noted that the results of this analysis of spen function appears contradictory to those from a prior study that suggested spen antagonizes EGFR output (Kuang et al., 2000). Specifically, Kuang and colleagues reported elevated EGFR signaling in spen mutant embryos, as evidenced 123 by increased numbers of midline glial cells and loss of the EGFR pathway repressor Yan. However, these results have not been reproducible (F. Chen and I. Rebay, unpublished). spen’s involvement in the EGFR pathway was confirmed by eye-based genetic modifier screens to identify novel components of the pathway. spen alleles were isolated as enhancers of the rough eye phenotype associated with eye-specific expression of a constitutively-active Yan (sev-yanACT) (Rebay et al., 2000) and with loss-of-function alleles of the receptor tyrosine kinase (RTK) antagonist corkscrew (csw) (Firth et al., 2000). Subsequent in-depth analysis of spen in the eye has confirmed that it functions positively in or in parallel to the EGFR pathway (see Chapter 3). It should be noted that spen alleles were also pulled-out as enhancers of the rough eye phenotype associated with eye-specifically-driven activated Raf kinase (sev-RaftorY9) (Dickson et al., 1992; Dickson et al., 1996). This would indicate that spen [E(Raf)2A] could be a negative regulator of EGFR signaling, but the nature of this interaction is unclear. Spen is required for multiple processes during Drosophila development In addition to spen’s role in the embryonic nervous system, spen is required in multiple contexts at various stages of Drosophila development. Investigations into spen’s role in various tissues may provide fruitful information in the understanding of Spen function; however, the pleiotropic effects of spen, combined with its involvement in multiple signaling pathways, may complicate the analysis. Multiple independent genetic screens designed to identify regulators of the cell cycle recovered alleles of spen, suggesting possible involvement in proliferation control. spen alleles were isolated as enhancers of the rough eye phenotype associated with dE2F and dDP overexpression (GMR-dE2FdDP), which causes extra cell divisions through induction of the S phase (Du et al., 1996; Staehling-Hampton et al., 1999). spen alleles were also identified in a screen using ectopic Cyclin E (sev-CycE), which promotes entry into S phase (Lane et al., 2000). Thus, spen may act to regulate proliferation during eye development, although the particular mechanisms by which spen controls the cell cycle have yet to be determined. In addition, loss 124 of spen in the eye does not lead to dysregulation of the cell cycle (see Appendix II), although it is possible that genetic redundancy with other factors could mask a role for spen in this context. In the developing larva, spen is required during wing development for the positive regulation of Wingless (Wg)/Wnt signaling (Lin et al., 2003). spen was also shown to be involved in the regulation of planar cell polarity during wing and sensory bristle development (Mace and Tugores, 2004). Thus, spen is not only involved in Notch and EGFR signaling but may be important in other pathways. spen is also required in homeotic/Hox function during embryogenesis. In a screen to identify genes that interact with the homeotic gene Deformed (Dfd), spen (or polycephalon/poc) alleles were identified as promoting sclerite development in the head and suppressing head-like sclerite development in the thorax (Gellon et al., 1997; Wiellette et al., 1999). Head-like thoracic sclerites in spen mutants are a result of a wound-healing process initiated due to the defective integrity of the epidermal integument (Mace et al., 2005). The epithelial injury seen in spen mutants promotes the activity of Mitogen activated protein kinase (MAPK) as part of an injury response (Mace et al., 2005). In summary, the cases described above illustrate a wide role for spen in Drosophila development. Structure and molecular function of Spen family proteins In order to understand the function of spen, molecular gene cloning and identification of the encoded protein was performed independently by three different laboratories (Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999). The spen gene spans a 44.2-kb genomic region that is predicted to encode three large, differentially-spliced isoforms that are translated into 5560, 5533, and 5476-amino acid proteins of 600, 597, and 591-kDa molecular weight, respectively. These isoforms encode ubiquitously-expressed nuclear proteins (Fig. 2-1) (Chen and Rebay, 2000; Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999). The ubiquitously-expressed nuclear proteins encoded by the spen gene (Chen and Rebay, 2000; Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999) provide the founding members of 125 Fig. 2-1. The spen gene encodes three large proteins containing RRM and SPOC domains. (A) spen gene region (in green), transcript/CDS isoforms, and protein schematic. Spen has three tandem RNA-recognition motifs (RRM) and a Spen Paralog Ortholog C-terminus (SPOC). (B) Alignment of evolutionarily-conserved SPOC domain amongst Spen family members. 126 a family of proteins conserved from worm to human that is divided into two classes: large Spen family proteins >300 kDa (Grande class) or small Spen family proteins <125 kDa (Pequeño class). The genome of a given metazoan species encodes at least one Grande and one Pequeño Spen protein (Fig. 2-2). Spen family proteins are characterized by two domains: three tandem RNA-recognition motifs (RRMs) located at the N-terminus and a C-terminal conserved region termed the Spen Paralog Ortholog C-terminus (SPOC) (Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999). Very recently, a Spen-like protein, OsRRM, was identified in rice Oryza sativa to be required for growth. OsRRM contains two RRMs at the N-terminus and a SPOC domain located in the middle of the protein (Chen et al., 2007). In the following section, we describe studies that focus on elucidating the contribution of the RRMs and SPOC domain to Spen function. Fig. 2-2. Spen family of proteins. Spen family proteins have RRMs and a SPOC domain. There are two classes of proteins: Grande (large) and Pequeño (small) Spen family proteins. 127 The RRMs RRM-domain proteins mediate a wide variety of biological processes including mRNA splicing, stability, modification, transport, and translation (reviewed in Siomi and Dreyfuss, 1997)). In addition to interacting with RNA, RRM domains can also serve as protein-protein or protein-DNA interaction motifs (Samuels et al., 1998; reviewed in Maris et al., 2005). Given the diverse capabilities of RRM-domain proteins, it is important to understand how Spen family RRMs function. In Drosophila, the role for these RRMs is unknown except for structure-function analysis in S2 cultured cells which suggests that the RRMs are necessary for proper nuclear targeting of Spen (I. Rebay, unpublished data). The human Spen ortholog SMRT/HDAC1 Associated Repressor Protein (Sharp) and its paralog One Twenty-Two 1 (Ott1)/ RNA-Binding protein Motif 15 (Rbm15) was identified through protein pull-down assays in a complex with the spliceosome, suggesting a role for Spen-like proteins in splicing (Zhou et al., 2002). Indeed, a Pequeño-class human homolog of Spen, Ott3, was identified as a factor that acts post-transcriptionally to repress the accumulation of alternatively spliced mRNAs without affecting constitutively spliced mRNAs (Hiriart et al., 2005). This post-transcriptional repression is mediated through the binding of Ott3’s SPOC domain and the N-terminus of the EpsteinBarr virus mRNA export factor EB2. Furthermore, Sharp and Ott1 were also found to bind to EB2 (Hiriart et al., 2005). It has not been described whether the RRMs are actively involved in binding to mRNA in this context or if mRNA-binding occurs solely through EB2. In mouse, Ott1 also binds to RNA sequences of murine retrotransposons termed RNA transport elements (RTEs) (Lindtner et al., 2006). In this system, Ott1 binds in tandem with the mRNA export receptor NXF1 to stimulate RTE-mediated mRNA export and expression (Lindtner et al., 2006). RNA-binding by Spen-like proteins is not limited to mRNA processing, export, and/or splicing. For example, the RRMs of Sharp bind to a non-coding RNA termed SRA, a steroid receptor co-activator, known to associate with activator complexes that potentiate steroid hormone action (Lanz et al., 1999; Shi et al., 2001). Sharp sequesters this co-activator, thereby 128 129 Fig. 2-3. The human Spen ortholog Sharp represses nuclear receptor transcriptional activation and sequesters the non-coding RNA SRA. (A) Ligand-bound nuclear receptors recruit an activation complex that includes the Steroid Receptor Activator (SRA) non-coding RNA. (B) Sharp binds to nuclear receptors via a receptor interaction domain (RID) and recruits SMRT and HDAC to mediate repression. (C) Sharp also inhibits nuclear receptor activation by sequestering SRA from p72/p68. (D) A combined model for Sharp SRA sequestration and recruitment of a repressor complex inhibiting SRA-stimulated steroid receptor transcriptional activity (Fig. 2-3) (Shi et al., 2001). Therefore, Spen family proteins may indirectly regulate transcription. Spen family RRMs may play more direct roles in the transcriptional regulation process. In the case of the murine Spen ortholog Msx2-Interacting Nuclear Target (Mint), its RRMs selectively bind G/T-rich DNA sequences. For example, Sharp interacts with the proximal rat osteocalcin (OC) promoter to repress transcription during craniofacial development and bone morphogenesis (Newberry et al., 1999). This mode of regulation requires Mint’s interaction with the homeodomain transcription factor Msx2 via Mint’s SPOC domain (Newberry et al., 1999). These observations suggest that Spen family proteins may function in a transcriptional regulatory capacity. Indeed, Mint binds a 42-bp region of the OC promoter termed the OC Fibrinogen growth factor (FGF) Response Element, or OCFRE. In osteoblasts, Mint functions as a scaffold on the DNA that organizes and mediates transcriptional responses. In the active transcriptional state, FGF induces the master regulator of osteoblast differentiation Runx2 to bind the OCFRE and synergize with Mint. This generates a transcriptional ‘on’ state (Sierra et al., 2004). However, in the absence of FGF signaling, Msx2 binds to the OCFRE and Mint to elicit repression of osteoblast differentiation genes (Fig. 2-4) (Sierra et al., 2004). These observations suggest that Spen family proteins participate in the switch between transcriptional ‘off’ and ‘on’ states. The analysis above concerning Spen family RRMs indicates a variety of roles for these motifs. It is unclear whether the examples described above are species-specific functions of Spen family proteins or whether a Spen family protein of a given species has multiple roles in RNAand DNA-binding. The SPOC domain The Spen SPOC domain is a 168-amino acid long conserved region widely implicated as a protein interaction motif. Strongly-conserved SPOC domains are located at the extreme C-terminal region of the protein; however, loosely-conserved SPOC domains have also been 130 Fig. 2-4. Mint functions as a DNA scaffold mediating transcriptional activation and repression. (A) FGF signaling promotes Runx/CBF transcriptional activation of the osteocalcin gene. (B) Msx2, in combination with Mint, inhibits Runx/CBF activation of transcription. 131 identified at various positions in more distantly-related proteins (Sanchez-Pulido et al., 2004). The studies described in the section focus on the strongly-conserved SPOC domains that are found in Spen family proteins. Although a few studies have implicated Spen family proteins as cell death- or proteasome-associated proteins (Li et al., 2006; Sanchez-Pulido et al., 2004), there is mounting evidence in support of a primary role for Spen family proteins as regulators of transcription. For example, human Sharp was identified initially in a yeast two-hybrid screen as a transcriptional co-repressor that interacted with another co-repressor Steroid mediator for retinoid- and thyroidhormone receptors (SMRT)/Nuclear receptor co-repressor (NCoR) and Histone deacetylase 1 (HDAC1) (Shi et al., 2001). A nuclear receptor interaction domain (RID) was mapped to the middle of the Sharp protein and shown to mediate binding to the estrogen receptor, glutamate receptor, and the peroxisome-proliferator activated receptor-δ (Shi et al., 2001; Shi et al., 2002). These experiments led to the model where, in the absence of inducing hormone, Sharp is bound to the nuclear receptor leading to the recruitment of SMRT/NCoR and members of the nucleosome remodeling and histone deacetylase (NuRD) complex via the SPOC domain (Fig. 4-3) (Shi et al., 2001). This produces a transcriptionally silenced state. In the presence of inducing hormone, however, liganded nuclear receptor is no longer bound by Sharp (Shi et al., 2002). Therefore, co-activator proteins are recruited to the nuclear receptor to potentiate target transcription. In a similar manner, transcriptional repression from the carboxylated-sterol nuclear receptor DAF-12 is regulated by the Spen family protein DIN-1S during lipid metabolism, larval development, and aging in nematodes (Ludewig et al., 2004). Spen family proteins do not only bind to nuclear receptors to mediate repression, but they may also bind to an array of other transcription factors including the crystallin binding protein 1 (CRYBP1) in the regulation of Col2α expression (Yang et al., 2005) and RBP-J(κ) in the regulation of Notch pathway target genes (Kuroda et al., 2003; Oswald et al., 2002). Crystal structure analysis and site-directed mutagenesis of key residues affirms SharpSMRT interactions. The positively-charged, basic patch on the surface of the SPOC domain 132 binds directly to SMRT’s negatively-charged, acidic C-terminus (Ariyoshi and Schwabe, 2003). Mutation of key interaction residues on either protein abrogates association between Sharp and SMRT. The high conservation of the interacting regions indicates that transcriptional corepressor interaction may be common with Spen family proteins (Fig. 2-1B). The importance of the SPOC domain in Spen family protein function indicates that it may be an ideal target for modulating Spen’s activity. For example, over-expression of the Mint SPOC fragment alone abrogated the endogenous Mint from repressing transcription due to homodimerization of the SPOC domain (Li et al., 2005). Homodimerization of the SPOC domain may thus provide a mechanism for the attenuation of Spen’s co-repressor action. Indeed, we observed that over-expression of the Spen SPOC domain functions as a dominant-negative in vivo (see Chapter 4). How then might the monomer and homodimer states of Spen-like proteins be regulated? Some insight into this may follow from a yeast two-hybrid screen in which the p21-activated kinase 1 (Pak1) was identified as an interactor of Sharp. Pak1 is required for the phosphorylation of Sharp upstream of SPOC domain and this phosphorylation activity is required for the repression activity of Sharp (Vadlamudi et al., 2005). Although this remains to proven, phosphorylation at this upstream site may promote SPOC monomer formation. Spen family proteins are involved in developmental pathways and disease Transcriptional regulation is a key to affecting the inputs and outputs of signal transduction pathways during development. Studies of vertebrate Spen family proteins have provided key insight into how Spen family proteins function with respect to these signaling pathways. For example, Spen family proteins function as co-repressors of Notch signaling pathway targets (Fig. 2-5) (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005; reviewed in Tanigaki and Honjo, 2007). As in the case of the regulation of nuclear receptor gene targets, Sharp and Mint switch a transcription factor from eliciting an active to a repressed transcriptional state for Notch signaling pathway targets. Both Sharp and Mint bind to RBP133 Fig. 2-5. Sharp functions as a transcriptional repressor during Notch signaling. (A) Sharp interacts with RBP-J/Su(H) and SMRT to mediate repression. (B) Activated Notch competes for binding to RBP-J and recruits an activation complex for target gene transcription. 134 J(κ), the vertebrate ortholog of Drosophila Su(H) (Kuroda et al., 2003; Oswald et al., 2002). RBP-J(κ) is a bi-functional transcription factor that acts either to recruit a repressor complex to limit transcriptional activation of Notch pathway genes or to recruit an activator complex in the presence of intracellular Notch for target gene transactivation (Hsieh et al., 1999; Kao et al., 1998; Zhou and Hayward, 2001). Sharp and Mint facilitate RBP-J(κ)-mediated repression of genes like the Enhancer of Split [E(spl)] proteins HES and HEY by recruiting SMRT and Cterminal Binding Protein (CtBP)/CtBP-Interacting Protein (CtIP) repressor complexes to RBPJ(κ) (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005). Loss of Spen family gene function in vertebrates has severe consequences during development. Analysis of the Mint knock-out mouse showed that the Mint null is embryonic lethal with lethality beginning at E12 and death occurring by E14.5 (Kuroda et al., 2003; Tsuji et al., 2007). The associated phenotypes have been attributed to defects in Notch signaling. Errors in heart, muscle, and pancreas development as well as splenic B cell (Kuroda et al., 2003) and thymocyte T cell differentiation (Tsuji et al., 2007) are all similar to defects observed in mice defective for Notch signaling components in these tissues (Apelqvist et al., 1999; McCright et al., 2001; Pui et al., 1999; Tanigaki et al., 2002). The Pequeño class Spen-like protein Ott1 also regulates blood development via Notch. Ott1 is highly expressed in hematopoietic stem cells (HSCs) (Ma et al., 2007). The Ott1 mouse knock-out leads to defects in differentiation during lymphopoiesis and myelopoiesis (Ma et al., 2007; Raffel et al., 2007). As in the case for Sharp and Mint, Ott1 binds to RBP-J(κ) to mediate Notch pathway target repression. Regulation of transcription appears to be cell-type specific. In non-HSCs, Ott1 acts to repress Notch pathway target genes, whereas in HSCs, Ott1 has the opposite effect: Ott1 promotes Notch transactivation in HSCs (Ma et al., 2007). Ott1, therefore, facilitates the Notch differentiation program in HSCs. Interestingly, the Pequeño class Drosophila protein Nito opposes Spen function during eye development (Jemc and Rebay, 2006) which suggests that antagonism between Grande and Pequeño class Spen proteins may be conserved through the vertebrate lineage. 135 In humans, genetic defects in Ott1 leads to acute megakaryoblastic leukemia (AMKL) in children or to rare cases of acute myeloid leukemia (AML) in adults (Kawaguchi et al., 2005; Ma et al., 2001; Mercher et al., 2002; Mercher et al., 2001). The recurrent translocation of chromosome 1p13 to chromosome 22q13 [t(1;22)(p13;q13)] leads to an in-frame, almost full-length translational fusion of Ott1 with the Megakaryocytic Acute Leukemia (Mal)/ Megakaryoblastic Leukemia 1 (Mkl1) transcription factor (Ma et al., 2001; Mercher et al., 2001). The resultant Ott1-Mal fusion protein is likely to deregulate transcriptional targets of Ott1 and Mal. For example, the Ott1-Mal fusion may lead to the recruitment of Ott1-interacting transcriptional repressors or activators to target genes bound by Mal. Spen family proteins may affect cancer progression through other signaling pathways. In Drosophila, one aspect of Spen function is to positively regulate the Wg/Wnt signaling pathway (Lin et al., 2003). In humans, a subset of cancers results from the constitutive activation of the Wnt effector T-cell factor (TCF) through stabilization of the TCF activator β-catenin (β-cat) (Morin, 1999; Polakis, 2000). In human and mouse cancers, elevated β-cat/TCF activation of target genes is attributed to high Sharp and Mint levels, respectively (Feng et al., 2007). Sharp overexpression and RNA interference-mediated knock-down in colon cancer cells confirm Sharp’s role in cancer progression (Feng et al., 2007). In summary, Spen family proteins are essential during signal transduction processes and perturbation of Spen protein function may adversely affect development and cause disease. Concluding Remarks I have described in the discussion our current understanding of Spen molecular function in vertebrates and in Drosophila development. In order to better identify the function of Spen, it is important to understand the role of Spen family proteins in other systems, to draw hypotheses of Spen through this homology, and to test these experiments. My discussion has described a variety of functions for Spen from RNA processing and export to protein degradation to 136 transcriptional regulation in signal transduction. In the experiments described hereafter in this thesis, I describe a primary role for Spen in the regulation of signal transduction in the contexts of the EGFR and Notch signaling pathways. I then describe a screen to identify genetic interactors of spen in order to identify novel partners in Spen function. By using the power of genetics in Drosophila, my aim is to provide observations and models that will be helpful in understanding the role of Spen family proteins in signal transduction during development. References Apelqvist, A., Li, H., Sommer, L., Beatus, P., Anderson, D. J., Honjo, T., Hrabe de Angelis, M., Lendahl, U., and Edlund, H. (1999). Notch signalling controls pancreatic cell differentiation. Nature 400, 877-881. Ariyoshi, M., and Schwabe, J. W. (2003). A conserved structural motif reveals the essential transcriptional repression function of Spen proteins and their role in developmental signaling. Genes Dev 17, 1909-1920. Artavanis-Tsakonas, S., Rand, M. D., and Lake, R. J. (1999). Notch signaling: cell fate control and signal integration in development. Science 284, 770-776. Chen, F., and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10, 943-946. Chen, S. Y., Wang, Z. Y., and Cai, X. L. (2007). OsRRM, a Spen-like rice gene expressed specifically in the endosperm. Cell Res. Dickson, B., Sprenger, F., Morrison, D., and Hafen, E. (1992). Raf functions downstream of Ras1 in the Sevenless signal transduction pathway. Nature 360, 600-603. Dickson, B. J., van der Straten, A., Dominguez, M., and Hafen, E. (1996). Mutations Modulating Raf signaling in Drosophila eye development. Genetics 142, 163-171. Du, W., Xie, J. E., and Dyson, N. (1996). Ectopic expression of dE2F and dDP induces cell proliferation and death in the Drosophila eye. Embo J 15, 3684-3692. Feng, Y., Bommer, G. T., Zhai, Y., Akyol, A., Hinoi, T., Winer, I., Lin, H. V., Cadigan, K. M., Cho, K. R., and Fearon, E. R. (2007). Drosophila split ends homologue SHARP functions as a positive regulator of Wnt/beta-catenin/T-cell factor signaling in neoplastic transformation. Cancer Res 67, 482-491. Firth, L., Manchester, J., Lorenzen, J. A., Baron, M., and Perkins, L. A. (2000). Identification of genomic regions that interact with a viable allele of the Drosophila protein tyrosine phosphatase corkscrew. Genetics 156, 733-748. Gellon, G., Harding, K. W., McGinnis, N., Martin, M. M., and McGinnis, W. (1997). A genetic screen for modifiers of Deformed homeotic function identifies novel genes required for head development. Development 124, 3321-3331. Granderath, S., and Klambt, C. (1999). Glia development in the embryonic CNS of Drosophila. Curr Opin Neurobiol 9, 531-536. Hiriart, E., Gruffat, H., Buisson, M., Mikaelian, I., Keppler, S., Meresse, P., Mercher, T., Bernard, O. A., Sergeant, A., and Manet, E. (2005). Interaction of the Epstein Barr Virus (EBV) 137 mRNA export factor EB2 with human Spen proteins SHARP, OTT1 and a novel member of the family, OTT3, links Spen proteins with splicing regulation and mRNA export. J Biol Chem. Hsieh, J. J., Zhou, S., Chen, L., Young, D. B., and Hayward, S. D. (1999). CIR, a corepressor linking the DNA binding factor CBF1 to the histone deacetylase complex. Proc Natl Acad Sci U S A 96, 23-28. Hummel, T., Menne, T., Scholz, H., Granderath, S., Giesen, K., and Klambt, C. (1997). CNS midline development in Drosophila. Perspect Dev Neurobiol 4, 357-368. Jacobs, J. R. (2000). The midline glia of Drosophila: a molecular genetic model for the developmental functions of glia. Prog Neurobiol 62, 475-508. Jarman, A. P., Grau, Y., Jan, L. Y., and Jan, Y. N. (1993). atonal is a proneural gene that directs chordotonal organ formation in the Drosophila peripheral nervous system. Cell 73, 13071321. Jemc, J., and Rebay, I. (2006). Characterization of the split ends-like gene spenito reveals functional antagonism between SPOC family members during Drosophila eye development. Genetics 173, 279-286. Kao, H. Y., Ordentlich, P., Koyano-Nakagawa, N., Tang, Z., Downes, M., Kintner, C. R., Evans, R. M., and Kadesch, T. (1998). A histone deacetylase corepressor complex regulates the Notch signal transduction pathway. Genes Dev 12, 2269-2277. Kawaguchi, H., Hitzler, J. K., Ma, Z., and Morris, S. W. (2005). RBM15 and MKL1 mutational screening in megakaryoblastic leukemia cell lines and clinical samples. Leukemia 19, 14921494. Klambt, C., Jacobs, J. R., and Goodman, C. S. (1991). The midline of the Drosophila central nervous system: a model for the genetic analysis of cell fate, cell migration, and growth cone guidance. Cell 64, 801-815. Kolodziej, P. A., Jan, L. Y., and Jan, Y. N. (1995). Mutations that affect the length, fasciculation, or ventral orientation of specific sensory axons in the Drosophila embryo. Neuron 15, 273286. Kuang, B., Wu, S. C., Shin, Y., Luo, L., and Kolodziej, P. (2000). split ends encodes large nuclear proteins that regulate neuronal cell fate and axon extension in the Drosophila embryo. Development 127, 1517-1529. Kuroda, K., Han, H., Tani, S., Tanigaki, K., Tun, T., Furukawa, T., Taniguchi, Y., Kurooka, H., Hamada, Y., Toyokuni, S., and Honjo, T. (2003). Regulation of Marginal Zone B Cell Development by MINT, a Suppressor of Notch/RBP-J Signaling Pathway. Immunity 18, 301312. Lane, M. E., Elend, M., Heidmann, D., Herr, A., Marzodko, S., Herzig, A., and Lehner, C. F. (2000). A screen for modifiers of cyclin E function in Drosophila melanogaster identifies Cdk2 mutations, revealing the insignificance of putative phosphorylation sites in Cdk2. Genetics 155, 233-244. Lanz, R. B., McKenna, N. J., Onate, S. A., Albrecht, U., Wong, J., Tsai, S. Y., Tsai, M. J., and O’Malley, B. W. (1999). A steroid receptor coactivator, SRA, functions as an RNA and is present in an SRC-1 complex. Cell 97, 17-27. Li, J., Wang, J., Yang, X., Qin, H., Dong, X., Zhu, Y., Liang, L., Liang, Y., and Han, H. (2006). The Spen Homolog Msx2-Interacting Nuclear Target Protein Interacts with the E2 UbiquitinConjugating Enzyme UbcH8. Mol Cell Biochem 288, 151-157. 138 Li, J., Yang, X., Qin, H., Zhou, P., Liang, Y., and Han, H. (2005). The C terminus of MINT forms homodimers and abrogates MINT-mediated transcriptional repression. Biochim Biophys Acta 1729, 50-56. Lin, H. V., Doroquez, D. B., Cho, S., Chen, F., Rebay, I., and Cadigan, K. M. (2003). Splits ends is a tissue/promoter specific regulator of Wingless signaling. Development 130, 3125-3135. Lindtner, S., Zolotukhin, A. S., Uranishi, H., Bear, J., Kulkarni, V., Smulevitch, S., Samiotaki, M., Panayotou, G., Felber, B. K., and Pavlakis, G. N. (2006). RNA-binding motif protein 15 binds to the RNA transport element RTE and provides a direct link to the NXF1 export pathway. J Biol Chem 281, 36915-36928. Ludewig, A. H., Kober-Eisermann, C., Weitzel, C., Bethke, A., Neubert, K., Gerisch, B., Hutter, H., and Antebi, A. (2004). A novel nuclear receptor/coregulator complex controls C. elegans lipid metabolism, larval development, and aging. Genes Dev 18, 2120-2133. Ma, X., Renda, M. J., Wang, L., Cheng, E. C., Niu, C., Morris, S. W., Chi, A. S., and Krause, D. S. (2007). Rbm15 Modulates Notch-induced Transcriptional Activation and Affects Myeloid Differentiation. Mol Cell Biol. Ma, Z., Morris, S. W., Valentine, V., Li, M., Herbrick, J. A., Cui, X., Bouman, D., Li, Y., Mehta, P. K., Nizetic, D., et al. (2001). Fusion of two novel genes, RBM15 and MKL1, in the t(1;22)(p13;q13) of acute megakaryoblastic leukemia. Nat Genet 28, 220-221. Mace, K., and Tugores, A. (2004). The product of the split ends gene is required for the maintenance of positional information during Drosophila development. BMC Dev Biol 4, 15. Mace, K. A., Pearson, J. C., and McGinnis, W. (2005). An epidermal barrier wound repair pathway in Drosophila is mediated by grainy head. Science 308, 381-385. Maris, C., Dominguez, C., and Allain, F. H. (2005). The RNA recognition motif, a plastic RNAbinding platform to regulate post-transcriptional gene expression. Febs J 272, 2118-2131. McCright, B., Gao, X., Shen, L., Lozier, J., Lan, Y., Maguire, M., Herzlinger, D., Weinmaster, G., Jiang, R., and Gridley, T. (2001). Defects in development of the kidney, heart and eye vasculature in mice homozygous for a hypomorphic Notch2 mutation. Development 128, 491-502. Mercher, T., Busson-Le Coniat, M., Khac, F. N., Ballerini, P., Mauchauffe, M., Bui, H., Pellegrino, B., Radford, I., Valensi, F., Mugneret, F., et al. (2002). Recurrence of OTT-MAL fusion in t(1;22) of infant AML-M7. Genes Chromosomes Cancer 33, 22-28. Mercher, T., Coniat, M. B., Monni, R., Mauchauffe, M., Khac, F. N., Gressin, L., Mugneret, F., Leblanc, T., Dastugue, N., Berger, R., and Bernard, O. A. (2001). Involvement of a human gene related to the Drosophila spen gene in the recurrent t(1;22) translocation of acute megakaryocytic leukemia. Proc Natl Acad Sci U S A 98, 5776-5779. Morin, P. J. (1999). beta-catenin signaling and cancer. Bioessays 21, 1021-1030. Newberry, E. P., Latifi, T., and Towler, D. A. (1999). The RRM domain of MINT, a novel Msx2 binding protein, recognizes and regulates the rat osteocalcin promoter. Biochemistry 38, 10678-10690. Okabe, M., and Okano, H. (1997). Two-step induction of chordotonal organ precursors in Drosophila embryogenesis. Development 124, 1045-1053. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W., et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J 21, 5417-5426. Oswald, F., Winkler, M., Cao, Y., Astrahantseff, K., Bourteele, S., Knochel, W., and Borggrefe, T. 139 (2005). RBP-Jkappa/SHARP recruits CtIP/CtBP corepressors to silence Notch target genes. Mol Cell Biol 25, 10379-10390. Polakis, P. (2000). Wnt signaling and cancer. Genes Dev 14, 1837-1851. Pui, J. C., Allman, D., Xu, L., DeRocco, S., Karnell, F. G., Bakkour, S., Lee, J. Y., Kadesch, T., Hardy, R. R., Aster, J. C., and Pear, W. S. (1999). Notch1 expression in early lymphopoiesis influences B versus T lineage determination. Immunity 11, 299-308. Raffel, G. D., Mercher, T., Shigematsu, H., Williams, I. R., Cullen, D. E., Akashi, K., Bernard, O. A., and Gilliland, D. G. (2007). Ott1(Rbm15) has pleiotropic roles in hematopoietic development. Proc Natl Acad Sci U S A. Rebay, I., Chen, F., Hsiao, F., Kolodziej, P. A., Kuang, B. H., Laverty, T., Suh, C., Voas, M., Williams, A., and Rubin, G. M. (2000). A genetic screen for novel components of the Ras/Mitogen-activated protein kinase signaling pathway that interact with the yan gene of Drosophila identifies split ends, a new RNA recognition motif- containing protein. Genetics 154, 695-712. Samuels, M., Deshpande, G., and Schedl, P. (1998). Activities of the Sex-lethal protein in RNA binding and protein:protein interactions. Nucleic Acids Res 26, 2625-2637. Sanchez-Pulido, L., Rojas, A. M., van Wely, K. H., Martinez, A. C., and Valencia, A. (2004). SPOC: a widely distributed domain associated with cancer, apoptosis and transcription. BMC Bioinformatics 5, 91. Seeger, M., Tear, G., Ferres-Marco, D., and Goodman, C. S. (1993). Mutations affecting growth cone guidance in Drosophila: genes necessary for guidance toward or away from the midline. Neuron 10, 409-426. Shi, Y., Downes, M., Xie, W., Kao, H. Y., Ordentlich, P., Tsai, C. C., Hon, M., and Evans, R. M. (2001). Sharp, an inducible cofactor that integrates nuclear receptor repression and activation. Genes Dev 15, 1140-1151. Shi, Y., Hon, M., and Evans, R. M. (2002). The peroxisome proliferator-activated receptor delta, an integrator of transcriptional repression and nuclear receptor signaling. Proc Natl Acad Sci U S A 99, 2613-2618. Sierra, O. L., Cheng, S. L., Loewy, A. P., Charlton-Kachigian, N., and Towler, D. A. (2004). MINT, the Msx2 interacting nuclear matrix target, enhances Runx2-dependent activation of the osteocalcin fibroblast growth factor response element. J Biol Chem 279, 32913-32923. Siomi, H., and Dreyfuss, G. (1997). RNA-binding proteins as regulators of gene expression. Curr Opin Genet Dev 7, 345-353. Staehling-Hampton, K., Ciampa, P. J., Brook, A., and Dyson, N. (1999). A genetic screen for modifiers of E2F in Drosophila melanogaster. Genetics 153, 275-287. Stemerdink, C., and Jacobs, J. R. (1997). Argos and Spitz group genes function to regulate midline glial cell number in Drosophila embryos. Development 124, 3787-3796. Tanigaki, K., Han, H., Yamamoto, N., Tashiro, K., Ikegawa, M., Kuroda, K., Suzuki, A., Nakano, T., and Honjo, T. (2002). Notch-RBP-J signaling is involved in cell fate determination of marginal zone B cells. Nat Immunol 3, 443-450. Tanigaki, K., and Honjo, T. (2007). Regulation of lymphocyte development by Notch signaling. Nat Immunol 8, 451-456. Tsuji, M., Shinkura, R., Kuroda, K., Yabe, D., and Honjo, T. (2007). Msx2-interacting nuclear target protein (Mint) deficiency reveals negative regulation of early thymocyte differentiation by Notch/RBP-J signaling. Proc Natl Acad Sci U S A 104, 1610-1615. 140 Vadlamudi, R. K., Manavathi, B., Singh, R. R., Nguyen, D., Li, F., and Kumar, R. (2005). An essential role of Pak1 phosphorylation of SHARP in Notch signaling. Oncogene 24, 45914596. Wiellette, E. L., Harding, K. W., Mace, K. A., Ronshaugen, M. R., Wang, F. Y., and McGinnis, W. (1999). spen encodes an RNP motif protein that interacts with Hox pathways to repress the development of head-like sclerites in the Drosophila trunk. Development 126, 5373-5385. Yang, X., Li, J., Qin, H., Yang, H., Zhou, P., Liang, Y., and Han, H. (2005). Mint represses transactivation of the type II collagen gene enhancer through interaction with alpha Acrystallin-binding protein 1. J Biol Chem 280, 18710-18716. Zhou, S., and Hayward, S. D. (2001). Nuclear localization of CBF1 is regulated by interactions with the SMRT corepressor complex. Mol Cell Biol 21, 6222-6232. Zhou, Z., Licklider, L. J., Gygi, S. P., and Reed, R. (2002). Comprehensive proteomic analysis of the human spliceosome. Nature 419, 182-185. 141 142 Chapter 3. Split ends antagonizes the Notch and potentiates the EGFR signaling pathways during Drosophila eye development David B. Doroquez, Terry L. Orr-Weaver, and Ilaria Rebay Accepted for publication in Mechanisms of Development 143 Abstract The Notch and Epidermal Growth Factor Receptor (EGFR) signaling pathways interact cooperatively and antagonistically to regulate many aspects of Drosophila development, including the eye. How output from these two signaling networks is fine-tuned to achieve the precise balance needed for specific inductive interactions and patterning events remains an open and important question. Previously, we reported that the gene split ends (spen) functions within or parallel to the EGFR pathway during midline glial cell development in the embryonic central nervous system. Here, we report that the cellular defects caused by loss of spen function in the developing eye imaginal disc place spen as both an antagonist of the Notch pathway and a positive contributor to EGFR signaling during retinal cell differentiation. Specifically, loss of spen results in broadened expression of Scabrous, ectopic activation of Notch signaling, and a corresponding reduction in Atonal expression at the morphogenetic furrow. Consistent with Spen’s role in antagonizing Notch signaling, reduction of spen levels is sufficient to suppress Notch-dependent phenotypes. At least in part due to loss of Spen-dependent down-regulation of Notch signaling, loss of spen also dampens EGFR signaling as evidenced by reduced activity of MAP kinase (MAPK). This reduced MAPK activity in turn leads to a failure to limit expression of the EGFR pathway antagonist and the ETS-domain transcriptional repressor Yan and to a corresponding loss of cell fate specification in spen mutant ommatidia. We propose that Spen plays a role in modulating output from the Notch and EGFR pathways to ensure appropriate patterning during eye development. Introduction Signaling pathways are responsible for a host of cellular functions during development, including cell fate specification, morphogenesis, proliferation, differentiation, polarity 144 establishment, programmed cell death, and motility. These signaling pathways do not operate in isolation, but integrate with other networks to elicit specific activities (reviewed in Doroquez and Rebay, 2006; Hurlbut et al., 2007; Sundaram and Han, 1996; Sundaram, 2005; Tan and Kim, 1999; Voas and Rebay, 2003). Thus, a major challenge in molecular developmental genetics is to identify those genes that facilitate such coordination. The compound eye of the fruitfly Drosophila melanogaster is a paradigm for gaining insight into mechanisms that govern combinatorial signaling (Freeman, 1997; Ready et al., 1976; Voas and Rebay, 2003). The adult eye is composed of ~750 regularly repeated, light-sensing ommatidia that each contains eight photoreceptors (R cells) and 12 accessory cells (Ready et al., 1976). The eye develops from an imaginal disc primordium that originates embryonically and proliferates until the third larval instar, when cells arrest at the G1 stage of the cell cycle and apically constrict to form the morphogenetic furrow (MF) at the disc’s posterior edge (Wolff and Ready, 1991). As the MF travels anteriorly, Notch signaling establishes and then refines an initially broad expression pattern of the proneural bHLH transcription factor Atonal (Ato) for specification of evenly-spaced R8 photoreceptors, the founder cells of developing ommatidia, in processes known as proneural enhancement and lateral inhibition, respectively (Baker et al., 1996; Baker and Yu, 1997; Baonza and Freeman, 2001; Dokucu et al., 1996; Jarman et al., 1994; Li and Baker, 2001). Reiterative signaling by the Epidermal Growth Factor Receptor (EGFR)/Ras/MAP Kinase (MAPK) pathway (hereafter termed EGFR pathway) is then required for progressive and sequential recruitment of cell types, with the remaining neuronal photoreceptor cells added in a stereotypical pair-wise manner (R2/5, R3/4, R1/6, and finally R7), followed by non-neuronal cone and pigment cells (Freeman, 1996). Mechanistically, in response to EGFR activation, dualphosphorylated ERK (dpERK/activated MAPK) translocates into the nucleus to phosphorylate proteins including two ETS-domain transcription factors, Pointed-P2 (Pnt-P2) and Yan. Phosphorylation of Pnt-P2 potentiates its activator function, whereas phosphorylation abrogates Yan-mediated repression (Brunner et al., 1994; O’Neill et al., 1994; Rebay and Rubin, 1995). 145 This results in genes formerly repressed by Yan becoming activated by Pnt-P2, thereby initiating a cellular response appropriate to the original stimulus. The Notch and EGFR pathways interact intimately through genetic cooperation or antagonism; however, the molecular mechanisms that underlie these relationships remain largely unclear (reviewed in Doroquez and Rebay, 2006; Sundaram, 2005). One way by which the Notch pathway interfaces with the EGFR pathway is through regulation of Ato at the MF (Baonza and Freeman, 2001; Chen and Chien, 1999). Here, Ato is required for the expression of rhomboid (rho), which processes TGF-α-like ligands that activate EGFR (Baonza and Freeman, 2001; Lee et al., 2001; Urban et al., 2001; Urban et al., 2002), and for the activity of dpERK (Chen and Chien, 1999). Therefore, Notch-mediated lateral inhibition represses ato and limits EGFR signaling output. Since both Notch and EGFR networks orchestrate virtually all aspects of eye development, numerous mechanisms will undoubtedly be required for precise and context-appropriate coordination of pathway outputs. We initially identified split ends (spen) in a screen directed to uncover novel downstream effectors and regulators of the EGFR pathway (Rebay et al., 2000). Spen belongs to a family of nuclear proteins defined by two domains: (1) three RNA-recognition motifs (RRMs) at the Nterminus and (2) a C-terminal domain termed the Spen Paralog and Ortholog C-terminus (SPOC) (Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999). Consistent with the conservation of signaling pathways with which Spen has been implicated, Spen orthologs have been reported from worms to humans (Ludewig et al., 2004; Ma et al., 2001; Mercher et al., 2001; Newberry et al., 1999; Shi et al., 2001). Our genetic analysis showed that Spen functions in a positive manner with respect to EGFR signaling during Drosophila embryonic central nervous system development (Chen and Rebay, 2000; Rebay et al., 2000) whereas a parallel study suggested spen antagonizes EGFR output during embryonic neural development (Kuang et al., 2000). Additional studies have suggested that Spen may regulate a variety of other processes including the cell cycle, planar cell polarity, and cell death (Dickson et al., 1996; Firth et al., 2000; Kuang et al., 2000; Lane et al., 146 2000; Lin et al., 2003; Mace and Tugores, 2004; Mace et al., 2005; Sanchez-Pulido et al., 2004; Staehling-Hampton et al., 1999). This pleiotropy of developmental requirements positions Spen as an appealing candidate for integrating information from multiple signaling networks and the Drosophila eye provides an ideal context in which to explore Spen’s role during development. Spen family proteins have been best characterized as transcriptional co-repressors (Ariyoshi and Schwabe, 2003; Kuroda et al., 2003; Li et al., 2005; Ludewig et al., 2004; Ma et al., 2007; Oswald et al., 2002; Oswald et al., 2005; Shi, 2002; Shi et al., 2001; Tsuji et al., 2007; Vadlamudi et al., 2005; Yang et al., 2005). Notably, the mammalian orthologs human SHARP and mouse MINT interact with the Notch pathway transcriptional effector RBP-J(κ), an ortholog of Drosophila Suppressor of Hairless [Su(H)], and recruit a repressor complex (Ariyoshi and Schwabe, 2003; Kuroda et al., 2003; Li et al., 2005; Ma et al., 2007; Oswald et al., 2002; Shi et al., 2001; Tsuji et al., 2007; Yang et al., 2005). When Notch signaling is initiated, the repressor complex is competed away from RBP-J(κ) by Notchintra, allowing for activators to promote expression from target genes (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005). The developmental significance of spen function with respect to Notch signaling remains poorly understood in Drosophila. In this study, we explore the requirement for spen function during Drosophila eye development, toward the goal of elucidating how Spen influences output from the Notch and the EGFR pathways. Using somatic mosaic analysis to induce patches of spen mutant tissue in the developing eye imaginal disc, we examined expression of a collection of molecular markers that provide sensitive readouts of both Notch and EGFR pathway activity. Our results indicate that spen functions as a negative regulator of Notch signaling at the MF. Consistent with the elevated Notch signaling that occurs in spen mutant tissue, we observe a concomitant loss of Ato and dpERK at the MF, and thus reduced output from the EGFR pathway as evidenced further by cell fate specification defects. Together, our results reveal a role for Spen as an antagonist of Notch and a promoter of EGFR signaling during Drosophila retinal development. 147 Materials and Methods Fly Strains and Genetics Fly stocks were maintained at 25ºC and obtained from the Bloomington Stock Center. w1118 was used as the control strain, unless otherwise indicated. For the generation of mitotic clones in the eye imaginal disc, the following genotype was analyzed: P[ey-FLP/+ ; spenAH393 P[FRT]40A/P[w+, ubi-GFPnls] P[FRT]40A. For tangential sections, adult eyes were fixed, embedded in plastic, sectioned, and mounted as described (Wolff, 2000). Mutant tissue was recognizable by the absence of red eye pigment (Xu and Rubin, 1993). To generate eye discs predominantly mutant for spen, the EGUF/hid cell lethal technique (Stowers and Schwarz, 1999) was employed to produce the genotype: spenAH393 P[FRT]40 P[GMR-hid] l(2)cl-L31 P[FRT]40A ; P[ey-Gal4] P[UAS-FLP]. yan post-transcriptional regulation by Spen was analyzed in the following genotype: spenAH393 P[FRT]40A/P[ubi-GFPnls] P[FRT]40A ; P[dpp3-GAL4] P[UASFLP]/P[UAS-YanWT]. To analyze adult wing phenotypes, N54l9/+ and N54l9/+ ; spenAH393/+ flies were generated. Wings were fixed in 70% ethanol and mounted in Aquamount (BDH). Microscopy and image acquisition was performed on a Zeiss AxioPhot using a Spot digital imaging system (Diagnostic Instruments). Immunohistochemistry and Immunofluorescence Eye and wing imaginal discs were dissected from wandering third instar larvae, fixed with 4% paraformaldehyde in PBT (0.1% Triton X-100 in Phosphate-Buffered Saline [PBS]) for 15 min, washed three times in PBT (5 min each), incubated in primary antibody overnight in PBT with 5% normal goat serum (PNT) at 4ºC, and then washed three times in PBT (5 min each). Discs were incubated in secondary antibody for 2 hr in PNT at room temperature, washed three times in PBT (5 min each), and mounted in Vectashield mountant (Vector Laboratories). 148 Microscopy and image acquisition was performed on a Zeiss LSM510 confocal microsope. For immunofluorescence, the following primary antibodies were used: 1:5,000 rabbit anti-Atonal (Jarman et al., 1993), 1:30 mouse anti-Scabrous (Sca1, DSHB), 1:20 mouse antiNotchintra (C17.9C6, DSHB), 1:10 mouse anti-E(spl)-bHLH mAb323 (Jennings et al., 1994), 1:200 mouse anti-dpERK (M8159, Sigma), 1:200 mouse anti-Yan (8B12H9, DSHB), 1:50 rat anti-Elav (7E8A10, Developmental Studies Hybridoma Bank [DSHB]), 1:20,000 rabbit anti-βGalactosidase (Cappel Laboratories), 1:1000 guinea pig anti-Senseless (Nolo et al., 2000), 1:300 rabbit anti-Spalt, 1:500 rabbit anti-BarH1 (Higashijima et al., 1992), 1:10 mouse anti-Prospero (MR1A, DSHB), and 1:100 mouse anti-Cut (2B10, DSHB). Secondary antibodies were Alexa Fluor 568-conjugated goat anti-mouse IgG (Molecular Probes), or appropriate Rhodamine RedX- and Cy5-conjugated antibodies (Jackson ImmunoResearch). Quantitative RT-PCR Forty pairs of eye imaginal discs were dissected from control and spen/cell lethal clone third instar larvae. Total RNA was isolated using Trizol Reagent (Invitrogen) according to the manufacturer’s protocol, treated with DNase I (Invitrogen), and used as template to generate cDNA using the random or oligo-dT primers provided in the Reverse Transcription System (Promega). Quantitative/real-time PCR from 1 µL cDNA template was performed using 2× ABI SYBR Green Master Mix (Applied Biosystems), 80 nM forward primer, and 80 nM reverse primer in 25 µL reactions for 35 thermocycling reactions. Each experimental reaction was performed in triplicate using the relative quantitation method (ABI) or alongside four ten-fold dilutions of standard [wild-type eye disc cDNA] and no-template control reactions (in triplicate) as previously described in (Claycomb et al., 2002). For each sample, relative fluorescence was measured in comparison to standard curves. Mean and standard deviations of the triplicate reactions were calculated (ABI Prism 7300 software). Fold-expression of the experimental transcript was normalized to RpS17 expression for each sample. Analysis to determine statistical significance was performed with paired, two-tailed Student’s t-tests. 149 PCR primers are as follows: sca forward 5’- AACCGATTTCCCTAAACCAACC-3’ sca reverse 5’- CTTGATCTCTTTGGCATGCGACT-3’ aos forward 5’-CTTCCGTGACTACACTTGGACTT-3’ aos reverse 5’-CTATCTGCTCCGTCACATTCAAC-3’ RpS17 forward 5’-GTACGAACCAAGACGGTGAAGA-3’, RpS17 reverse 5’-GGCGAGTGTAGTACTTCTCGATG-3’, yan forward 5’-GGTATTAGCAGTGCCAGCAGTAA-3’, yan reverse 5’-ACTCCACCACTGGTCTCAGAGTA-3’ Results spen is required for proper proneural gene expression at the morphogenetic furrow To explore a coordinate role for spen in Notch and EGFR signaling, we chose to study its function in the developing compound eye, a tissue whose proper patterning requires reiterative inputs from both pathways. To circumvent the embryonic lethality of loss-of-function spen alleles (Chen and Rebay, 2000; Gellon et al., 1997; Kuang et al., 2000; Wiellette et al., 1999), we generated and analyzed somatic mosaic spen clones (see Materials and Methods). Because we did not observe proliferative or other defects suggestive of a critical role early in eye development (data not shown), we focused our investigation on the requirement for spen during the specification and differentiation events that occur in and posterior to the MF. We began by asking whether loss of spen altered the expression pattern of the proneural protein Ato. The pattern of proneural Ato in the eye occurs in four stages in wild-type imaginal discs (reviewed in Doroquez and Rebay, 2006; Frankfort and Mardon, 2002). First, Ato is expressed in a broad D-V stripe in the progressing MF during a process termed proneural enhancement (Stage 1) (Jarman et al., 1995). Ato is then restricted posteriorly via lateral 150 inhibition into the intermediate groups (IGs), the earliest stage of ommatidial cluster formation (Stage 2), of which two or three cells become equipotent in their ability to become the founding photoreceptor R8 (Stage 3) (Fig. 3-1A-C/Ato, GFP-positive tissue). A single cell within this R8 equivalence group becomes the R8 and maintains Ato expression (Stage 4) (Dokucu et al., 1996). In the majority of spen clones that cross the MF, Ato is significantly reduced during Stages 1-3 (Fig. 3-1A-C, GFP-negative tissue). However, in all cases, Ato expression ‘recovers’ and becomes refined to single cells—the presumptive R8s—that are spaced abnormally as compared to wild-type tissue (Fig. 3-1A-C). These data suggest that Spen is required for the maintenance of Ato proneural expression at the MF. Loss of spen leads to elevated levels of Sca protein and transcript Like Ato, the fibrinogen-related protein Scabrous (Sca) is expressed in the IGs, but is secreted to nearby cells for the restriction of Ato expression during lateral inhibition (Baker et al., 1990; Baker and Zitron, 1995; Lee et al., 1996; Mlodzik et al., 1990). It is proposed that endocytosis of Sca in nearby cells allows for Sca to prevent endocytic degradation of Notch, effectively activating the Notch pathway (Chou and Chien, 2002; Li et al., 2003; Powell et al., 2001). Since we observed down-regulation of Ato in spen eye clones, we hypothesized that spen loss could promote Ato inhibition via broadening or elevating Sca expression. Indeed, in spen clones, the expression domain of Sca is dramatically broadened at the MF, especially in the IG zone where Ato expression is down-regulated (Fig. 3-1A, GFP-negative tissue; Fig. 3-S1A,B [Supplementary Data]). To test whether the increase in Sca expression reflected increased transcription, we performed qRT-PCR to compare levels of sca transcript levels in spen mutant discs to those in wild-type discs. We find that sca mRNA levels are significantly elevated (1.6-fold; p = 0.009) in spen mutant discs as compared to wild-type (Fig. 3-2A). This suggests that in wild-type tissue, Spen directly or indirectly represses sca transcription or mRNA stability. 151 152 Fig. 3-1. Up-regulation of Notch pathway components in spen clones affects Atonal expression. Confocal micrograph of spen eye disc clones stained (in red) for (A) Scabrous, (B) Notchintra, (C) E(spl)-bHLH (mAb323). Discs were also co-stained for Ato (blue). spen mutant tissue lacks GFP (green). Ato staining is reduced in spen clones (A, arrow; B). Sca expression is broadened in spen clones (A, bracket) as compared to its restricted expression in IG cells (A, arrowhead). Notch expression is elevated in clones (B, arrow). E(spl)-bHLH expression is broadened in spen clones (C, bracket) and elevated posterior to the MF. E(spl)-bHLH is normally restricted to alternating groups of cells (C, arrows) complementary to Ato-positive IG cell expression in the MF. Discs are oriented anterior to the left. Scale bar = 20 µm. Loss of spen results in elevated Notch signaling at the morphogenetic furrow Because Sca facilitates Notch-mediated lateral inhibition (Baker and Zitron, 1995; Lee et al., 1996), broadly expressed Sca in spen clones would be predicted to activate the Notch pathway ectopically. To test this, markers for activation of Notch signaling were measured to determine the status of this pathway. First, we examined Notch expression with a monoclonal antibody specific to the intracellular domain of Notch that detects the cleaved, activated form of Notch (Notchintra) as well as full-length membrane-bound Notch (Fehon et al., 1990). In wildtype eye disc tissue, Notch is expressed predominantly apically throughout the eye imaginal disc tissue with a peak of expression at the MF and tapering anteriorly (Fig. 3-1B, GFP-positive tissue) (Baker and Yu, 1998; Baker and Zitron, 1995; Kidd et al., 1989). We found Notch protein levels elevated in spen clones with significant up-regulation at the MF (Fig. 3-1B, GFP-negative tissue). We also detected elevated levels of Notch using an antibody specific for the extracellular domain of Notch (Diederich et al., 1994) but did not detect significant change in Notch transcripts in spen mutants (Fig. 3-S1C,D and data not shown), consistent with the previous demonstration that elevated Sca expression stabilizes Notch protein levels (Powell et al., 2001). Because elevated receptor levels need not imply increased signaling output, we asked whether the Notch pathway was activated in spen clones. First, we examined the expression of the E(spl)-bHLH transcription factors, whose expression provides a direct read-out of Notch signal transduction (Bailey and Posakony, 1995; Baker et al., 1996; Dokucu et al., 1996; Jennings et al., 1994; Lecourtois and Schweisguth, 1995). With an antibody that interacts with four of the seven E(spl) bHLH proteins [E(spl)-mδ, -mγ, -mβ, and -m3], these proteins are detected at the MF in an alternating pattern with Ato-positive IG cells and in cells posterior to the MF (Fig. 3-1C, GFP-positive tissue; Fig. 3-S1E,F). The MF expression of E(spl)-bHLHs detected in wild-type tissue is consistent with a role for E(spl)-bHLHs in the repression of ato during Notch-mediated lateral inhibition (Baker et al., 1996; Dokucu et al., 1996). In spen clones, however, the alternating E(spl)-bHLH/Ato expression is disrupted such that E(spl)-bHLH is elevated and broadened at the MF (Fig. 3-1C, GFP-negative tissue; Fig. 3-S1E,F). Consistent 153 with the activation of Notch signaling in this tissue, we also detected elevated levels of E(spl)bHLH transcripts in spen mutant discs (Fig. 3-S2A,B). The expanded E(spl)-bHLH pattern is consistent with the repression of Ato observed in spen clones. In addition to broadening of E(spl)-bHLH expression at the MF, we observe modestly elevated levels of E(spl)-bHLH posterior to the MF in spen clones as compared to wild-type tissue (Fig. 3-1C; Fig. 3-S1E,F), indicating that loss of spen leads to prolonged hyperactivation of the Notch pathway. Reduction of spen suppresses Notch haploinsufficiency phenotypes in the wing The results just described are all consistent with Spen antagonizing the Notch pathway. To test this model further, we looked for dose-sensitive genetic interactions between spen and Notch. Although Notch null mutations are homozygous lethal, heterozygosity for Notch results in a dominant loss of tissue at the distal wing margin ([N54l9/+] 88% notched wing, n = 256) as compared to wild-type adult wings (0% notched wing, n = 100) (Fig. 3-2B,C) (Mohler, 1956), providing a sensitized genetic background in which to test for interactions. Consistent with our model that predicts a reduction in spen should elevate Notch levels, we found that adult flies doubly heterozygous for Notch and spen showed complete suppression of the wing notching phenotype ([N54l9/+; spenAH393/+] 0% notched wing, n = 48) (Fig. 3-2D,E). Loss of spen reduces dpERK levels in the developing eye What might be the functional consequences of reduced Ato expression and up-regulated Notch signaling in spen mutant eye tissue? Given that the Notch and EGFR pathways often operate antagonistically, we asked whether output from the EGFR pathway was perturbed. Specifically we asked whether dpERK (the dually-phosphorylated and activated form of MAPK) levels are affected. In wild-type discs, dpERK is most prominent in the IGs at the MF, and its expression pattern depends on Ato (Baonza et al., 2001; Chen and Chien, 1999). Posterior to the MF, dpERK is expressed at lower levels where it is predominantly cytoplasmic but also transits to the nucleus to phosphorylate target proteins (Fig. 3-3, GFP-positive tissue; Fig. 3-S3A,B) 154 Fig. 3-2. Antagonism of Notch by Spen: loss of spen leads to elevated sca transcript levels and heterozygous reduction of spen suppresses the Notch wing phenotype. (A) qRT-PCR was performed to measure relative sca mRNA levels in wild-type and spen/cell lethal eye discs (three samples, mean + st. dev.). Elevated levels are significant (*), p = 0.009. (B-E) Adult wing phenotypes of (B) wild-type, (C) Notch54l9/+, (D,E) Notch54l9/+; spenAH393/+ flies. Notch heterozygotes have loss of distal wing margin tissue (C, arrowhead). This is suppressed in Notch/+;spen/+ double-heterozygotes (D,E), where 16.7% of flies (n = 48) have aberrant LIII wing veins targeting to the margin (E, arrow). Scale bar = 0.3 mm. (Chen and Chien, 1999; Gabay et al., 1997; Kumar et al., 2003; Kumar et al., 1998; Spencer et al., 1998; Yang and Baker, 2003). Consistent with the loss of Ato expression in the IGs, we find that the dpERK expression in the IGs is drastically reduced in spen mutant clones (Fig. 3-3, arrow). dpERK levels posterior to the MF appear unchanged (Fig. 3-3, arrowhead). The reduced dpERK levels that result from loss of spen suggest Spen may potentiate EGFR signaling via antagonism of the Notch pathway (see Discussion). 155 Fig. 3-3. dpERK expression is lost at the MF in spen clones. Confocal micrograph of eye discs oriented with anterior to the left in spen clones stained for dpERK (red). spen mutant tissue lacks GFP (green). dpERK is lost at the MF in spen clones (B, arrow) but appears unchanged posterior to the MF (B, arrowhead). Scale bar = 20 µm. Transcriptional output from the EGFR pathway is perturbed in the absence of spen Do the reduced dpERK levels in spen mutant tissue dampen output from the EGFR signaling pathway? In wild-type animals, upon stimulation of EGFR signaling, phosphorylation of the transcriptional repressor Yan by dpERK leads to Yan nuclear export and degradation (Rebay and Rubin, 1995; Tootle et al., 2003). Therefore, we asked if Yan protein levels in wildtype versus spen mutant imaginal disc clones were different. In wild-type eye discs, Yan is expressed at the MF and in the basally located nuclei of undifferentiated cells posterior to the MF (Fig. 3-4A), but is down-regulated in the apically localized nuclei of differentiating cells (Fig. 3-4C) (Lai and Rubin, 1992). Thus, in developing wild-type eye discs, Yan expression is complementary to that of the pan-neuronal nuclear marker Elav, consistent with Yan’s function to block neuronal differentiation (Fig. 3-4A,C) (O’Neill et al., 1994; Rebay and Rubin, 1995). In spen clones, Yan expression is elevated in cells both at and posterior to the MF (Fig. 3-4B,D; Fig. 3-S3C,D). Western blot analysis of eye discs mutant for spen confirms that Yan protein levels are elevated relative to wild-type (data not shown). Posterior to the MF, Yan up-regulation is evident 156 157 Fig. 3-4. Yan is up-regulated in spen clones. Confocal micrographs of third instar eye imaginal discs oriented anterior to the left in wild-type (A,C) and spen clones (B,D) at basal (A,B) and apical (C,D) levels. Discs were stained for Yan (red) and Elav (blue), which are complementary in expression. spen mutant tissue is marked by lack of GFP (B,D; green). Yan is up-regulated in spen clones both in (B, arrowhead) and posterior (D, arrow) to the MF. Scale bar = 20 µm. throughout the clone in basally-localized nuclei (Fig. 3-4B) and in a subset of apical nuclei (Fig. 3-4D). As in wild-type, Yan and Elav expression in spen clones is largely complementary despite Yan’s encroachment into the apical plane, consistent with up-regulated Yan blocking these cells from adopting a neuronal fate (Fig. 3-4B,D). Mechanistically, elevated Yan protein levels should inappropriately repress expression of EGFR pathway target genes. For example, expression of argos (aos), which encodes a feedback inhibitor of EGFR signaling, is normally repressed by Yan in the absence of pathway activation and activated by Pnt upon pathway stimulation (Freeman, 1994; Gabay et al., 1996; Golembo et al., 1996; Sawamoto et al., 1996; Sawamoto et al., 1994; Schweitzer et al., 1995). Using quantitative real-time RT-PCR (qRT-PCR), we find that aos mRNA levels are significantly reduced approximately two-fold (p = 0.001) in spen mutant eye discs relative to wild-type, indicating that transcriptional output downstream of the EGFR signaling pathway is compromised (Fig. 3-5A). yan is not transcriptionally regulated by Spen Our results suggest that Spen regulates Yan levels post-translationally or posttranscriptionally, but it is alternatively possible that yan is regulated transcriptionally by Spen. This seems reasonable for two reasons: first, Notch signaling has been previously shown to transcriptionally activate yan expression directly (Rohrbaugh et al., 2002); and second, Spen family proteins have been implicated as transcriptional repressors of Notch pathway target genes (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005). To address this question, we asked whether loss of spen changes the expression of an enhancer-trap reporter, yanP[lacZ], that recapitulates the eye-specific yan mRNA expression (Lai and Rubin, 1992; Rohrbaugh et al., 2002). If Spen negatively regulates yan transcription, β-Gal reporter signal should be elevated in spen mutant clones. Analysis of spen clones generated in the yanP[lacZ] background detected no change in β-GAL levels relative to wild-type control tissue 158 Fig. 3-5. Loss of spen leads to reduced EGFR pathway target expression, but does not alter yan transcript levels. (A,C) qRT-PCR was performed to measure relative (A) argos and (C) yan mRNA levels in wild-type (w1118 and Oregon R, respectively) and spen/cell lethal eye discs (three samples, mean + st. dev.). The reduced aos transcript levels in spen discs is significant (*), p = 0.001, but yan transcript levels do not change, p = 0.602. (B) Confocal micrograph of spen clones generated in eye discs carrying the yanP[lacZ] enhancer-trap. β-Galactosidase levels are indistinguishable between control (GFP-positive) and spen mutant (GFP-negative) tissue. Discs are oriented anterior to the left. Scale bar = 20 µm. (Fig. 3-5B). In situ hybridization analysis similarly showed no change in yan expression in spen clones (data not shown). To confirm the above finding and to further analyze whether Spen might negatively regulate Yan expression by affecting the production or stability of yan mRNA, we examined yan mRNA levels via qRT-PCR analysis of spen mutant eye discs. We find that yan transcript levels in spen mutants are not significantly different than those measured in wild-type eye discs (Fig. 35C). Together, these data indicate that Spen does not inhibit yan mRNA production or stability. 159 Having ruled out transcriptional regulation of yan by Spen, we next asked if Spen regulates Yan post-transcriptionally. To investigate this, we compared the levels of Yan expressed from a cDNA transgene lacking all introns, the entire 5’UTR and the majority (all but 249 bp) of the 3’UTR in the presence or absence of spen. Because epitope tags significantly compromise Yan subcellular localization and function (T.L. Tootle and I. Rebay, unpublished results), we performed these experiments with untagged transgene constructs and detected expression with anti-Yan antibody (Rebay and Rubin, 1995). This precluded us from performing the experiment in the developing eye, as it would be impossible to distinguish endogenous from ectopically-expressed Yan. Therefore, we expressed wild-type Yan (YanWT) in the developing wing imaginal discs where, like the eye, Spen is ubiquitously expressed (Lin et al., 2003) and the EGFR signaling pathway is active (Gabay et al., 1997). However, Yan is not expressed in the wing (data not shown), allowing us to follow specifically the protein expressed from the cDNA transgene. To circumvent the larval lethality that results from ubiquitous expression of YanWT, we expressed UAS-yan transgenes under control of the dpp-GAL4 driver, which targets expression to a stripe of cells along the anterior-posterior (A-P) margin of the wing imaginal disc (Fig. 36A) (Morimura et al., 1996). As a control, we co-expressed YanWT and GFP and compared the two expression patterns. Interestingly, YanWT protein expression is limited to only a subset of GFP-positive cells; specifically, YanWT is not detected at the lateral edges of the dpp stripe (Fig. 3-6A; Fig. 3-S4A,B). Based on prior work showing that endogenous EGFR signaling induces robust down-regulation of Yan expressed from this transgene in a variety of developmental contexts (Rebay and Rubin, 1995), we reasoned that a similar phenomenon was occurring in the wing. Supporting the idea, dpERK is elevated at the lateral sides of the A-P border of the wing pouch (Gabay et al., 1997), precisely the region where ectopic dpp>YanWT is lost. Therefore, to test our model, we drove expression of the dpERK-unresponsive YanACT transgene (Rebay and Rubin, 1995) and found a close correlation between Yan-positive and GFP-positive nuclei (Fig. 3-6B; Fig. 3-S4C,D). These data show that accumulation of ectopic Yan protein in the wing 160 Fig. 3-6. Yan is up-regulated in spen wing clones that express ectopic Yan from a cDNA transgene. (A,B) Confocal micrographs of third instar wing imaginal discs that coexpress NLS-tagged GFP with (A) YanWT or (B) YanACT under the control of the dpp-GAL4 driver. dpp>YanWT (A, arrow) is expressed in a subset of cells where dpp>GFPNLS is driven (B, arrowhead). However, dpp>YanACT co-localizes well (B, arrow) with dpp>GFPNLS. (C) dpp>YanWT is expressed in a spen clones background. spen clones lack GFP (green). More red Yan-positive cells are seen in spen mutant tissue (C, arrowhead) than in control tissue (C, arrow). Discs are oriented dorsal up. Scale bar = 20 µm. 161 is regulated by EGFR signaling as it is in the eye, suggesting that the wing provides a relevant context in which to investigate Yan post-translational regulation by Spen. To ask if loss of spen function affects Yan protein levels when expressed under an exogenous promoter, we next generated spen clones in wing imaginal discs expressing dpp>YanWT. We found that in spen mutant wing clones, ectopic Yan is more broadly expressed compared to wild-type cells (Fig. 3-6C, Fig. 3-S4E,F). Importantly, generation of spen clones in an otherwise wild-type wing disc is not sufficient to induce expression of yan from the endogenous locus (data not shown). The observation that ectopic expression of Yan is required to observe up-regulation upon loss of spen and that the wing experiments are performed with a cDNA transgene, argue that Spen-mediated regulation of Yan occurs post-translationally, most likely via activation of ERK. spen is required for patterning of neuronal and non-neuronal cell types during eye development Because we observed elevated Notch and reduced EGFR pathway signaling activities in spen mutant clones, we were interested in how this disrupted signaling might affect recruitment and differentiation of the different cell types that comprise each ommatidium of the eye. To begin, we analyzed the phenotype of spen mutant clones in adult eyes. In wild-type adult eyes, tangential sections reveal regularly-spaced ommatidia, each containing eight trapezoidallyarrayed photoreceptor R cells (only seven of which are observed in any given plane) separated by a non-neuronal pigment cell lattice (Fig. 3-7A). Disorganization of ommatidia in spen mutant tissue reflects loss of R cells and likely disruption of the supporting lattice of non-neuronal accessory cells (Fig. 3-7B). The spen phenotype is highly penetrant between ommatidia. Sixtypercent of spen mutant ommatidia have reduced number of rhabdomeres. Four percent of spen mutant ommatidia have more than seven rhabdomeres, likely due to the failure of R7 and/or R8 positioning or the splitting of rhabdomeres within an ommatidium (data not shown). Of the 29 percent that showed seven rhabdomeres within tangential sections, 91 percent had abnormal 162 Fig. 3-7. Loss of spen disrupts morphology of the adult eye. Tangential sections of adult compound eyes from (A) wild-type (Oregon R) and (B) spen clones. Photoreceptor loss and ommatidial disorganization is evident in spen mutant tissue, which is marked by lack of the white gene product. Scale bar = 10 µm. rhabdomere morphology, orientation, and/or positioning of rhabdomeres in a trapezoidal pattern within an ommatidium. Thus, spen is required for the proper patterning and complement of ommatidial cell types, consistent with earlier descriptions of spen mutant eyes (Dickson et al., 1996). Considering the prominent role EGFR signaling plays in specifying cell fates in the developing eye (Freeman, 1997; Schweitzer and Shilo, 1997; Shilo and Raz, 1991), we investigated whether differentiation of specific cell types was compromised by loss of spen by examining expression of markers for neuronal and non-neuronal cell specification in third instar eye imaginal discs. In wild-type tissue, R cells, positive for the pan-neuronal nuclear marker Elav (Fig. 3-8A) (Robinow and White, 1991), are recruited sequentially to each ommatidium posterior to the MF, followed by addition of non-neuronal Cut-positive cone cells (Fig. 3-8K) (Blochlinger et al., 1990). In spen mutant discs, fewer Elav-positive cells are present relative to 163 164 Fig. 3-8. Loss of neuronal and non-neuronal ommatidial cell types in spen mutant tissue. Confocal micrographs of third instar eye imaginal discs in wild-type (A,C,E,G,I,K) and spen clones (B,D,F,H,J,L). Discs, oriented anterior to the left, were stained (red) for Elav (neuronal nuclei; A,B), Senseless (R8; C,D), Spalt (R3/4; E,F), BarH1 (R1/6; G,H), Prospero (R7; I,J), or Cut (cone; K,L). Elav (blue) was used to co-stain for ommatidial clusters (E-J). spen mutant tissue is marked by lack of GFP (green). In spen clones, there are fewer Elav-positive R cells per ommatidial cluster (B, arrows). The initial R8 spacing defect in spen clones (D, arrow) is later resolved (D, arrowhead). spen clones have ommatidial clusters that lack specific R cells [R3/4 (F, arrow), R1/6 (H, arrow), R7 (J, arrow)] and also have reduced expression of cone cell marker (L, asterisk). Scale bar = 20 µm. wild-type, suggesting that R cell specification is perturbed (Fig. 3-8B, arrows). spen’s effect on R cell development is quite penetrant, as demonstrated by analysis of later pupal stages in which the morphology of the tissue makes the reduced numbers of Elav-positive cells easier to view (Fig. 3-S5A-D). To determine which photoreceptors failed to develop within spen mutant ommatidia, cell-type specific R cell markers were examined. R8 is the first photoreceptor to differentiate in each ommatidium as marked by Senseless (Sens) expression (Fig. 3-8C) (Nolo et al., 2000). In spen clones, R8 spacing appears irregular (Fig. 3-8D, arrow) near the MF as compared to WT; however, this defect is resolved more posterior to the MF and R8 differentiation does not seem compromised in spen mutant tissue (Fig. 3-8D, arrowhead). The R8 spacing defect is consistent with Spen’s role in affecting Notch-mediated lateral inhibition of Ato at the MF (Fig. 3-1A). Furthermore, the R8 spacing defect seen here is also consistent with reduced EGFR signaling in spen tissue, as it has been previously shown that loss of dpERK activity in EGFR mutant clones at the MF leads to abnormal R8 spacing (Baonza and Freeman, 2001). We also analyzed the differentiation of R3/4, R1/6, and R7 as determined by Spalt (de Celis and Barrio, 2000), BarH1 (Higashijima et al., 1992), and Prospero (Pros) (Spana and Doe, 1995) expression, respectively (Fig. 3-8E-J). In spen mutants, loss of R3 and R4 occurs, but at fairly low penetrance (Fig. 3-8F, arrow). Loss of R1 and R6 in spen clones is more striking, although not fully penetrant between ommatidia (Fig. 3-8H, arrow). Like R3/4 and R1/6, R7 differentiation is also compromised in spen mutants (Fig. 3-8J, arrow). These results are consistent with the terminal phenotype of penetrant but irregularly patterned photoreceptor loss observed in adult eye sections (Fig. 3-7B). In contrast to variable loss of Elav-positive cells, loss or severe reduction in Cut expression is completely penetrant (Fig. 3-8L, asterisk), suggesting a role for spen in cone cell differentiation. As expected, during pupal development, fewer than the normal complement of four cone cells are present within most ommatidia (Fig. 3-S5E,F). Together these data indicate a requirement for spen in specification and differentiation 165 of neuronal and non-neuronal cell types during eye development, a function consistent with Spen playing a positive role within the EGFR pathway. Discussion In this study, we demonstrate that loss of spen perturbs the normal balance between the EGFR and Notch pathways as evidenced by the patterning disruptions and aberrant expression of multiple pathway components. These findings raise the question of whether Spen functions primarily in the Notch pathway, primarily in the EGFR pathway, or as a critical component of both. Although definite resolution is difficult given the extensive and intricate feedback regulation within and between these two signaling networks, we propose a model in which Spen-mediated antagonism of the Notch pathway regulates the signaling flow through the EGFR pathway to achieve proper retinal cell fate specification. Spen antagonizes Notch signaling Loss of spen results in hyperactivation of the Notch pathway as evidenced by elevated levels of both Notch and its transcriptional targets, the E(spl)-bHLHs. Therefore, a normal function of Spen in the developing eye is to limit the activity of Notch. Consistent with this model, we observed that heterozygous reduction of spen is sufficient to suppress the heterozygous Notch wing margin phenotype. However, loss of spen does not lead to the antineurogenic phenotypes typically associated with overexpression/overactivation of canonical members of the Notch pathway (reviewed in Artavanis-Tsakonas et al., 1999), suggesting that although Notch signaling output is elevated, the increase is below the threshold needed to achieve such phenotypes. Consistent with this interpretation, recruitment of the initial R8 photoreceptor neuron, a process influenced at multiple stages by Notch signaling, occurs normally in the absence of spen. 166 Where might Spen interface with the Notch signaling pathway? The striking increase in Sca expression in spen mutant clones at the MF is consistent with Spen regulating Notch activation by limiting the expression of sca either through transcriptional repression or by destabilizing the transcript. This suggests that in the Drosophila eye Spen may have an upstream role in the Notch pathway in contrast to the downstream role described for Spen mammalian orthologs (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005) On the other hand, because of extensive feedback regulation in Notch signaling, it is plausible that Spen interfaces with the network at a more downstream point. For example, ectopic expression of Notchintra was shown to promote Sca expression (Baker and Yu, 1997), which in turn is proposed to activate Notch signaling by preventing Notch degradation (Chou and Chien, 2002; Li et al., 2003; Powell et al., 2001). Additionally, although we did not detect such a role with respect to yan, it is possible that Spen limits Notchintra/Su(H)-mediated transactivation at the level of transcriptional repression of other Notch pathway targets, including the E(spl)-bHLHs, as is the case for the mammalian Spen orthologs (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005). This latter mechanism might also be relevant posterior the MF, where Notch signaling remains elevated as judged by increased levels of both Notch and the E(spl)-bHLHs in spen mutant clones, but where Sca is no longer expressed Although pinpointing where Spen interfaces with the Notch signaling pathway remains a challenge, the simplest interpretation of our data is that at the MF, Spen either directly or indirectly regulates Sca expression to restrict Notch pathway output. Posterior to the MF, as discussed below, mutual antagonism between the Notch and EGFR pathways may stabilize the initial signaling imbalance independent of Sca, leading to sustained up-regulation of Notch and down-regulation of EGFR output in spen mutant tissue. Elevated Notch pathway output leads to reduced EGFR signaling in spen mutant tissue What might the consequences of a moderate increase in Notch pathway output be? Given the extensive functional antagonism that has been previously reported between the Notch 167 and EGFR pathways in the eye (reviewed in Doroquez and Rebay, 2006; Sundaram, 2005), a likely outcome is that the increased Notch signaling in spen mutant tissue would dampen EGFR pathway output. Supporting the idea that spen plays a positive role with respect to EGFR signaling, the cell fate specification defects observed in spen mutant clones are highly reminiscent of phenotypes associated with hypomorphic mutants in positive components of the EGFR pathway (Dickson et al., 1992; Dominguez and de Celis, 1998; O’Neill et al., 1994; Silver et al., 2004; Simon et al., 1991). Thus, the defective specification of neuronal and non-neuronal cell types and the perturbed R8 spacing adjacent to the MF all suggest reduced, but not ablated, EGFR pathway function in spen clones. Lending further support to a model in which elevated Notch signaling in spen clones dampens EGFR pathway output, both Ato and dpERK expression at the MF are reduced. Because previous work has shown that Ato is required for activation of the EGFR pathway at the MF (Baonza et al., 2001; Chen and Chien, 1999), one possibility is that Spen stabilizes dpERK levels at the MF by antagonizing Notch-mediated lateral inhibition to ensure appropriate Ato expression. Another plausible mechanism for Spen-mediated regulation of dpERK activity would be downstream of or in parallel to Ras as we had proposed previously (Chen and Rebay, 2000). In this scenario, Spen might mediate transcriptional repression of an inhibitor such as a MAPK phosphatase (reviewed in Farooq and Zhou, 2004). However, qRT-PCR analysis in imaginal discs predominantly mutant for spen do not indicate a role for Spen in regulating the expression of two characterized Drosophila MAPK phosphatases—dMKP3 and PTP-ER (D.B. Doroquez and I. Rebay, unpublished observations; (Karim and Rubin, 1999; Kim et al., 2004; Kim et al., 2002; Rintelen et al., 2003). Thus, validation of such a mechanism will require identification of other MAPK phosphatases or pathway inhibitors that might be regulated by Spen It should be noted that the results of our analysis of spen function in the eye appear contradictory to those from a prior study that suggested spen antagonizes EGFR output and promotes Notch signaling during embryonic neural development (Kuang et al., 2000). 168 Specifically, the authors reported elevated EGFR signaling in spen maternal/zygotic null embryos, as evidenced by increased numbers of midline glial cells and loss of Yan expression. However, we have been unable to reproduce these results (F. Chen and I. Rebay, unpublished). On the contrary, our analysis of spen function during midline glial cell development in the embryonic central nervous system was entirely consistent with a role for spen as a positive contributor to EGFR signaling (Chen and Rebay, 2000). Thus, at least with respect to EGFR signaling, we believe Spen serves an analogous role in multiple developing tissues. With respect to Notch signaling, Kuang and colleagues report a strong reduction in E(spl)-bHLH expression throughout the embryo but no change in Notch levels, exactly opposite to our findings in the eye. Additional work will be needed to determine whether and how spen interfaces with the Notch pathway during embryogenesis, and whether distinct or identical mechanisms operate in retinal versus embryonic neural development. Spen modulates EGFR and Notch signaling posterior to the morphogenetic furrow It is not yet clear whether spen’s role in Notch-EGFR interactions posterior to the MF is identical to its role in events occurring at the MF. The failure to down-regulate Yan and the resulting cell fate specification defects show that EGFR signaling posterior to the MF is compromised in spen mutant tissue. Given that Yan up-regulation in spen clones does not result from loss of Spen-mediated transcriptional repression, but rather reflects loss of posttranslational control, two models for Spen function seem likely. First, if our inability to detect changes in dpERK protein levels posterior to the MF in situ accurately indicates unaltered dpERK levels, then the ability of dpERK to phosphorylate Yan must be compromised in spen mutants. Alternatively, dpERK levels may be sufficiently reduced to increase Yan stability, but the change may be below our immunohistochemical detection threshold. In terms of the signals that impinge on dpERK, whereas Notch signaling and Ato expression are critical for proper dpERK expression in the MF, reiterative EGFR signaling takes over posterior to the MF to maintain dpERK activity (Freeman, 1996). Thus, it is possible that a 169 Spen-dependent, Notch-independent mechanism may regulate EGFR output posterior to the MF. Alternatively, because Notch, E(spl) and Yan expression are all elevated in spen mutant tissue both in and posterior to the MF, Spen-mediated antagonism of Notch signaling may be relevant to EGFR regulation in both contexts. An extension of this idea that results in perhaps the most appealing model is that the initial increase in Notch output at the MF dampens EGFR signaling, which in turn leads to elevated Notch signaling in more posterior regions resulting in reduced EGFR output. In this way, the initial signaling imbalance created by loss of spen at the MF could be maintained over the entire eye disc through mutual antagonism and feedback regulation between the Notch and EGFR pathways. Conclusions In summary, we have analyzed the requirement for spen in regulating the EGFR and Notch pathways during Drosophila eye development and propose that increased Notch pathway activity upon loss of spen may be sufficient to dampen EGFR signaling, but not to disrupt other downstream effects of Notch signaling. Therefore, because the effects of spen loss appear to be at a threshold below the production of bona fide Notch-related phenotypes, we suggest that Spen plays a subtle role in the regulation of the Notch pathway or functions redundantly alongside other components. An equally likely hypothesis is that Spen regulates the Notch and EGFR pathways separately and that the phenotypes we have reported reflect a composite of independent disruptions to both signaling networks. Although much of the literature focuses on a primary role for Spen family proteins as corepressors, recent findings suggest members of this family may also regulate non-coding RNA sequestration, mRNA export, RNA splicing, and proteolysis (Hiriart et al., 2005; Li et al., 2006; Lindtner et al., 2006; Shi et al., 2001). Therefore, future identification of the precise molecular mechanisms by which Spen interfaces with the EGFR and Notch pathways may reveal novel modes of interaction between these two critical and conserved signaling networks. 170 Acknowledgements We would like to thank R.G. Fehon, P.A. Garrity, J.C. Jemc, P. Vivekanand, S. Maitra for comments on the manuscript; J.A. Lees and the Fehon, Orr-Weaver, and Rebay labs for stimulating discussion. We are grateful to A. Mukherjee, G. Hurlbut, and the ArtavanisTsakonas and Garrity labs for reagents, and to N. Watson (WIBR/Keck Microscopy Facility) for microscopy assistance. D.B.D. was supported by a National Science Foundation Graduate Fellowship. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. References Ariyoshi, M., and Schwabe, J. W. (2003). A conserved structural motif reveals the essential transcriptional repression function of Spen proteins and their role in developmental signaling. Genes Dev 17, 1909-1920. Artavanis-Tsakonas, S., Rand, M. D., and Lake, R. J. (1999). Notch signaling: cell fate control and signal integration in development. Science 284, 770-776. Bailey, A. M., and Posakony, J. W. (1995). Suppressor of hairless directly activates transcription of enhancer of split complex genes in response to Notch receptor activity. Genes Dev 9, 2609-2622. Baker, N. E., Mlodzik, M., and Rubin, G. M. (1990). Spacing differentiation in the developing Drosophila eye: a fibrinogen-related lateral inhibitor encoded by scabrous. Science 250, 1370-1377. Baker, N. E., Yu, S., and Han, D. (1996). Evolution of proneural atonal expression during distinct regulatory phases in the developing Drosophila eye. Curr Biol 6, 1290-1301. Baker, N. E., and Yu, S. Y. (1997). Proneural function of neurogenic genes in the developing Drosophila eye. Curr Biol 7, 122-132. Baker, N. E., and Yu, S. Y. (1998). The R8-photoreceptor equivalence group in Drosophila: fate choice precedes regulated Delta transcription and is independent of Notch gene dose. Mech Dev 74, 3-14. Baker, N. E., and Zitron, A. E. (1995). Drosophila eye development: Notch and Delta amplify a neurogenic pattern conferred on the morphogenetic furrow by scabrous. Mech Dev 49, 173189. Baonza, A., Casci, T., and Freeman, M. (2001). A primary role for the epidermal growth factor receptor in ommatidial spacing in the Drosophila eye. Curr Biol 11, 396-404. Baonza, A., and Freeman, M. (2001). Notch signalling and the initiation of neural development 171 in the Drosophila eye. Development 128, 3889-3898. Blochlinger, K., Bodmer, R., Jan, L. Y., and Jan, Y. N. (1990). Patterns of expression of cut, a protein required for external sensory organ development in wild-type and cut mutant Drosophila embryos. Genes Dev 4, 1322-1331. Brunner, D., Ducker, K., Oellers, N., Hafen, E., Scholz, H., and Klambt, C. (1994). The ETS domain protein pointed-P2 is a target of MAP kinase in the sevenless signal transduction pathway. Nature 370, 386-389. Chen, C. K., and Chien, C. T. (1999). Negative regulation of atonal in proneural cluster formation of Drosophila R8 photoreceptors. Proc Natl Acad Sci U S A 96, 5055-5060. Chen, F., and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10, 943-946. Chou, Y. H., and Chien, C. T. (2002). Scabrous controls ommatidial rotation in the Drosophila compound eye. Dev Cell 3, 839-850. Claycomb, J. M., MacAlpine, D. M., Evans, J. G., Bell, S. P., and Orr-Weaver, T. L. (2002). Visualization of replication initiation and elongation in Drosophila. J Cell Biol 159, 225-236. de Celis, J. F., and Barrio, R. (2000). Function of the spalt/spalt-related gene complex in positioning the veins in the Drosophila wing. Mech Dev 91, 31-41. Dickson, B., Sprenger, F., Morrison, D., and Hafen, E. (1992). Raf functions downstream of Ras1 in the Sevenless signal transduction pathway. Nature 360, 600-603. Dickson, B. J., van der Straten, A., Dominguez, M., and Hafen, E. (1996). Mutations Modulating Raf signaling in Drosophila eye development. Genetics 142, 163-171. Diederich, R. J., Matsuno, K., Hing, H., and Artavanis-Tsakonas, S. (1994). Cytosolic interaction between deltex and Notch ankyrin repeats implicates deltex in the Notch signaling pathway. Development 120, 473-481. Dokucu, M. E., Zipursky, S. L., and Cagan, R. L. (1996). Atonal, rough and the resolution of proneural clusters in the developing Drosophila retina. Development 122, 4139-4147. Dominguez, M., and de Celis, J. F. (1998). A dorsal/ventral boundary established by Notch controls growth and polarity in the Drosophila eye. Nature 396, 276-278. Doroquez, D. B., and Rebay, I. (2006). Signal Integration During Development: Mechanisms of EGFR and Notch Pathway Function and Cross-Talk. Crit Rev Biochem Mol Biol 41, 339385. Farooq, A., and Zhou, M. M. (2004). Structure and regulation of MAPK phosphatases. Cell Signal 16, 769-779. Fehon, R. G., Kooh, P. J., Rebay, I., Regan, C. L., Xu, T., Muskavitch, M. A., and ArtavanisTsakonas, S. (1990). Molecular interactions between the protein products of the neurogenic loci Notch and Delta, two EGF-homologous genes in Drosophila. Cell 61, 523-534. Firth, L., Manchester, J., Lorenzen, J. A., Baron, M., and Perkins, L. A. (2000). Identification of genomic regions that interact with a viable allele of the Drosophila protein tyrosine phosphatase corkscrew. Genetics 156, 733-748. Frankfort, B. J., and Mardon, G. (2002). R8 development in the Drosophila eye: a paradigm for neural selection and differentiation. Development 129, 1295-1306. Freeman, M. (1994). Misexpression of the Drosophila argos gene, a secreted regulator of cell determination. Development 120, 2297-2304. Freeman, M. (1996). Reiterative use of the EGF receptor triggers differentiation of all cell types in the Drosophila eye. Cell 87, 651-660. 172 Freeman, M. (1997). Cell determination strategies in the Drosophila eye. Development 124, 261270. Gabay, L., Scholz, H., Golembo, M., Klaes, A., Shilo, B. Z., and Klambt, C. (1996). EGF receptor signaling induces pointed P1 transcription and inactivates Yan protein in the Drosophila embryonic ventral ectoderm. Development 122, 3355-3362. Gabay, L., Seger, R., and Shilo, B. Z. (1997). In situ activation pattern of Drosophila EGF receptor pathway during development. Science 277, 1103-1106. Gellon, G., Harding, K. W., McGinnis, N., Martin, M. M., and McGinnis, W. (1997). A genetic screen for modifiers of Deformed homeotic function identifies novel genes required for head development. Development 124, 3321-3331. Golembo, M., Schweitzer, R., Freeman, M., and Shilo, B. Z. (1996). Argos transcription is induced by the Drosophila EGF receptor pathway to form an inhibitory feedback loop. Development 122, 223-230. Higashijima, S., Kojima, T., Michiue, T., Ishimaru, S., Emori, Y., and Saigo, K. (1992). Dual Bar homeo box genes of Drosophila required in two photoreceptor cells, R1 and R6, and primary pigment cells for normal eye development. Genes Dev 6, 50-60. Hiriart, E., Gruffat, H., Buisson, M., Mikaelian, I., Keppler, S., Meresse, P., Mercher, T., Bernard, O. A., Sergeant, A., and Manet, E. (2005). Interaction of the Epstein Barr Virus (EBV) mRNA export factor EB2 with human Spen proteins SHARP, OTT1 and a novel member of the family, OTT3, links Spen proteins with splicing regulation and mRNA export. J Biol Chem. Hurlbut, G. D., Kankel, M. W., Lake, R. J., and Artavanis-Tsakonas, S. (2007). Crossing paths with Notch in the hyper-network. Curr Opin Cell Biol 19, 166-175. Jarman, A. P., Grau, Y., Jan, L. Y., and Jan, Y. N. (1993). atonal is a proneural gene that directs chordotonal organ formation in the Drosophila peripheral nervous system. Cell 73, 13071321. Jarman, A. P., Grell, E. H., Ackerman, L., Jan, L. Y., and Jan, Y. N. (1994). Atonal is the proneural gene for Drosophila photoreceptors. Nature 369, 398-400. Jarman, A. P., Sun, Y., Jan, L. Y., and Jan, Y. N. (1995). Role of the proneural gene, atonal, in formation of Drosophila chordotonal organs and photoreceptors. Development 121, 20192030. Jennings, B., Preiss, A., Delidakis, C., and Bray, S. (1994). The Notch signalling pathway is required for Enhancer of split bHLH protein expression during neurogenesis in the Drosophila embryo. Development 120, 3537-3548. Karim, F. D., and Rubin, G. M. (1999). PTP-ER, a novel tyrosine phosphatase, functions downstream of Ras1 to downregulate MAP kinase during Drosophila eye development. Mol Cell 3, 741-750. Kidd, S., Baylies, M. K., Gasic, G. P., and Young, M. W. (1989). Structure and distribution of the Notch protein in developing Drosophila. Genes Dev 3, 1113-1129. Kim, M., Cha, G. H., Kim, S., Lee, J. H., Park, J., Koh, H., Choi, K. Y., and Chung, J. (2004). MKP-3 has essential roles as a negative regulator of the Ras/mitogen-activated protein kinase pathway during Drosophila development. Mol Cell Biol 24, 573-583. Kim, S. H., Kwon, H. B., Kim, Y. S., Ryu, J. H., Kim, K. S., Ahn, Y., Lee, W. J., and Choi, K. Y. (2002). Isolation and characterization of a Drosophila homologue of mitogen-activated protein kinase phosphatase-3 which has a high substrate specificity towards extracellular173 signal-regulated kinase. Biochem J 361, 143-151. Kuang, B., Wu, S. C., Shin, Y., Luo, L., and Kolodziej, P. (2000). split ends encodes large nuclear proteins that regulate neuronal cell fate and axon extension in the Drosophila embryo. Development 127, 1517-1529. Kumar, J. P., Hsiung, F., Powers, M. A., and Moses, K. (2003). Nuclear translocation of activated MAP kinase is developmentally regulated in the developing Drosophila eye. Development 130, 3703-3714. Kumar, J. P., Tio, M., Hsiung, F., Akopyan, S., Gabay, L., Seger, R., Shilo, B. Z., and Moses, K. (1998). Dissecting the roles of the Drosophila EGF receptor in eye development and MAP kinase activation. Development 125, 3875-3885. Kuroda, K., Han, H., Tani, S., Tanigaki, K., Tun, T., Furukawa, T., Taniguchi, Y., Kurooka, H., Hamada, Y., Toyokuni, S., and Honjo, T. (2003). Regulation of Marginal Zone B Cell Development by MINT, a Suppressor of Notch/RBP-J Signaling Pathway. Immunity 18, 301312. Lai, Z. C., and Rubin, G. M. (1992). Negative control of photoreceptor development in Drosophila by the product of the yan gene, an ETS domain protein. Cell 70, 609-620. Lane, M. E., Elend, M., Heidmann, D., Herr, A., Marzodko, S., Herzig, A., and Lehner, C. F. (2000). A screen for modifiers of cyclin E function in Drosophila melanogaster identifies Cdk2 mutations, revealing the insignificance of putative phosphorylation sites in Cdk2. Genetics 155, 233-244. Lecourtois, M., and Schweisguth, F. (1995). The neurogenic suppressor of hairless DNAbinding protein mediates the transcriptional activation of the enhancer of split complex genes triggered by Notch signaling. Genes Dev 9, 2598-2608. Lee, E. C., Hu, X., Yu, S. Y., and Baker, N. E. (1996). The scabrous gene encodes a secreted glycoprotein dimer and regulates proneural development in Drosophila eyes. Mol Cell Biol 16, 1179-1188. Lee, J. R., Urban, S., Garvey, C. F., and Freeman, M. (2001). Regulated intracellular ligand transport and proteolysis control EGF signal activation in Drosophila. Cell 107, 161-171. Li, J., Wang, J., Yang, X., Qin, H., Dong, X., Zhu, Y., Liang, L., Liang, Y., and Han, H. (2006). The Spen Homolog Msx2-Interacting Nuclear Target Protein Interacts with the E2 UbiquitinConjugating Enzyme UbcH8. Mol Cell Biochem 288, 151-157. Li, J., Yang, X., Qin, H., Zhou, P., Liang, Y., and Han, H. (2005). The C terminus of MINT forms homodimers and abrogates MINT-mediated transcriptional repression. Biochim Biophys Acta 1729, 50-56. Li, Y., and Baker, N. E. (2001). Proneural enhancement by Notch overcomes Suppressor-ofHairless repressor function in the developing Drosophila eye. Curr Biol 11, 330-338. Li, Y., Fetchko, M., Lai, Z. C., and Baker, N. E. (2003). Scabrous and Gp150 are endosomal proteins that regulate Notch activity. Development 130, 2819-2827. Lin, H. V., Doroquez, D. B., Cho, S., Chen, F., Rebay, I., and Cadigan, K. M. (2003). Splits ends is a tissue/promoter specific regulator of Wingless signaling. Development 130, 3125-3135. Lindtner, S., Zolotukhin, A. S., Uranishi, H., Bear, J., Kulkarni, V., Smulevitch, S., Samiotaki, M., Panayotou, G., Felber, B. K., and Pavlakis, G. N. (2006). RNA-binding motif protein 15 binds to the RNA transport element RTE and provides a direct link to the NXF1 export pathway. J Biol Chem 281, 36915-36928. Ludewig, A. H., Kober-Eisermann, C., Weitzel, C., Bethke, A., Neubert, K., Gerisch, B., Hutter, 174 H., and Antebi, A. (2004). A novel nuclear receptor/coregulator complex controls C. elegans lipid metabolism, larval development, and aging. Genes Dev 18, 2120-2133. Ma, X., Renda, M. J., Wang, L., Cheng, E. C., Niu, C., Morris, S. W., Chi, A. S., and Krause, D. S. (2007). Rbm15 Modulates Notch-induced Transcriptional Activation and Affects Myeloid Differentiation. Mol Cell Biol. Ma, Z., Morris, S. W., Valentine, V., Li, M., Herbrick, J. A., Cui, X., Bouman, D., Li, Y., Mehta, P. K., Nizetic, D., et al. (2001). Fusion of two novel genes, RBM15 and MKL1, in the t(1;22)(p13;q13) of acute megakaryoblastic leukemia. Nat Genet 28, 220-221. Mace, K., and Tugores, A. (2004). The product of the split ends gene is required for the maintenance of positional information during Drosophila development. BMC Dev Biol 4, 15. Mace, K. A., Pearson, J. C., and McGinnis, W. (2005). An epidermal barrier wound repair pathway in Drosophila is mediated by grainy head. Science 308, 381-385. Mercher, T., Coniat, M. B., Monni, R., Mauchauffe, M., Khac, F. N., Gressin, L., Mugneret, F., Leblanc, T., Dastugue, N., Berger, R., and Bernard, O. A. (2001). Involvement of a human gene related to the Drosophila spen gene in the recurrent t(1;22) translocation of acute megakaryocytic leukemia. Proc Natl Acad Sci U S A 98, 5776-5779. Mlodzik, M., Baker, N. E., and Rubin, G. M. (1990). Isolation and expression of scabrous, a gene regulating neurogenesis in Drosophila. Genes Dev 4, 1848-1861. Mohler, J. D. (1956). New mutants report. Drosophila Information Service 30, 78-79. Morimura, S., Maves, L., Chen, Y., and Hoffmann, F. M. (1996). decapentaplegic overexpression affects Drosophila wing and leg imaginal disc development and wingless expression. Dev Biol 177, 136-151. Newberry, E. P., Latifi, T., and Towler, D. A. (1999). The RRM domain of MINT, a novel Msx2 binding protein, recognizes and regulates the rat osteocalcin promoter. Biochemistry 38, 10678-10690. Nolo, R., Abbott, L. A., and Bellen, H. J. (2000). Senseless, a Zn finger transcription factor, is necessary and sufficient for sensory organ development in Drosophila. Cell 102, 349-362. O’Neill, E. M., Rebay, I., Tjian, R., and Rubin, G. M. (1994). The activities of two Ets-related transcription factors required for Drosophila eye development are modulated by the Ras/ MAPK pathway. Cell 78, 137-147. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W., et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J 21, 5417-5426. Oswald, F., Winkler, M., Cao, Y., Astrahantseff, K., Bourteele, S., Knochel, W., and Borggrefe, T. (2005). RBP-Jkappa/SHARP recruits CtIP/CtBP corepressors to silence Notch target genes. Mol Cell Biol 25, 10379-10390. Powell, P. A., Wesley, C., Spencer, S., and Cagan, R. L. (2001). Scabrous complexes with Notch to mediate boundary formation. Nature 409, 626-630. Ready, D. F., Hanson, T. E., and Benzer, S. (1976). Development of the Drosophila retina, a neurocrystalline lattice. Dev Biol 53, 217-240. Rebay, I., Chen, F., Hsiao, F., Kolodziej, P. A., Kuang, B. H., Laverty, T., Suh, C., Voas, M., Williams, A., and Rubin, G. M. (2000). A genetic screen for novel components of the Ras/Mitogen-activated protein kinase signaling pathway that interact with the yan gene of Drosophila identifies split ends, a new RNA recognition motif- containing protein. Genetics 154, 695-712. 175 Rebay, I., and Rubin, G. M. (1995). Yan functions as a general inhibitor of differentiation and is negatively regulated by activation of the Ras1/MAPK pathway. Cell 81, 857-866. Rintelen, F., Hafen, E., and Nairz, K. (2003). The Drosophila dual-specificity ERK phosphatase DMKP3 cooperates with the ERK tyrosine phosphatase PTP-ER. Development 130, 34793490. Robinow, S., and White, K. (1991). Characterization and spatial distribution of the ELAV protein during Drosophila melanogaster development. J Neurobiol 22, 443-461. Rohrbaugh, M., Ramos, E., Nguyen, D., Price, M., Wen, Y., and Lai, Z. C. (2002). Notch activation of yan expression is antagonized by RTK/pointed signaling in the Drosophila eye. Curr Biol 12, 576-581. Sanchez-Pulido, L., Rojas, A. M., van Wely, K. H., Martinez, A. C., and Valencia, A. (2004). SPOC: a widely distributed domain associated with cancer, apoptosis and transcription. BMC Bioinformatics 5, 91. Sawamoto, K., Okabe, M., Tanimura, T., Mikoshiba, K., Nishida, Y., and Okano, H. (1996). The Drosophila secreted protein Argos regulates signal transduction in the Ras/MAPK pathway. Dev Biol 178, 13-22. Sawamoto, K., Okano, H., Kobayakawa, Y., Hayashi, S., Mikoshiba, K., and Tanimura, T. (1994). The function of argos in regulating cell fate decisions during Drosophila eye and wing vein development. Dev Biol 164, 267-276. Schweitzer, R., Howes, R., Smith, R., Shilo, B. Z., and Freeman, M. (1995). Inhibition of Drosophila EGF receptor activation by the secreted protein Argos. Nature 376, 699-702. Schweitzer, R., and Shilo, B. Z. (1997). A thousand and one roles for the Drosophila EGF receptor. Trends Genet 13, 191-196. Shi, Y. (2002). Mechanisms of caspase activation and inhibition during apoptosis. Mol Cell 9, 459-470. Shi, Y., Downes, M., Xie, W., Kao, H. Y., Ordentlich, P., Tsai, C. C., Hon, M., and Evans, R. M. (2001). Sharp, an inducible cofactor that integrates nuclear receptor repression and activation. Genes Dev 15, 1140-1151. Shilo, B. Z., and Raz, E. (1991). Developmental control by the Drosophila EGF receptor homolog DER. Trends Genet 7, 388-392. Silver, S. J., Chen, F., Doyon, L., Zink, A. W., and Rebay, I. (2004). New class of Son-ofsevenless (Sos) alleles highlights the complexities of Sos function. Genesis 39, 263-272. Simon, M. A., Bowtell, D. D., Dodson, G. S., Laverty, T. R., and Rubin, G. M. (1991). Ras1 and a putative guanine nucleotide exchange factor perform crucial steps in signaling by the sevenless protein tyrosine kinase. Cell 67, 701-716. Spana, E. P., and Doe, C. Q. (1995). The prospero transcription factor is asymmetrically localized to the cell cortex during neuroblast mitosis in Drosophila. Development 121, 3187-3195. Spencer, S. A., Powell, P. A., Miller, D. T., and Cagan, R. L. (1998). Regulation of EGF receptor signaling establishes pattern across the developing Drosophila retina. Development 125, 4777-4790. Staehling-Hampton, K., Ciampa, P. J., Brook, A., and Dyson, N. (1999). A genetic screen for modifiers of E2F in Drosophila melanogaster. Genetics 153, 275-287. Stowers, R. S., and Schwarz, T. L. (1999). A genetic method for generating Drosophila eyes composed exclusively of mitotic clones of a single genotype. Genetics 152, 1631-1639. Sundaram, M., and Han, M. (1996). Control and integration of cell signaling pathways during C. 176 elegans vulval development. Bioessays 18, 473-480. Sundaram, M. V. (2005). The love-hate relationship between Ras and Notch. Genes Dev 19, 1825-1839. Tan, P. B., and Kim, S. K. (1999). Signaling specificity: the RTK/RAS/MAP kinase pathway in metazoans. Trends Genet 15, 145-149. Tootle, T. L., Lee, P. S., and Rebay, I. (2003). CRM1-mediated nuclear export and regulated activity of the Receptor Tyrosine Kinase antagonist YAN require specific interactions with MAE. Development 130, 845-857. Tsuji, M., Shinkura, R., Kuroda, K., Yabe, D., and Honjo, T. (2007). Msx2-interacting nuclear target protein (Mint) deficiency reveals negative regulation of early thymocyte differentiation by Notch/RBP-J signaling. Proc Natl Acad Sci U S A 104, 1610-1615. Urban, S., Lee, J. R., and Freeman, M. (2001). Drosophila rhomboid-1 defines a family of putative intramembrane serine proteases. Cell 107, 173-182. Urban, S., Lee, J. R., and Freeman, M. (2002). A family of Rhomboid intramembrane proteases activates all Drosophila membrane-tethered EGF ligands. Embo J 21, 4277-4286. Vadlamudi, R. K., Manavathi, B., Singh, R. R., Nguyen, D., Li, F., and Kumar, R. (2005). An essential role of Pak1 phosphorylation of SHARP in Notch signaling. Oncogene 24, 45914596. Voas, M., and Rebay, I. (2003). Signal integration during development: insights from the Drosophila eye. Mech Dev. Wiellette, E. L., Harding, K. W., Mace, K. A., Ronshaugen, M. R., Wang, F. Y., and McGinnis, W. (1999). spen encodes an RNP motif protein that interacts with Hox pathways to repress the development of head-like sclerites in the Drosophila trunk. Development 126, 5373-5385. Wolff, T. (2000). Histological techniques for the Drosophila eye part I: larva and pupa. In Drosophila Protocols, W. Sullivan, M. Ashburner, and R. S. Hawley, eds. (Cold Spring Harbor, NY, Cold Spring Harbor Laboratory Press), pp. 201-227. Wolff, T., and Ready, D. F. (1991). The beginning of pattern formation in the Drosophila compound eye: the morphogenetic furrow and the second mitotic wave. Development 113, 841-850. Xu, T., and Rubin, G. M. (1993). Analysis of genetic mosaics in developing and adult Drosophila tissues. Development 117, 1223-1237. Yang, L., and Baker, N. E. (2003). Cell cycle withdrawal, progression, and cell survival regulation by EGFR and its effectors in the differentiating Drosophila eye. Dev Cell 4, 359369. Yang, X., Li, J., Qin, H., Yang, H., Zhou, P., Liang, Y., and Han, H. (2005). Mint represses transactivation of the type II collagen gene enhancer through interaction with alpha Acrystallin-binding protein 1. J Biol Chem 280, 18710-18716. 177 Supplementary Data Fig. 3-S1. Up-regulation of Notch pathway components in spen clones. Confocal micrograph of spen eye discs clones stained (in red) for (A,B) Scabrous, (C,D) extracellular Notch (NotchEC, 1:200 C458.2H, DSHB) and (E,F) E(spl)bHLH (mAb323). spen mutant tissue lacks GFP (green). Sca expression is broadened in spen clones (A,B, brackets) as compared to its restricted expression in IG cells. Notch expression is elevated in clones (C,D, arrows). E(spl)-bHLH expression is broadened in spen clones (E,F, brackets) and elevated posterior to the MF. E(spl)-bHLH is normally restricted to alternating groups of cells (E, arrowheads). Discs are oriented anterior to the left. Scale bars = 20 µm. 178 Fig. 3-S2. E(spl)-bHLH transcripts are elevated in spen mutant discs. qRTPCR was performed to measure relative (A) E(spl)m3 and (B) E(spl)-mβ mRNA levels in wild-type and spen/ cell lethal eye discs (three samples, mean + st. dev.). Elevated levels are significant (*), p = 0.0026 and 0.018, respectively. 179 Fig. 3-S3. spen loss affects EGFR signaling pathway components. (A,B) Confocal micrographs of third instar eye imaginal discs oriented with anterior to the left in spen clones stained (in red) for (A,B) dpERK or (C,D) Yan. spen mutant tissue lacks GFP (green). dpERK is lost at the MF in spen clones (A,B, asterisks *) but appears unchanged posterior to the MF. Yan is up-regulated in spen clones both in and posterior to the MF (B,D, arrows). Scale bars = 20 µm. 180 Fig. 3-S4. Yan is up-regulated in spen wing clones that express ectopic Yan from a cDNA transgene. (A,B) confocal migraphs of third instar wing imaginal discs that co-expression NLS-tagged GFP with (A,B) YanWT or (B) YanACT under the control of the dpp-GAL4 driver. dpp>YanWT (A,B, arrows) is expressed in a subset of cells where dpp>GFPNLS is driven (A,B, arrowheads). However, dpp>YanACT colocalizes well (C,D, arrows) with dpp>GFPNLS. (E,F) dpp>YanWT is expressed in a spen clones background. spen clones lack GFP (green). More red Yan-positive cells are seen in spen mutant tissue (E,F, arrowheads) than in control tissue (E,F, arrow). Discs are oriented dorsal up. Scale bars = 20 µm. 181 Fig. 3-S5. Loss of organization and cell types in spen mutant pupal tissue. Confocal micrographs of eye imaginal discs 42 hr APF (after puparium formation) in wild-type (A,C,E) and spen clones (B,D,F). Discs were stained for (A,B) DE-cadherin (DE-cad, 1:50 DCAD2, DSHB), (C,D) Elav, and (E,F) Cut. spen mutant tissue is marked by lack of GFP (green). C,E and D,F correspond to the same sample, respectively. DE-cad (A,B) is a cell boundary marker that especially marks cone, pigment, and bristle cells. spen mutant tissue (B) shows loss of these cell types and loss of the regular ommatidial organization found in wild-type (A). In spen clones, there are fewer Elav-positive R cells (D) and Cut-positive cone cells (F). Scale bars = 10 µm. 182 Chapter 4. A screen for genetic modifiers of the rough eye associated with dominant-negative Spen in Drosophila David B. Doroquez, Fangli Chen, Eric Henckels, Emily Ludwig, Ilaria Rebay D. Doroquez performed the primary work. F. Chen generated the spenDN constructs and performed an initial screen pass through the deficiency kit. E. Henckels and E. Ludwig helped perform X-ray and EP screening. 183 Abstract Split ends (Spen) is the founding member of a family of proteins that are characterized to be transcriptional co-repressors. We previously identified Spen as a positive regulator of the Epidermal Growth Factor Receptor (EGFR) signaling pathway and as a negative regulator of Notch signaling during Drosophila eye development. In this report, we have over-expressed nuclear localization sequence (NLS)-tagged Spen C-terminus in the eye and identify that this represents a dominant-negative protein (SpenDN). The C-terminus includes the evolutionarily conserved SPOC domain that was shown in vertebrates to be required for transcriptional repression. SpenDN is likely to abrogate Spen function by homodimerizing with endogenous Spen. Here, we use eye-specifically expressed spenDN as a sensitized background to identify dominant genetic interactors. Our results reveal interactions with genes in the EGFR and Notch pathways and also suggest roles for Spen in chromatin regulation and programmed cell death. Introduction Signaling pathways require activator and repressor complexes to affect transcriptional regulation during development (reviewed in Doroquez and Rebay, 2006; Lai, 2002; Lee et al., 2001; Whitmarsh, 2006; Wotton and Massague, 2001). To understand how signal transduction moderates the switch between gene activation and repression, it is critical both to elucidate the components of activator and repressor complexes and to identify the specific processes affected by these transcriptional regulators. The compound eye of the fruitfly Drosophila melanogaster is a model system for understanding mechanisms of signal transduction interaction with transcriptional regulatory complexes. The adult eye is composed of ~750 regularly repeated, light-sensing ommatidia that each contains eight photoreceptor R cells and 12 accessory cells (Ready et al., 1976). The eye is particularly well-suited for genetic analysis because of its dispensability for viability and because 184 mutations that cause loss or gain of cell types invariably perturbs the structure, leading to obvious phenotypes. The basis for these defects may be identified through genetic and molecular analyses (reviewed in St Johnston, 2002). Epidermal Growth Factor Receptor (EGFR) and Notch signaling pathways interact closely during eye development (reviewed in Doroquez and Rebay, 2006; Hurlbut et al., 2007; Sundaram, 2005). It was in this context that we identified the protein Split ends (Spen) as a positive regulator of EGFR signaling and as an antagonist of the Notch pathway (see Chapter 3). Spen is the founding member of a family of nuclear proteins defined by two domains of interest: (1) three RNA-recognition motifs (RRMs) at the N-terminus and (2) a C-terminal domain termed the Spen Paralog and Ortholog C-terminus (SPOC) (Fig. 4-1A) (Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999). Structure-function analysis implicates Spen family proteins as transcriptional co- repressors (Ariyoshi and Schwabe, 2003; Kuroda et al., 2003; Li et al., 2005; Ludewig et al., 2004; Ma et al., 2007; Oswald et al., 2002; Oswald et al., 2005; Shi, 2002; Shi et al., 2001; Tsuji et al., 2007; Vadlamudi et al., 2005; Yang et al., 2005). The mammalian orthologs human Sharp and mouse Mint bind to the Notch pathway transcriptional effector Recombination signal Binding Protein-J [RBP-J(κ)], the vertebrate ortholog of Drosophila Suppressor of Hairless [Su(H)], and recruit a repressor complex (Ariyoshi and Schwabe, 2003; Kuroda et al., 2003; Li et al., 2005; Ma et al., 2007; Oswald et al., 2002; Shi et al., 2001; Tsuji et al., 2007; Yang et al., 2005). When Notch signaling is initiated, Notchintra and co-activators compete the repressor complex away from RBP-J(κ) for the activation of gene expression (Kuroda et al., 2003; Oswald et al., 2002; Oswald et al., 2005). Therefore, Spen family proteins are involved in a switch between ‘on’ and ‘off’ transcriptional states during development. The domain responsible for recruiting the repressor complex in vertebrates is the conserved 168-amino acid SPOC domain (Kuroda et al., 2003; Shi et al., 2001). The Sharp SPOC domain binds to the Steroid mediator for retinoid- and thyroid-hormone receptors (SMRT)/Nuclear receptor co-repressor (NCoR). The binding recruits members of the 185 nucleosome remodeling and histone deacetylase (NuRD) complex (Shi et al., 2001). Crystal structure analysis and site-directed mutagenesis confirms that Sharp and SMRT directly interact. The positively-charged, basic patch on the SPOC domain (Fig. 4-1C) binds directly to SMRT’s negatively-charged, acidic C-terminus (Ariyoshi and Schwabe, 2003). Spen’s identity with Sharp and Mint in the basic patch is significant (Fig. 4-1C). Additionally, Smrter (Smr), the functional homolog of SMRT, has the conserved C-terminal motif required to bind the basic region of the SPOC domain (Ariyoshi and Schwabe, 2003; Tsai et al., 1999). Thus, Drosophila Spen most likely also functions as a transcriptional repressor. We reported previously that over-expression of a Spen protein containing the RRMs but lacking the C-terminus functions as a dominant-negative (Chen and Chien, 1999). Here, we report that over-expression of nuclear-localized Spen C-terminus also causes a dominantnegative phenotype when expressed in the eye. In order to identify components related to Spen function and to understand the processes that Spen may be involved in, we used this dominantnegative spen as a sensitized background for genetic modifier screens in the eye. These screens are modeled after classical screens that have identified EGFR and Notch components in the eye (reviewed in St Johnston, 2002). Based on the results of the screen, the modifiers confirm a role for Spen in the EGFR and Notch signaling pathways as well as suggest a role in regulating cell death and chromatin state. Materials and Methods Fly Strains and Genetics Fly stocks were maintained at 25ºC and obtained from the Bloomington Stock Center. w1118 was used as the control strain, unless otherwise indicated. For the generation of mosaic mitotic clones for adult eye analysis, the following genotype was analyzed: P[ey-FLP/+ ; spenAH393 P[FRT]40A/P[w+, ubi-GFPnls] P[FRT]40A. For duplication analysis, the following 186 Fig. 4-1. Spen protein, transgenes, and conserved SPOC domain. (A) The spen gene encodes proteins ~5,500 amino acids in length. The protein includes three RNA-recognition motifs (RRMs), a coiled-coil domain (C-C), and a Spen Paralog and Ortholog C-terminus (SPOC). (B) cDNA encoding NLS-tagged Spen C-terminus was inserted into pUAST and pSev P-element vectors to generate dominant-negative spen (spenDN) transgenes. (C) The SPOC domain is conserved amongst Grande- and Pequeño-class Spen family proteins. High identity is found in the basic region. [Symbol] conservation: [*] identical, [:] and [.] degrees of sequence similarity. Residue (color): non-polar (red); polar (green); acidic (blue); basic (magenta). 187 genotype was analyzed: Dp(2;1)G146dpp/+ ; CyO P[sev-spenDN]/+. For analysis of smr interaction with spenDN, the following genotype was analyzed. smrP[BG01648]/+ ; CyO P[sevspenDN]/+. The smr allele is a P-element insertion within the 3’ end of the smr ORF; the nature of this allele has not been reported. Pilot F1 Deficiency Dominant Modifier Screen P[UAS-spenDN] lines were constructed by PCR amplifying the C-terminal 2.8-kb of the spen-coding region (amino acids 4540-5476) using the following oligos: 5’-ggaagatctatgccgaagaagaagcgcaaggtggttgccgccagtcatttggcacc-3’ 5’-ccgctcgagttagacagtagcgatgacaatcag-3’ PCR products were digested with BglII and XhoI, ligated into the pUAST vector (downstream of 5× UAS) and injected into w1118 embryos to obtain transgenics. The first primer contains a nuclear localization signal (PKKKRKV). To facilitate screening, the P[UAS-spenDN]3-8-5-1 transgenic line was recombined with the P[sev-Gal4]2B driver and balanced with TM6B on chromosome III. For screening, chromosome II- and III-balanced male flies with genomic deficiencies from the Bloomington Deficiency Kit were crossed to P[sev-Gal4]2B P[UASspenDN]3-8-5-1 virgin females and F1 progeny were scored for enhancement or suppression of the sev>spenDN rough-eye phenotype under a dissecting microscope. To determine sev­ promoterspecific interactions, dominant modification was tested with CyO P[sev-yanACT]/+ (Rebay et al., 2000) and CyO P[sev-RasV12]/+ (Karim et al., 1996). X-ray mutagenesis F1 dominant modifier screen P[sev-spenDN] lines were constructed by restriction digesting cDNA encoding NLS- tagged Spen(aa4540-5476) from pUAST-spenDN and ligating into the pSev vector (downstream of the sevenless (sev) promoter and upstream of the sev enhancer) and injected into w1118 embryos to obtain transgenics. To facilitate chromosomal linkage analysis and establishment of balanced stocks, the P[sev-spenDN] transgenes were mobilized onto CyO and TM6B balancer 188 chromosomes. Independent insertion events in which the spenDN phenotype co-segregated with the balancer were screened for strength, dose sensitivity, and variability of phenotype. Based on these criteria, another round of mobilization was performed to insert a second copy of sev-spenDN onto CyO 1×P[sev-spenDN] (SCS1) to generate the CyO 2×P[sev-spenDN] (SCS2). SCS2 was suitable for use in the modifier screen. The TM6B 1×P[sev-spenDN] (STS18) line was selected for use in determining chromosomal linkage. X-ray mutagenesis was performed similarly to the GMR/sev-yanACT screen described in Rebay et al., 2000. Specifically, 100 male w1118 flies isogenic for chromosomes II and III were irradiated with 5000 or 6000 R of X-rays (110 kVp, 16:40 or 20:00, 30.5 cm) using a Faxitron Cabinet X-ray System (Model 43855D) and then mated to 100 w1118; Sco/SCS2 virgin females. The eyes of F1 progeny were examined, as above, for enhancement or suppression of the sevspenDN phenotype. Potential modifiers were backcrossed to w1118; Sco/SCS2 and the F2 progeny were re-scored for enhancement or suppression. In the F2 generation, linkage of the mutation to chromosome II could be determined if all flies bearing the SCS2 were non-Sco and showed the appropriate modification; a balanced stock was established. If the modifier segregated randomly with respect to the SCS2 chromosome, this indicated chromosome III linkage. To establish a balanced stock, w1118; +/SCS2 ; mut/+ (mut = modifier) were crossed to w1118; TM3/STS18. In the next generation, a balanced stock of w1118; mut/STS18 was established by selecting STS18 flies showing the appropriate modification. Only chromosomes II and III were screened for modification because X-chromosome-linked modifiers are difficult to maintain in the spenDN backgrounds. Complementation tests based on lethality were performed among all mutations mapping to the same chromosome. Alleles that complement all other mutations on that chromosome are considered single hits. Mutations that are viable were not mapped; it is not possible to easily identify which genes are affected. 189 P Lethal and EP F1 Dominant Modifier Kit Screens An F1 dominant modifier screen of the sev-spenDN rough-eye phenotype was performed by crossing w1118; Sco/SCS2 virgin females with male flies that carry a Berkeley Drosophila Genome Project (BDGP) lethal P-element insertion (Spradling et al., 1999) or that carry an EnhancerPromoter (EP) P-element insertion (Rorth, 1996) on chromosomes II or III. The location of and the identity of the gene affected by each P-element insertion are listed in FlyBase (http://www. flybase.org). The F1­ progeny were screened for enhancement or suppression of the spenDN rougheye phenotype as described above. Histology Flies were prepared for scanning electron microscopy (SEM) by fixation in 1% glutaraldehyde/1% paraformaldehyde in 0.1 M sodium phosphate buffer for 2 hr as described (Tootle et al., 2003). The fixed tissue was dehydrated through an ethanol series. Samples were critical point dried, sputter coated, and images were acquired on a scanning electron microscope (JEOL 5600LV). Fixation and tangential sections of adult eyes was performed as previously described (Tomlinson and Ready, 1987). Results Over-expression of the Spen C-terminus is dominant-negative Structure-function analysis of Spen family proteins showed that the SPOC domain is required for their function (Kuroda et al., 2003; Shi et al., 2001) and over-expression of the mouse Spen-ortholog Mint’s SPOC domain in vitro abrogates transcriptional repression attributed to full-length Mint (Li et al., 2005). We were interested in determining whether over-expression of fragments of Spen in vivo could affect normal Spen function. We showed previously that over-expression from a transgene that encodes a protein that lacks the Spen C190 terminus (Spen∆C) functions as a dominant-negative allele of spen (Chen and Rebay, 2000)(see Appendix V). Here we performed the reciprocal experiment to determine if over-expression of the C-terminus that includes the SPOC domain can affect normal Spen function. Because the eye is easily amenable to observe failures in developmental patterning attributed to spen loss (see Chapter 3), we over-expressed the Spen C-terminus in the eye. In order to do this, we placed cDNA encoding SV40 large T antigen nuclear localization signal (NLS)-tagged spen C-terminal 937-amino acids (aa 4540-5476) under the control of eye-specific sevenless (sev) regulatory elements in a transposable P-element vector (sev-spenC-NLS=sev-spenDN; Fig. 4-1A,B) and generated transgenic flies (Lin et al., 2003)(see Materials and Methods). The sev promoter and enhancer drives expression of the transgene specifically in the developing R3, R4, and R7 photoreceptor neurons as well as in the four non-neuronal cone cell precursors (Bowtell et al., 1988; Fortini et al., 1992; Tomlinson and Ready, 1987). Untagged Spen Cterminus (SpenC) does not localize correctly to the nucleus in Drosophila cultured S2 cells and thus an NLS-tag is required to localize the spen C-terminus appropriately (F. Chen and I. Rebay, unpublished results). We observed that sev-spenC-NLS flies have dominant defects and bear eyes that are reduced in size compared to wild-type. sev-spenC-NLS eyes are rough and lack the highly regular pattern of ommatidia, normal hexagonal morphology, and regular spacing of ommatidial bristles observed in wild-type. In addition, a partial transformation of eye tissue into head cuticle is observed by the increased area of hair and cuticle along the ventral-posterior region of the eye (Fig. 42A,B). Some ommatidia also had lens defects that manifest as black holes in the ommatidial lens. Multiple lines that were obtained exhibited similar defects overall but with varying degrees of severity (data not shown). Over-expression of untagged spenC did not result in an observable phenotype (F. Chen and I. Rebay, unpublished results). The rough-eye defects observed in sev-spenC-NLS flies are similar to those seen in flies that have lost spen function in eye tissue (see Chapter 3). In order to understand the patterning defects at cellular resolution, we compared tangential sections of sev-spenC-NLS adult fly eyes 191 192 Fig. 4-2. sev-spenDN functions as a dominant-negative in the eye. Scanning electron micrographs (SEM) for adult eyes of (A) wildtype, (B) sev-spenDN/+, (C) spenAH393/sev-spenDN, and (D) Dp(spen)/+; sev-spenDN/+. Tangential sections for adult eyes of (E) wildtype, (F) spen mutant clones, (G) sev-spenDN/+, (H) spenAH393/sev-spenDN, (I) 2× sev-spenDN, and (J) sev-spenDN; TAF110SY3-1(I53). with those of wild-type and spen loss-of-function. Wild-type adult eyes have regularly-spaced ommatidia, each containing eight trapezoidally-arrayed photoreceptor R cells (only seven of which are observed in any given plane). Each ommatidium is separated by a non-neuronal pigment cell lattice that forms a hexagonally-shaped boundary around the ommatidium (Fig. 42E). To circumvent the embryonic lethality of loss-of-function spen alleles (Chen and Rebay, 2000; Gellon et al., 1997; Kuang et al., 2000; Wiellette et al., 1999), we generated and analyzed somatic mosaic spen clones (see Materials and Methods). spen mutant clones have ommatidia that are disorganized and have lost photoreceptor cell types (Fig. 4-2F; see Chapter 3). In addition to having fewer R cells, sev-spenC-NLS ommatidia have ommatidial fusions, and appear defective in the supporting non-neuronal lattice (Fig. 4-2G). The results above indicate that over-expression of spenC-NLS leads to a phenotype similar to homozygous loss-of-function spen in the eye. Therefore, we tested if sev-spenC-NLS acted as a dominant-negative allele of spen by two independent methods. First, we hypothesized that if sev-spenC-NLS acted as a dominant-negative, a two-fold reduction of endogenous spen would exacerbate the rough-eye phenotype. spen heterozygotes do not show an observable change in eye patterning (data not shown), but flies that express sev-spenC-NLS in this background (Fig. 4-2C,H) have more severe eye defects than compared with the transgene expressed in a wildtype background (Fig. 4-2B,G). sev-spenC-NLS, spen-/+ flies have smaller eyes (Fig. 4-2C) and the ommatidia have fewer cell types (Fig. 4-2H). Second, we hypothesized that a chromosomal duplication that includes the spen locus [Dp(2;1)G146] (Araujo and Bier, 2000) would suppress the activity of sev-spenC-NLS. We found that the spen duplication moderately restored the eye size, ommatidial morphology, and bristle spacing (Fig. 4-2D). In summary, these data suggest that the over-expression of the Spen C-terminus in the eye is dominant-negative. We thus renamed sevspenC-NLS to sev-spenDN. We were interested in determining whether sev-spenDN might provide a suitable background for a genetic interaction screen. First, we confirmed the dosage sensitivity of the sev-spenDN transgene. Two copies of sev-spenDN exacerbates the size reduction, roughness, and 193 loss of cell types of the eye (data not shown; Fig. 4-2I). Heterozygous loss of the gene that encodes the TATA binding protein-associated factor TAF110 promoter-specifically suppresses sev promoter-driven transgenes in a dominant manner (Rebay et al., 2000). We also found that reduction of TAF110 levels also suppressed the rough-eye phenotype of sev-spenDN (Fig. 4-2J). We also asked whether the functional homolog of the co-repressor SMRT, Drosophila Smr, genetically interacted with spen. Since spenDN activity is likely a result of a failure to repress target genes of a Spen repressor complex, reduction of a prospective component of this complex would enhance the spenDN phenotype. In accord with this hypothesis, we found that a P-element insertion mutant of smr enhanced the sev-spenDN rough-eye phenotype (Table 4-2). Potential genetic interactors arise from pilot sev>spenDN modifier screen One way to study the function of a gene is to understand its genetic interactors. As demonstrated above, sev-spenDN provides an ideal system for a genetic screen because the readily observable phenotypes in the eye can be dominantly suppressed or enhanced and are dosagesensitive (Fig. 4-2). We therefore employed screens to identify modifiers of the sev-spenDN rough-eye phenotype (Fig. 4-3). The premise of these screens is that a two-fold reduction in activity of a relevant gene will enhance or suppress the sev-spenDN phenotype. This can be readily observed in the F1 generation (Fig. 4-3). F1 modifier screens such as this are highly efficient and allow for large numbers of flies to be screened (Dickson et al., 1996; Karim et al., 1996; Rebay et al., 2000; Simon et al., 1991; reviewed in St Johnston, 2002). We first performed a pilot screen with spenDN cDNA inserted downstream of 5× Upstream Activating Sequences (UAS) (Fig. 4-1B) and drove expression of the resulting UAS-spenDN transgene with sev-GAL42B driver. sev>spenDN flies have similar eye defects as sev-spenDN (Fig. 4-4B,D). We crossed 256 chromosomal deficiencies from the Bloomington Drosophila Stock Center Deficiency Kit each to sev>spenDN and screened the progeny for enhancement and suppression of the sev>spenDN rough-eye phenotype. We identified 19 moderate-tostrong suppressors and 26 moderate-to-strong enhancers (Table 4-1) that are spread out along 194 Fig. 4-3. Screens for F1 dominant modifiers of the spenDN rough-eye phenotype. Cross schemes for (A) deficiency kit screen with sev>spenDN, (B) P-element insertion screen with sev-spenDN (lethal and EP collection), and (C) X-ray mutagenesis screen with sev-spenDN. Enhancement or suppression of the spenDN rough-eye phenotype is scored in the F­1 generation. 195 196 Fig. 4-4. brm modifies of sev>spenDN . SEM of adults eyes of (A) wild-type and (B) sev>spenDN Tangential sections for adult eyes of (C) wild-type, (D) sev-spenDN/+, (E) sevspenDN/+; brmI21/+. 197 Fig. 4-5. Deficiencies that modify the sev>spenDN rough-eye phenotype. Deficiency regions that suppress (red) or enhance (blue) the rough-eye phenotype along chromosomes X, II, and III. 198 table 4-1A Deficiency stocks that suppress sev>spenDn rough-eye phenotype Deficiency Deficiency Breakpoint Suppression Selected Candidate Suppressors roughest (rst), Notch (N), mnt, flapwing Df()N- 0C02-0;0E0-0 Moderate (flw) Df()C2 0E;09C-D Moderate flw Df(2L)C144 23A01-02;23C03-05 Strong Chd1 Df(2L)J-H 27C02-09;28B03-04 Strong Pvf2, Pvf3, Wnt4, wg Df(2L)XE-2750 28B02; 28D03 Strong Df(2L)Trf-C6R31 28DE Moderate/Strong Df(2L)r10, cn[1] 35D01;36A06-07 Moderate/Strong CycE Df(2L)H20 36A08-09;36E01-02 Moderate grp, CadN, CadN2 Df(2R)Np F0;D09-E0 Moderate ced-6, Wnt2 Df(2R)w-0n A0-07; E02-0 Strong ced-6, Wnt2 Df(2R)k00 C0-0; C0-0 Strong Df(2R)X-2 D0-02;9A Moderate partner of paired (ppa) Df(2R)or-BR 9D0-0;0B0-0 Moderate yorkie (yki) Df(3L)AC1 67A2;67D13 Moderate/Strong Nedd2-like caspase (Nc) Df(3L)XS533 76B04;77B Moderate/Strong Mi-2 Df(3R)P115 89B07-08;89E07-08 Moderate/Strong Df(3R)DG2 89E01-F04;91B01-B02 Moderate/Strong abdA, Abd-B, MESK4, osa Df(R)C 9E0-0;90A0-07 Strong Abd-B, MESK4 Df(R)0 9E0;99A0-0 Moderate string (stg) 199 table 4-1B Deficiency stocks that enhance sev>spenDn rough-eye phenotype Deficiency Deficiency Breakpoint enhancement Selected Candidate enhancers Df()A 0D0-E0;0F0 Moderate mnt Df()BA2- 0F0;0A Moderate Df()C2 07D0;07D0-0 Strong Df()b B0-C0 Moderate Df()B A02;A0 Moderate BarH1 Df(1)BK10 16A02;16C07-10 Moderate BarH1, par-6 Df()N9 7A0;A02 Moderate Pvf1, wengen (wgn) Df(2L)net-PMF 21A01;21B07-08 Strong smo, kis, spen Df(2L)sc19-4 25A05;25E05 Strong tkv Df(2L)TE35BC-24 35B04-06;35F01-07 Moderate Su(H), Snail (Sna) Df(2L)TW161 38A06-B01;40A04-B01 Moderate Fs(2)Ket, e2f2 Df(2R)cn9 2E;C Strong prickle (pk), Hey Df(2R)CX D07-09;7F- Strong lola Df(2R)PC A;F Moderate edl/mae Df(2R)AA2 F09-7;7D-2 Strong shotgun (shg) Df(2R)9AD 9A0-0;9D0-0 Strong Df(3L)vin7 68C08-11;69B04-05 Strong Df(3L)brm11 71F01-04; 72D01-10 Moderate/Strong brm, PDCD-5 Df(3L)st-f13 72C01-D01;73A03-04 Strong brm, PDCD-5 Df(3L)81k19 73A03;74F Strong Keren (Krn) Df(3L)kto2 76B01-02; 76D05 Strong Mi-2, Su(z)12 Df(3L)ME107 77F03;78C08-09 Strong fringe (fng), Polycomb (Pc) Df(R)p-XT0 A02;C0-02 Strong neuralized (neur) Df(R)H-B79 92B0;92F Strong Hairless (H) Df(R)mbc-0 9A0-07;9C0- Strong Df(R)slo 9A02-07; 9D02-0 Strong chromosomes X, II, and III (Fig. 4-5). One caveat to the Deficiency Kit pilot screen is that the deficiencies delete many genes (those with known or unknown function). Listed in Tables 1A and 1B are selected candidate modifiers that have known function and for which allele stocks are available. As proof-of-principle, we identified a deficiency [Df(2L)net-PMF] that deletes the gene region encoding the spen locus as a strong enhancer of sev>spenDN (Table 4-1B). It is formally possible that loss of other genes by this deficiency may also serve as enhancers. Candidate sev-spenDN modifiers deleted by deficiencies are associated with various processes and signaling pathways. Modifying deficiencies delete the gene that encodes the EGFR pathway ligand Keren [Krn; Df(3L)81k19] and the gene that encodes Mae [Df(2R)PC4], a regulator of EGFR pathway transcription factors (Table 4-1B). Components of the Notch signaling pathway are candidate modifiers: Notch [Df(1)N-8], Su(H) [Df(2L)TE35BC-24], and Hairless [Df(3R)p-XT103], and Neuralized [Df(3R)p-XT103] (Table 4-1). This further indicates Spen’s interaction with the EGFR and Notch pathways. We also identified deficiencies that deleted genes that encode cell death regulators: yorkie [Df(2R)or-BR6], Nedd2-like caspase [Df(3L)AC1], and PDCD-5 [Df(3L)st-f13, Df(3L)brm11]. These observations may suggest a role for Spen in programmed cell death (PCD). Two genes deleted by deficiencies that may be of interest are brahma (brm) [Df(3L)st- f13, Df(3L)brm11] and osa [Df(3R)DG2]. Brahma and Osa are components of a Swi/Snf-related Brm chromatin remodeling complex in Drosophila that promotes gene expression (Collins et al., 1999). Interestingly, brm, osa, and spen alleles were identified together in a screen for genetic modifiers of the rough-eye phenotype associated with E2F/DP over-overexpression (StaehlingHampton et al., 1999). We therefore tested loss-of-function or hypomorphic mutations in brahma and osa to determine if specific alleles could modify the sev>spenDN phenotype. The brm2, brmI21, osa2, and osa0090 alleles each dominantly suppressed sev>spenDN and sev-spenDN (Table 4-2). Although the deficiency deleting brm acted as an enhancer of sev>spenDN (Table 41A), we identified it here as a suppressor (Table 4-2; Fig. 4-4E). This indicates another caveat 200 table 4-2 Candidate gene approach for interactors with spenDn rough-eye phenotype gene/Allele Role sev2B>spenDn sev-spenDn sev-yanACt sev-rasv12 chromatin mod/str supp mild/mod supp mild supp mod supp brm2 remodeling chromatin mod/str supp mod supp mild supp mod supp brmI21 remodeling chromatin mod/str supp mild supp mild supp mod enh osa2 remodeling chromatin mod/str supp mild/mod supp mild supp mild/mod enh osa0090 remodeling chromatin mod/str supp mild supp mod/str supp mor1 remodeling txnal mild enh smrBG01648 co-repressor of the Deficiency Kit screen: some deficiencies may delete both enhancers and suppressors. In such cases, loss of one modifier may mask the effect of another modifier. Given the direction of the interactions of brm and osa with sev>spenDN, these observations suggest that Osa and Brm chromatin remodeling may antagonize putative transcriptional repressor function of Spen. Genetic modifiers could specifically affect the sev regulatory elements of the sev-GAL4 or sev-spenDN transgenes. To determine whether or not the brahma and osa alleles suppressed in a promoter-specific manner, we crossed these alleles with sev-yanACT and sev-RasV12, which encode constitutive Yan repressor and activated Ras, respectively (Karim and Rubin, 1998; Rebay et al., 2000). These transgenes act oppositely to each other with respect to EGFR signaling. If brm and osa act in a promoter-specific manner, then the rough-eye phenotypes associated with each of these transgenes would be affected in the same direction. We observed that osa did not have promoter-specific interactions whereas the brm alleles suppressed both sev-yanACT and sev-RasV12 (Table 4-2). These data suggest that brm is a sev promoter-specific suppressor and may not be necessarily interact genetically with spenDN itself. However, given the observation that the Brmcomplex component Osa was identified as an interactor, we may not completely rule-out a role for Brm. In order to determine if other components of the Brm complex interacted with spen, we tested a hypomorphic allele of moira (mor) and found that it also suppressed sev>spenDN 201 table 4-3 P-element insertion stocks that modify sev-spenDn rough-eye phenotype Stock modification Chrom gene EP(2)09 enh 2 CG8929 EP(2)0992 enh 2 EP(2)02 enh 2 CG11395 EP(2)2 enh 2 toutatis molecular function chromatin remodeling EP(2)222 EP(2)227 EP(2)2 EP(2)29 EP(2) EP()0 EP()2 079 enh enh enh enh enh enh enh enh mild 2 2 2 2 2 Misexpression suppressor of ras 3 (MESR3) CG9047 grapes (grp) RluA-1 Cam (calmodulin) 0 enh mild 2 18w (18 wheeler) 0 enh mild 2 par-1 (CG30131) transmembrane receptor activity, cell adhesion molecule activity protein serine/threonine kinase activity, Dl localization 00 enh mild 2 CG32955 or CENP-ana (cana) kinesin motor activity, motor activity, plus-end-directed ATPase activity 0 enh mild 2 prod (proliferation disrupter) 09 enh mild 2 dap (dacapo) 00 09 enh mild enh mild 2 2 lamC (lamin C) and ttv (tout velu) CG8806 22 2 enh mild enh mild 2 2 CG30044 and fdl (fused lobes) l(2)k08611 mitotic chromosome condensation, cell proliferation cyclin dependent protein kinase inhibitor lamC(nuclear envelope reassembly) and ttv (acetylglucosaminyltransferase activity; glucuronosyltransferase activity) px19, PRELI, LEA type protein? fdl(beta-N-acetylhexosaminidase activity) - 0 enh mild l(3)06226 = Psa (Puromycin sensitive aminopeptidase) [EP(3)3486] cytosol alanyl aminopeptidase 0 097 enh mild enh mild l(3)L6750 = frc (fringe connection) Skuld (Skd) pyrimidine nucleotide sugar transporter (N receptor processing) RNAPII transcription mediator 0290 0 0 enh mild enh mild enh mild MRG15 (MORF-related Gene 15) Aats-Ile (Isoleucyl tRNA synthetase) Aats-Gln (Glutaminyl tRNA synthetase) chromo domain containing protein (transcriptional activation) isoleucine tRNA ligase glutamine tRNA ligase 9 enh mild sar1 7 7 7 20 2077 2097 20 20 22 enh mild enh mild enh mild enh mild enh mild enh mild enh mild enh mild enh mild l(3)07207 = crumbs [EP(3)1089] put (punt) Int6 Hsp83 between CG3967 and Shc CG32159 CG13062 Mi-2 CG3953 202 protein ser/thr kinase RNA processing calcium ion binding activity RAS small monomeric GTPase (component of COPII coated vesicle) cell polarity establishment, gammasecretase regulation for N signaling protein kinase (TGFbeta pathway) translation initiation factor chaperone Shc (SH2, PID, PH adaptor protein) ATP dependent DNA helicase metalloendopeptidase table 4-3 P-element insertion stocks that modify sev-spenDn rough-eye phenotype Stock modification Chrom gene 27 enh mild l(3)rJ880 27 enh mild awd (abnormal wing discs) 09 enh mild Ten-m (tenascin major) 2 enh mild gish (gilgamesh) molecular function nucleoside diphosphate kinase cell adhesion casein kinase I enh mild CG6685 and Nc (Nedd-2 like caspase) Nc(signaling (initiator) caspase/caspase-2/caspase) 0 0 enh mod enh mod 2 2 Star (S) Vha68-2 (CG4385) EGF receptor protein processing, protein-Golgi trafficking hydrogen exporting ATPase activity 070 enh mod 2 l(2)k08002 = 18w (CG8896 ~ tlr) 0 092 enh mod enh mod 2 2 l(2)k08915 (EP2237, CG4427) l(2)k10213 (=mir-14) 27 7 72 enh mod enh mod enh mod 2 2 CG8674 CG5953 pointed (pnt) 0 0 supp mild supp mild supp mild 2 2 Ark (Apaf-1 related killer) ena (enabled) and CG15118 bantam (ban) 079 supp mod 2 Ggamma1 protein binding activity, caspase activator activity actin binding activity miRNA, cell proliferation, growth Heterotrimeric G-protein GTPase activity 09 supp mod 2 l(2)k09837 = CG15609 actin binding activity, calcium ion binding activity, CH-domain cell adhesion/transmembrane receptor, toll-receptor like protein transcriptional activator, similar to TGFbeta inducible growth response protein 2 (TIEG-2) or kruppel-like factor suppresses cell death, similar to ATP synthetase mitochondrial F1 complex assembly factor 2 Ets transcription factor (Table 4-2). We cannot conclude at this point whether this is a promoter-specific interaction. In summary, our observations suggest that the Brm chromatin-remodeling complex may act to oppose Spen function. A P-element screen for modifiers of sev-spenDN The pilot screen demonstrated that genetic modifiers of sev>spenDN could be identified. Therefore, we proceeded with a scaled-up screen using gene-disrupting P-element insertional mutants (reviewed in Spradling et al., 1995). This screen did not use the GAL4>UAS system to drive expression of sev>spenDN but instead used the sev regulatory elements directly to drive over-expression of the spenDN transgene. This circumvents issues related to modification of the GAL4>UAS system itself and simplifies the genetics. We utilized two collections of P-element 203 insertional mutant collections: lethal P-element (or ‘P-lethal’) lines (Spradling et al., 1999) and Enhancer-Promoter (EP) lines (Rorth, 1996). The P-lethals are homozygous lethal and thus may represent loss-of-function mutations for essential genes (Spradling et al., 1999). The EP lines are P-elements that contain UAS sequences and may be used in over-expression screens (Rorth, 1996). In this case, we used the EP collection solely as insertional mutations. The attractive nature of these P-element collections is that the location of each insertion is known and the affected gene has been identified (FlyBase). For the following results, it is important to note that the genetic modifiers have not been yet been tested for promoter-specific interactions with sev regulatory elements as we had performed above and for the X-ray modifier screen (described in the next section). We crossed P-element insertion mutants from chromosomes II and III with sev-spenDN and looked for genetic modifiers in their progeny (Fig. 4-3B). From the EP collection, after screening 1,717 insertions, 11 enhancers and no suppressors were identified (Table 4-3). The enhancers include toutatis (tou) and grapes (grp) which encodes an Iswi-associated chromatin remodeling protein and a cell cycle checkpoint serine/threonine kinase, respectively (Fogarty et al., 1997; Vanolst et al., 2005). We screened 1,056 P-lethal insertions (666 Chr. II and 390 Chr. III) and identified 41 modifiers: 39 enhancers and 2 suppressors (Table 4-3). The enhancers include Star (S) and pointed (pnt), which encode positive components of the EGFR pathway that process EGFR ligand and activate target genes, respectively (Table 4-3)(reviewed in Doroquez and Rebay, 2006). We identified two cell polarity-related enhancers par-1 and crumbs involved in Notch signaling (Table 4-1). Par-1 is a protein serine/threonine kinase Par-1, which regulates proper localization of the Notch ligand Delta (Bayraktar et al., 2006). Crumbs is a transmembrane protein that limits γ-secretase’s activity for producing activated, intracellular Notch (Herranz et al., 2006). A number of cell death-related genes were identified as modifiers in the P-lethal screen. Nedd-2 like caspase (Nc) and the caspase activator Apaf-1 related killer (dArk) (Rodriguez et al., 1999)were identified as an enhancer and a suppressor, respectively (Table 4-3). We also 204 identified miR-14 and bantam (ban), which encode microRNAs that inhibit cell death (Table 4-3) (Brennecke et al., 2003; Xu et al., 2003). The result that we identified a number of apoptosisrelated genes suggests that spen may play a role in regulating cell death (see Appendix II). Another enhancer further suggests Spen’s role in regulating chromatin state. dMi-2 encodes an ATP-dependent DNA helicase involved in chromatin remodeling (Table 4-3)(Brehm et al., 2000). The genetic interaction observed here suggests a similar role for Spen and dMi-2 in chromatin regulation. X-ray screen identifies modifiers including new alleles of spen and pointed The P-lethal collection disrupts about 25 percent of the estimated 3600 genes on the autosomes required for viability (Spradling et al., 1999). Therefore, our screen for modifiers was not likely saturated. In order to identify new mutations that modify sev-spenDN, we performed a mutagenesis screen. We opted to use double-strand break-inducing X-rays as a mutagen in order to circumvent the male mosaicism associated with EMS-mutagenesis. Mosaicism arises when EMS mutated post-meiotic sperm undergo DNA repair in the embryo (reviewed in St Johnston, 2002). We mutagenized a total of 41,300 males that we crossed to an equivalent number of sevspenDN virgin females (Fig. 4-3C). We screened a total of 105,578 F1 flies (56,862 males; 48,716 females). We obtained a total of 60 modifiers: 49 enhancers and 11 suppressors (Table 4-4A; see Materials and Methods). We determined homozygous viability and chromosome linkage for each modifier: 31 modifiers on chromosome II, 26 modifiers on chromosome III, and one on the Y chromosome (Table 4-4A). We organized the homozygous lethal modifiers according to complementation groups. We also tested for promoter-specific interactions with sev-yanACT and sev-RasV12, as described above. These modifiers are termed Esdn and Ssdn for Enhancer of spenDN and suppressor of spenDN, respectively. Three enhancer complementation groups were identified for chromosome II (Table 4-4B). Five new alleles of spen were isolated, providing proof-of-principle of the screen. The two other Esdn2 complementation groups had two and three alleles, respectively. 205 These alleles and the single hits have not been mapped to a specific locus on chromosome II. Two of the single hits, KH2-331 and JL3-335, had promoter-specific interactions and were not pursued. On chromosome III, six new alleles of the enhancer pnt were isolated but the other enhancers did not fall into complementation groups. No complementation groups were identified for suppressors on chromosome II. The two chromosome III suppressors belong to a single complementation group that had promoter-specific interactions (Table 4-4B). We also determined if the modifiers had eye phenotypes in a wild-type background. Six modifiers had dominant rough-eye phenotypes. Three viable modifiers showed recessive rough-eye phenotypes (Table 4-4B). In summary, the X-ray screen identified new alleles of spen and pnt, while the other modifiers remain to be mapped and attributed to a gene. Discussion In this study, we demonstrated that over-expression of the C-terminus of Spen acts as a dominant-negative. Using SpenDN as a tool to identify genes related to Spen function we identify candidates that may elucidate the processes in which Spen functions during eye development. Eye-specific over-expression of the Spen C-terminus is dominant-negative We previously found that CNS midline glia cell (MGC)-specific expression of Spen∆C acted as a dominant-negative (Chen and Rebay, 2000). Similarly, over-expression of SpenDN in the MGCs produces a comparable phenotype suggesting it too functions as a dominant-negative in both eye and embryonic CNS development (see Appendix III). Further evidence of SpenDN’s activity as a dominant-negative protein is reported by Lin and colleagues (2003), who identified spen as a positive regulator of Wingless (Wg) signaling during wing development (Lin et al., 2003). 206 207 6 7 31 Suppressors lethal viable totals 2 2 0 26 iii 24 1 0 1 y 0 0 0 11 9 2 totals 49 34 15 Compl Chrom group Chromosome II Enhancers 2 spen 2 spen 2 spen 2 spen 2 spen 2 Esdn2-1 2 Esdn2-1 2 Esdn2-2 2 Esdn2-2 2 Esdn2-2 2 Esdn2 single 2 Esdn2 single 2 Esdn2 single 2 Esdn2 single 2 Esdn2 single 2 Esdn2 single 2 Esdn2 single 2 Esdn2 single modification Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Strong Enh - Mild Enh - Mod Enh - Mod Enh - Mod Enh - Strong Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Strong Enh - Mild Allele DV2-204 G2-38 MA2-357 NQ2-401 HH2-260 IP1-299 HC3-266 AI4-113 BN1-135 MX1-375 E1-22 FN1-227 IB1-284 IQ3-304 KE1-308 KW4-337 KH2-331 JL3-335 Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal lethal? No No No No No No No No No Yes No No No No No Yes Yes No eye-het? eyehom? N/A N/A N/A N/A N/A N/A No No No N/A N/A N/A N/A N/A N/A N/A N/A N/A table 4-4B x-ray-induced modifiers of sev-spenDn rough-eye phenotype ii 25 7 modification enhancers lethal viable table 4-4A x-ray screen for modifiers of sevspenDn rough-eye phenotype mild supp mod supp mild supp mild supp mod supp mild supp No. mod supp mild supp No. mod enh mod enh mild supp sev-rasv12 mild enh mild enh strong enh mod enh strong enh No. mild enh No. No. No. mod enh mod enh No. sev-yanACt No. No. No. No. No. No. No. No. No. No. Yes. Yes. No. Promoter Specific? 208 Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mod Enh - Mild Enh - Mod Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild modification Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mod Chromosome III Enhancers pnt IK1-298 pnt IM4-318 pnt LK2-344 pnt NL2-385 pnt Q1-44 pnt DR2-194 Esdn3 single BC1-123 Esdn3 single P4-76 Esdn3 single BI1-129 Esdn3 single BW1-141 Esdn3 single DC1-184 Esdn3 single MO2-368 Esdn3 single MU3-370 Allele BB1-118 EI3-222 HB1-252 IH2- LY1-354 ML2-364 O2- Supp - Mild Supp - Mild Supp - Mild Supp - Mild Supp - Mild Supp - Mild Supp - Mild Supp - Mild Compl group Esdn2 viable Esdn2 viable Esdn2 viable Esdn2 viable Esdn2 viable Esdn2 viable Esdn2 viable Chromosome II Suppressors 2 Ssdn2 single A2-6 2 Ssdn2 single EG2-208 2 Ssdn2 single EJ3-216 2 Ssdn2 single HD3-282 2 Ssdn2 single LI1-330 2 Ssdn2 single MS1-367 2 Ssdn2 single NI1-383 2 Ssdn2 viable GE2-20 Chrom 2 2 2 2 2 2 2 Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Lethal Viable lethal? Viable Viable Viable Viable Viable Viable Viable Yes No No No No No Yes No No No - No No No No No No No No eye-het? No No Yes No No No No N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A No eyehom? N/A N/A Yes No Yes Yes No table 4-4B x-ray-induced modifiers of sev-spenDn rough-eye phenotype mod supp No. No. mild supp mild supp No. No. No. enh - sev-rasv12 - mild enh No. No. No. mild enh No. No. No. mild supp - sev-yanACt - OK OK OK Yes. No. No. No. No. - Promoter Specific? - 209 Supp - Mild modification Enh - Mild Enh - Mod Enh - Mod Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mild Enh - Mod Enh - Mod Y Chromosome Suppressor Y Ssdny-1 DZ3-225 Allele NY1-389 HP1-278 MS3-369 CT2-7 GF-27 HT-2 J1-23 L3-43 LH3-393 CB3-163 I- Supp - Mild Supp - Mod Compl group Esdn3 single Esdn3 single Esdn3 single Esdn3 viable Esdn3 viable Esdn3 viable Esdn3 viable Esdn3 viable Esdn3 viable Esdn3 viable Esdn3 viable Chromosome III Suppressors Ssdn3-1 AH3-111 Ssdn3-1 AL2-89 Chrom Viable Lethal Lethal lethal? Lethal Lethal Lethal Viable Viable Viable Viable Viable Viable Viable Viable - - eye-het? No No No - - N/A N/A eyehom? N/A N/A N/A No No No No No Yes No table 4-4B x-ray-induced modifiers of sev-spenDn rough-eye phenotype - mod supp mild supp sev-rasv12 No. mild supp No. - - No. mild supp sev-yanACt No. No. No. - - Yes. Yes. Promoter Specific? OK OK OK - How do Spen∆C and SpenDN act as dominant-negatives at the molecular level? One model suggests that Spen-like proteins act as transcription factors bound to DNA via their RRMs (Sierra et al., 2004). Since Spen∆C proteins lack the repressive SPOC domain, Spen∆C fails to elicit transcriptional repression and simultaneously blocks endogenous Spen proteins from binding DNA. SpenDN proteins may function by titrating factors, such as co-repressors, that are bound by the SPOC domain away from endogenous Spen proteins. Another one of these titrated proteins may be endogenous Spen itself. The SPOC domain for the murine ortholog Mint was shown to homodimerize. In cultured cells, over-expression of the Mint SPOC fragment abrogates the transcriptional repression mediated by full-length Mint (Sierra et al., 2004). Thus homodimerization between the SPOCs of SpenDN and endogenous Spen may attenuate normal function. We found that eye-specific over-expression of SpenDN may not completely recapitulate spen loss-of-function. Expression of the SpenDN with the GMR- or sev-GAL4 drivers does not lead to an observable increase in Yan protein levels as is observed in spen loss-of-function. We did find that GMR>spen∆C had elevated Yan protein levels, but did not demonstrate a discernible change in eye phenotype at the dissecting scope level (see Appendix V). MGCdriven Spen∆C also does not have great penetrance in inducing a dominant-negative phenotype in a wild-type background (Chen and Rebay, 2000). Therefore, spen∆C and spenDN transgenes may each represent dominant-negative spen ‘hypomorphs’. It would be interesting to determine if simultaneous expression of these transgenes leads to a complete dominant-negative spen ‘amorph’. EGFR and Notch pathway components interact genetically with dominant-negative spen We have performed genetic screens each with the aim of identifying interactors of spen. We hypothesized that modifiers identified in these screens could interact with Spen at the molecular level or be involved in cellular processes that also require Spen. As a proof-ofprinciple, our x-ray modifier screen identified several alleles of the Ets-domain transcriptional 210 activator pnt. This underscores the observation that Spen acts closely with downstream components of the EGFR signaling pathway (Chen and Rebay, 2000)(see Chapter 2). Although a number of EGFR pathway components may be represented in the unidentified modifiers of our X-ray screen, we did not identify many EGFR pathway components in the P-element insertion screens. Our stock-blind screening method is predicted to identify EGFR pathway mutants represented in the P-lethal collection, including Egfr, sos, drk, Ras85D. We also identified few modifiers that relate to Notch signaling: crumbs and par-1. However, depending on where spen intersects with the EGFR pathway, our ability to get hits in upstream components of the pathway may be limited. To confirm interactions with genes in the EGFR and Notch pathway genetic crosses with specific alleles of EGFR and Notch components should be performed to determination modification of the spenDN phenotype. Spen may regulate chromatin state Modifiers from our screen suggest a role for Spen is chromatin regulation. A candidate approach identified the SMRT functional ortholog Smr likely be relevant to Spen function. Smr may be involved in a co-repressor complex with Spen. Crystal structure analysis indicates that the basic patch of the Spen SPOC domain is likely to bind directly to an acidic surface of Smr that is conserved among SMRT-like co-repressors (Ariyoshi and Schwabe, 2003). In addition to Smr, our results suggest that Mi-2 may be a member of Spen repressor complex. Interestingly, the human Spen ortholog Sharp was found to interact in a complex with Mi-2 in the NuRD repressor complex (Shi et al., 2001). Are other components shared between the Sharp and Spen co-repressor complexes? Attempts to answer this question genetically have been inconclusive. A loss-of-function allele for Rpd3, the Drosophila HDAC1, did not modify the rough-eye phenotype associated with sev-spenDN or spen mutant clones. Furthermore, a P-element insertion allele for the gene that encodes the Drosophila homolog of the NuRD core component MBD3, MBD-like, did not show modification of sev-spen or spen clones. However, the nature of this allele has not been characterized (see Appendix III). In order to rigorously test interaction of a 211 Spen repressor complex, experiments using NLS- and epitope-tagged Spen C-terminus should be performed to determine if Smr and NuRD complex components co-immunoprecipitate. Our screen also suggests a genetic interaction of spen with the Brahma chromatinremodeling complex that includes Brm, Osa, and Mor. This complex may include other chromatin-remodeling related proteins identified as interactors in the screen, including the Skd mediator, the nucleolar remodeling complex (NoRD)-associated protein Tou, and the chromatin-associated transcriptional activator MRG15. The direction of interaction for Brm complex mutants with spenDN indicates that the Brm complex interacts opposite to Spen function. This is consistent with the observation that the Brm complex facilitates the activation of gene transcription (Armstrong et al., 2002). Interestingly, the Brm complex was also found to act positively with the Notch signaling pathway during Drosophila sensory organ precursor development (Armstrong et al., 2005). In vertebrates, human Brm binds to CBF-1, a vertebrate Su(H) homolog to recruit the Brm complex to target genes activated by the Notch pathway (Kadam and Emerson, 2003). These insights therefore propose a role for Spen in the repression Notch pathway target genes whose activation is mediated Brm chromatin remodeling. Spen may be a regulator of cell death Our screens suggest a significant role for Spen in the regulation of cell death. At first glance, these results may not be a surprising, because first, the EGFR pathway is required for cell survival during eye development (reviewed in McNeill and Downward, 1999). MAPK phosphorylation limits the pro-apoptotic protein Hid from inhibiting Diap1, an inhibitor of caspases (Berger, 2001; Kurada and White, 1998; Wang et al., 1999). Second, loss of cell types in the adult eye suggests elevated apoptosis in loss-of-function spen mutant clones (see Chapter 3). Preliminary data indicates that more activated Caspase-3 cells may be present in spen mutant clones (see Appendix II). Third, the rough-eye phenotype of sev>spenDN is suppressible with the inhibitor of apoptosis P35 (data not shown)(Clem et al., 1991). 212 Considering the protective role for EGFR against PCD, our sev-spenDN screen would have predicted mutants in pro-apoptotic genes as suppressors and mutants in anti-apoptotic genes as enhancers. This was the case for the pro-apoptotic gene dArk and anti-apoptotic miR14. However, opposite of our prediction, we found that anti-apoptotic ban and pro-apoptotic Drosophila Nedd2 (Nc) were a suppressor and enhancer, respectively. Does this indicate that Spen has both anti- and pro-apoptotic roles? Interestingly, the SPOC domain shows some sequence and structural similarity to proteins of the Death inducer-obligator (DIO) family (Sanchez-Irizarry et al., 2004). DIO proteins are developmentally-regulated transcription factors upregulated during apoptosis. Ectopic expression of DIO proteins promotes cell death (GarciaDomingo et al., 2003). These insights suggest diverse roles for Spen in the regulation of PCD during Drosophila development. Conclusions and state of the sev-spenDN screen The screens we have performed should be considered pilot experiments demonstrating the suitability of spenDN as a background in which to identify genetic interactors. Although the X-ray screen did identify multiple alleles of spen and pnt, these loci may be hotspots for mutagenesis (Rebay, 2002) and thus not good indicators of saturation. The relatively few modifiers identified in our x-ray screen (~60 from over ~100,000 F1 animals screened) may be a result of excessively harsh mutagenesis. The amount of irradiation could be reduced (from 5000 R to 4000 R), as used in similar screens (Rebay et al., 2000). Furthermore, we did not screen through the complete collection of transposon insertions. Thus to expand the list of interactors, future endeavors should complete screening of the remaining one-third of the P-lethal lines, as well as other insertional collections using transposons such as PiggyBac that do not have the same locational biases as P-elements (reviewed in Spradling et al., 1995). Smaller deletions generated in an isogenic background, such as the DrosDel collection, may also prove to be useful in identifying modifiers. In conclusion, identifying genetic modifiers of spen provides an 213 important starting point for understanding Spen function, with the nature of the encoded gene product directing further molecular analyses. The power of molecular genetic analysis in the fruitfly makes it an ideal model system in which to analyze Spen function during development. Acknowledgements We would like to thank Paul Garrity, Terry Orr-Weaver, and the Rebay and Orr-Weaver labs for stimulating discussion. We are grateful to Fangli Chen for mentorship, guidance, and the P[sev-spenDN] construct, Laura Doyon for transgene injection, Eric Henckels (Tufts University) and Emily Ludwig (Williams College) for assisting in the genetic screen, Jessica Whited (Garrity Lab, MIT) and Kevin Cook (Bloomington Drosophila Stock Center) for sharing the P-element lethal stocks, and Nicki Watson (WIBR/Keck Microscopy Facility) for microscopy assistance. A National Science Foundation Graduate Fellowship supported D.B.D. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. References Araujo, H., and Bier, E. (2000). sog and dpp exert opposing maternal functions to modify toll signaling and pattern the dorsoventral axis of the Drosophila embryo. Development 127, 3631-3644. Ariyoshi, M., and Schwabe, J. W. (2003). A conserved structural motif reveals the essential transcriptional repression function of Spen proteins and their role in developmental signaling. Genes Dev 17, 1909-1920. Armstrong, J. A., Papoulas, O., Daubresse, G., Sperling, A. S., Lis, J. T., Scott, M. P., and Tamkun, J. W. (2002). The Drosophila BRM complex facilitates global transcription by RNA polymerase II. Embo J 21, 5245-5254. Armstrong, J. A., Sperling, A. S., Deuring, R., Manning, L., Moseley, S. L., Papoulas, O., Piatek, C. I., Doe, C. Q., and Tamkun, J. W. (2005). Genetic Screens for Enhancers of brahma Reveal Functional Interactions Between the BRM Chromatin-Remodeling Complex and the Delta-Notch Signal Transduction Pathway in Drosophila. Genetics 170, 1761-1774. Bayraktar, J., Zygmunt, D., and Carthew, R. W. (2006). Par-1 kinase establishes cell polarity and 214 functions in Notch signaling in the Drosophila embryo. J Cell Sci 119, 711-721. Berger, S. L. (2001). An embarrassment of niches: the many covalent modifications of histones in transcriptional regulation. Oncogene 20, 3007-3013. Bowtell, D. D., Simon, M. A., and Rubin, G. M. (1988). Nucleotide sequence and structure of the sevenless gene of Drosophila melanogaster. Genes Dev 2, 620-634. Brehm, A., Langst, G., Kehle, J., Clapier, C. R., Imhof, A., Eberharter, A., Muller, J., and Becker, P. B. (2000). dMi-2 and ISWI chromatin remodelling factors have distinct nucleosome binding and mobilization properties. Embo J 19, 4332-4341. Brennecke, J., Hipfner, D. R., Stark, A., Russell, R. B., and Cohen, S. M. (2003). bantam encodes a developmentally regulated microRNA that controls cell proliferation and regulates the proapoptotic gene hid in Drosophila. Cell 113, 25-36. Chen, C. K., and Chien, C. T. (1999). Negative regulation of atonal in proneural cluster formation of Drosophila R8 photoreceptors. Proc Natl Acad Sci U S A 96, 5055-5060. Chen, F., and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10, 943-946. Clem, R. J., Fechheimer, M., and Miller, L. K. (1991). Prevention of apoptosis by a baculovirus gene during infection of insect cells. Science 254, 1388-1390. Collins, R. T., Furukawa, T., Tanese, N., and Treisman, J. E. (1999). Osa associates with the Brahma chromatin remodeling complex and promotes the activation of some target genes. Embo J 18, 7029-7040. Dickson, B. J., van der Straten, A., Dominguez, M., and Hafen, E. (1996). Mutations Modulating Raf signaling in Drosophila eye development. Genetics 142, 163-171. Doroquez, D. B., and Rebay, I. (2006). Signal Integration During Development: Mechanisms of EGFR and Notch Pathway Function and Cross-Talk. Crit Rev Biochem Mol Biol 41, 339385. Fogarty, P., Campbell, S. D., Abu-Shumays, R., Phalle, B. S., Yu, K. R., Uy, G. L., Goldberg, M. L., and Sullivan, W. (1997). The Drosophila grapes gene is related to checkpoint gene chk1/ rad27 and is required for late syncytial division fidelity. Curr Biol 7, 418-426. Fortini, M. E., Simon, M. A., and Rubin, G. M. (1992). Signalling by the sevenless protein tyrosine kinase is mimicked by Ras1 activation. Nature 355, 559-561. Garcia-Domingo, D., Ramirez, D., Gonzalez de Buitrago, G., and Martinez, A. C. (2003). Death inducer-obliterator 1 triggers apoptosis after nuclear translocation and caspase upregulation. Mol Cell Biol 23, 3216-3225. Gellon, G., Harding, K. W., McGinnis, N., Martin, M. M., and McGinnis, W. (1997). A genetic screen for modifiers of Deformed homeotic function identifies novel genes required for head development. Development 124, 3321-3331. Herranz, H., Stamataki, E., Feiguin, F., and Milan, M. (2006). Self-refinement of Notch activity through the transmembrane protein Crumbs: modulation of gamma-secretase activity. EMBO Rep 7, 297-302. Hurlbut, G. D., Kankel, M. W., Lake, R. J., and Artavanis-Tsakonas, S. (2007). Crossing paths with Notch in the hyper-network. Curr Opin Cell Biol 19, 166-175. Kadam, S., and Emerson, B. M. (2003). Transcriptional specificity of human SWI/SNF BRG1 and BRM chromatin remodeling complexes. Mol Cell 11, 377-389. Karim, F. D., Chang, H. C., Therrien, M., Wassarman, D. A., Laverty, T., and Rubin, G. M. (1996). A screen for genes that function downstream of Ras1 during Drosophila eye 215 development. Genetics 143, 315-329. Karim, F. D., and Rubin, G. M. (1998). Ectopic expression of activated Ras1 induces hyperplastic growth and increased cell death in Drosophila imaginal tissues. Development 125, 1-9. Kuang, B., Wu, S. C., Shin, Y., Luo, L., and Kolodziej, P. (2000). split ends encodes large nuclear proteins that regulate neuronal cell fate and axon extension in the Drosophila embryo. Development 127, 1517-1529. Kurada, P., and White, K. (1998). Ras promotes cell survival in Drosophila by downregulating hid expression. Cell 95, 319-329. Kuroda, K., Han, H., Tani, S., Tanigaki, K., Tun, T., Furukawa, T., Taniguchi, Y., Kurooka, H., Hamada, Y., Toyokuni, S., and Honjo, T. (2003). Regulation of Marginal Zone B Cell Development by MINT, a Suppressor of Notch/RBP-J Signaling Pathway. Immunity 18, 301312. Lai, E. C. (2002). Keeping a good pathway down: transcriptional repression of Notch pathway target genes by CSL proteins. EMBO Rep 3, 840-845. Lee, J. W., Lee, Y. C., Na, S. Y., Jung, D. J., and Lee, S. K. (2001). Transcriptional coregulators of the nuclear receptor superfamily: coactivators and corepressors. Cell Mol Life Sci 58, 289297. Li, J., Yang, X., Qin, H., Zhou, P., Liang, Y., and Han, H. (2005). The C terminus of MINT forms homodimers and abrogates MINT-mediated transcriptional repression. Biochim Biophys Acta 1729, 50-56. Lin, H. V., Doroquez, D. B., Cho, S., Chen, F., Rebay, I., and Cadigan, K. M. (2003). Splits ends is a tissue/promoter specific regulator of Wingless signaling. Development 130, 3125-3135. Ludewig, A. H., Kober-Eisermann, C., Weitzel, C., Bethke, A., Neubert, K., Gerisch, B., Hutter, H., and Antebi, A. (2004). A novel nuclear receptor/coregulator complex controls C. elegans lipid metabolism, larval development, and aging. Genes Dev 18, 2120-2133. Ma, X., Renda, M. J., Wang, L., Cheng, E. C., Niu, C., Morris, S. W., Chi, A. S., and Krause, D. S. (2007). Rbm15 Modulates Notch-induced Transcriptional Activation and Affects Myeloid Differentiation. Mol Cell Biol. McNeill, H., and Downward, J. (1999). Apoptosis: Ras to the rescue in the fly eye. Curr Biol 9, R176-179. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W., et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J 21, 5417-5426. Oswald, F., Winkler, M., Cao, Y., Astrahantseff, K., Bourteele, S., Knochel, W., and Borggrefe, T. (2005). RBP-Jkappa/SHARP recruits CtIP/CtBP corepressors to silence Notch target genes. Mol Cell Biol 25, 10379-10390. Ready, D. F., Hanson, T. E., and Benzer, S. (1976). Development of the Drosophila retina, a neurocrystalline lattice. Dev Biol 53, 217-240. Rebay, I. (2002). Keeping the receptor tyrosine kinase signaling pathway in check: lessons from Drosophila. Rebay, I., Chen, F., Hsiao, F., Kolodziej, P. A., Kuang, B. H., Laverty, T., Suh, C., Voas, M., Williams, A., and Rubin, G. M. (2000). A genetic screen for novel components of the Ras/Mitogen-activated protein kinase signaling pathway that interact with the yan gene of Drosophila identifies split ends, a new RNA recognition motif- containing protein. Genetics 216 154, 695-712. Rodriguez, A., Oliver, H., Zou, H., Chen, P., Wang, X., and Abrams, J. M. (1999). Dark is a Drosophila homologue of Apaf-1/CED-4 and functions in an evolutionarily conserved death pathway. Nat Cell Biol 1, 272-279. Rorth, P. (1996). A modular misexpression screen in Drosophila detecting tissue-specific phenotypes. Proc Natl Acad Sci U S A 93, 12418-12422. Sanchez-Irizarry, C., Carpenter, A. C., Weng, A. P., Pear, W. S., Aster, J. C., and Blacklow, S. C. (2004). Notch subunit heterodimerization and prevention of ligand-independent proteolytic activation depend, respectively, on a novel domain and the LNR repeats. Mol Cell Biol 24, 9265-9273. Shi, Y. (2002). Mechanisms of caspase activation and inhibition during apoptosis. Mol Cell 9, 459-470. Shi, Y., Downes, M., Xie, W., Kao, H. Y., Ordentlich, P., Tsai, C. C., Hon, M., and Evans, R. M. (2001). Sharp, an inducible cofactor that integrates nuclear receptor repression and activation. Genes Dev 15, 1140-1151. Sierra, O. L., Cheng, S. L., Loewy, A. P., Charlton-Kachigian, N., and Towler, D. A. (2004). MINT, the Msx2 interacting nuclear matrix target, enhances Runx2-dependent activation of the osteocalcin fibroblast growth factor response element. J Biol Chem 279, 32913-32923. Simon, M. A., Bowtell, D. D., Dodson, G. S., Laverty, T. R., and Rubin, G. M. (1991). Ras1 and a putative guanine nucleotide exchange factor perform crucial steps in signaling by the sevenless protein tyrosine kinase. Cell 67, 701-716. Spradling, A. C., Stern, D., Beaton, A., Rhem, E. J., Laverty, T., Mozden, N., Misra, S., and Rubin, G. M. (1999). The Berkeley Drosophila Genome Project gene disruption project: Single P-element insertions mutating 25% of vital Drosophila genes. Genetics 153, 135-177. Spradling, A. C., Stern, D. M., Kiss, I., Roote, J., Laverty, T., and Rubin, G. M. (1995). Gene disruptions using P transposable elements: an integral component of the Drosophila genome project. Proc Natl Acad Sci U S A 92, 10824-10830. St Johnston, D. (2002). The art and design of genetic screens: Drosophila melanogaster. Nat Rev Genet 3, 176-188. Staehling-Hampton, K., Ciampa, P. J., Brook, A., and Dyson, N. (1999). A genetic screen for modifiers of E2F in Drosophila melanogaster. Genetics 153, 275-287. Sundaram, M. V. (2005). The love-hate relationship between Ras and Notch. Genes Dev 19, 1825-1839. Tomlinson, A., and Ready, D. F. (1987). Neuronal differentiation in the Drosophila ommatidium. Developmental Biology 120, 366-376. Tootle, T. L., Lee, P. S., and Rebay, I. (2003). CRM1-mediated nuclear export and regulated activity of the Receptor Tyrosine Kinase antagonist YAN require specific interactions with MAE. Development 130, 845-857. Tsai, C. C., Kao, H. Y., Yao, T. P., McKeown, M., and Evans, R. M. (1999). SMRTER, a Drosophila nuclear receptor coregulator, reveals that EcR- mediated repression is critical for development. Mol Cell 4, 175-186. Tsuji, M., Shinkura, R., Kuroda, K., Yabe, D., and Honjo, T. (2007). Msx2-interacting nuclear target protein (Mint) deficiency reveals negative regulation of early thymocyte differentiation by Notch/RBP-J signaling. Proc Natl Acad Sci U S A 104, 1610-1615. Vadlamudi, R. K., Manavathi, B., Singh, R. R., Nguyen, D., Li, F., and Kumar, R. (2005). An 217 essential role of Pak1 phosphorylation of SHARP in Notch signaling. Oncogene 24, 45914596. Vanolst, L., Fromental-Ramain, C., and Ramain, P. (2005). Toutatis, a TIP5-related protein, positively regulates Pannier function during Drosophila neural development. Development 132, 4327-4338. Wang, S. L., Hawkins, C. J., Yoo, S. J., Muller, H. A., and Hay, B. A. (1999). The Drosophila caspase inhibitor DIAP1 is essential for cell survival and is negatively regulated by HID. Cell 98, 453-463. Whitmarsh, A. J. (2006). Regulation of gene transcription by mitogen-activated protein kinase signaling pathways. Biochim Biophys Acta. Wiellette, E. L., Harding, K. W., Mace, K. A., Ronshaugen, M. R., Wang, F. Y., and McGinnis, W. (1999). spen encodes an RNP motif protein that interacts with Hox pathways to repress the development of head-like sclerites in the Drosophila trunk. Development 126, 5373-5385. Wotton, D., and Massague, J. (2001). Smad transcriptional corepressors in TGF beta family signaling. Curr Top Microbiol Immunol 254, 145-164. Xu, P., Vernooy, S. Y., Guo, M., and Hay, B. A. (2003). The Drosophila MicroRNA Mir-14 Suppresses Cell Death and Is Required for Normal Fat Metabolism. Curr Biol 13, 790-795. Yang, X., Li, J., Qin, H., Yang, H., Zhou, P., Liang, Y., and Han, H. (2005). Mint represses transactivation of the type II collagen gene enhancer through interaction with alpha Acrystallin-binding protein 1. J Biol Chem 280, 18710-18716. 218 Chapter V. Discussion and Future Directions David B. Doroquez and Ilaria Rebay 219 P roper patterning of the Drosophila eye requires a set of signaling pathways that are used reiteratively during the course of retinal development. The work described in this thesis advances the study of Drosophila eye development through the genetic analysis of split ends (spen). Spen is critical for regulating two signaling modules required for the eye: the Epidermal Growth Factor Receptor (EGFR) and Notch pathways. During morphogenetic furrow (MF) progression, Spen antagonizes Notch signaling via inhibition of scabrous (sca), whose gene product opposes an endosomal degradation pathway for the Notch receptor (Li et al., 2006). Spen also regulates activated Mitogen Activated Protein Kinase (actMAPK)/dual-phosphorylated Extracellular signal-Regulated Kinase (dpERK) levels to potentiate EGFR signaling, allowing the proper recruitment of cell types to each ommatidium. Our observations together implicate Spen in multiple activities. Our studies have taken good steps to elucidate the role of Spen during eye development, but further analysis will be required to frame a clearer picture of Spen function in this tissue. The requirement of spen for the antagonism of Notch signaling One question that arises from our analysis concerns the extent to which Spen functions in the Notch signaling pathway. We showed that heterozygous reduction of spen suppressed the wing phenotype associated with Notch haploinsufficiency. If one takes into account our observation of elevated Notch in spen clones then spen acts genetically upstream of Notch to inhibit Notch. Does the genetic interaction we see in the wing necessarily hold true for the eye? In order to rigorously test this, genetic epistasis experiments are required in the latter tissue. The analogous experiment in the eye would be to determine if dominant reduction in spen levels modifies the neurogenic phenotype associated with a temperature-sensitive allele of Notch (Dietrich and Campos-Ortega, 1984; Shellenbarger and Mohler, 1975). At the permissive temperature, Notchts1 flies have ommatidia with the normal complement of photoreceptors. However, when switched to the restrictive temperature as developing larvae, extra photoreceptors are observed (Table 5-1A)(Baker and Zitron, 1995). Heterozygous loss of spen would therefore 220 Fig. 5-1. Models for Spen regulation of the Notch pathway during eye development. (A) Spen may inhibit transcriptional activation of transcription factor X which potentiates sca expression. (B) Spen may inhibit sca transcription, a predicted feedback transcriptional target of the Notch pathway. 221 be hypothesized to suppress the Notchts phenotype if spen plays an antagonistic role in the Notch pathway (Table 5-1B). Similarly, Notchts or loss-of-function Notch may dominantly suppress the phenotype observed in spen mutant clones (Table 5-1C,D,E), namely elevated Enhancer of split basic-Helix-Loop-Helix proteins [E(spl)-bHLH], loss of proneural Atonal (Ato), and loss of dpERK at the MF. The tests described above would provide information regarding how spen and Notch genetically interact, but how does Spen molecularly feed into the Notch pathway? In vertebrate systems, the Spen-like proteins function as transcriptional repressors downstream of the pathway at the level of RBP-J(κ) [vertebrate Su(H)] (Kuroda et al., 2003; Ma et al., 2007; Oswald et al., 2002). Although we cannot rule out a role for Drosophila Spen at the level of Su(H), our evidence thus far shows that Spen interfaces with the Notch signaling pathway by limiting sca transcript. Although it is possible that Spen limits sca RNA stability, we propose that Spen inhibits the transcription of sca, a putative direct transcriptional target of the Notch pathway (Fig. 5-1A). The observations supporting this hypothesis are as follows: first, temperaturesensitive loss of Notch and Delta (Dl), which encodes Notch ligand, leads to reduced Sca levels (Baker and Yu, 1997; Baker and Zitron, 1995). Second, over-expression of activated Notch (Notchact)(Rebay et al., 1993) leads to elevated levels of sca transcript (Baker and Yu, 1997). Third, using a positional weight matrix derived from known Su(H) DNA-binding sites (Fig. 5-2A)(Bailey and Posakony, 1995; Lecourtois and Schweisguth, 1995; Marinescu et al., 2005; Wingender et al., 1996), we computationally analyzed potential cis-regulatory regions within the sca gene region (Cister: Cis-element Cluster Finder program)(Frith et al., 2001). We found three clusters of cis-regulatory sequences within sca predicted to be bound by Su(H). One cluster is positioned 5.3-kb upstream of the transcription start site; two clusters are located within the second intron (Fig. 5-2B). For reference, we analyzed predicted Su(H) binding sites within the E(spl)-mγ locus (Fig. 5-2C). The bioinformatics analyses described above suggest that Spen inhibits Notch/Su(H)- mediated transcriptional activation of sca, a putative direct target of the Notch signaling pathway. 222 223 Elevated E(spl)-bHLH; Ato loss; dpERK loss Look for ‘normal’ E(spl)-bHLH, Ato, dpERK levels Look for ‘normal’ E(spl)-bHLH, Ato, dpERK levels in spen clones Look for ‘normal’ E(spl)-bHLH, Ato, dpERK levels in spen clones spenAH393/ubi-GFPnls FRT40A ; ey-FLP/+ Notch54l9/+ ; spenAH393 FRT40A/ubi-GFPnls FRT40A ; ey-FLP/+ Notchts1/Y ; spenAH393 FRT40A/ubi-GFPnls FRT40A ; ey-FLP/+ spenAH393 FRT40A scaBP2/ubi-GFPnls FRT40A ; ey-FLP/+ scaA2-6/+ hs-FLP/+ ; scaA2-6/+ ; act<CD2<GAL4, UAS-GFPnls/UAS-Notchact hs-FLP/+ ; scaA2-6/+ ; act<CD2<GAL4, UAS-GFPnls/UAS-Notchdn spenAH393 FRT40A scaA2-6/ubi-GFPnls FRT40A ; ey-FLP/+ spen clones control Dominant suppression of spen by Notchlof/+? Notchts suppression of spen? Dominant Sca – suppression of spen? scalacZ enhancer-trap – control scalacZ enhanced with Notchact? scalacZ suppressed with Notchdn? scalacZ suppressed with spen clones? C D E F G H I J Look for reduced sca[E-GAL] expression Look for reduced sca[E-GAL] expression Look for elevated sca[E-GAL] expression sca[E-GAL] expression at and posterior to the MF Look for ‘normal’ number of photoreceptors Notchts1/Y ; spenAH393/+ Dominant suppression of Notchts by spenlof/+? B More photoreceptors within ommatidium Analysis Notchts1/Y Notchts – Control A table 5-1 epistasis and enhancer-trap analysis line Category genotype Fig. 5-2. Cis-element analysis of predicted Notch/Su(H) target genes. (A) Position-weight matrix for Su(H) binding sites. Gene regions, with transcript/CDS illustrated, showing Su(H) binding sites (and Ets binding site) for (B) sca, (C) E(spl)-mγ, (D) dMkp3, (E) PTP-ER, and (F) CG10089. 224 In order to test this hypothesis, it would be of interest to determine first how spen and sca interact genetically. As with the epistasis experiments described above, we would monitor whether a loss-of-function allele of sca could dominantly suppress the spen phenotype in mutant clones (Table 5-1F). Positive results would support the model in which spen interacts upstream to regulate sca. Next, to determine if sca is a transcriptional target of Notch, enhancer-trap analysis should be performed. β-Galactosidase (β-GAL) reporter expression for the scalacZ enhancertrap is observed at and posterior to the MF in eye imaginal discs (Table 5-1G)(Mlodzik et al., 1990). To determine Notch’s contribution, ectopic Notchact will be over-expressed clonally in the enhancer-trap background (Table 5-1H). If the Notch pathway activates sca transcriptionally then β-GAL reporter would be predicted to be elevated in the over-expression clones. Conversely, over-expression of dominant-negative Notch (Notchdn) would reduce reporter expression (Table 5-1I). In order to determine if spen is required to limit sca transcriptional expression, the sca enhancer-trap could be examined in spen mutant clones (Table 5-1J). In this case, β-GAL reporter is predicted to be elevated. Positive results that may be obtained from the above scalacZ enhancer-trap analysis would provide a motive to determine the relevance of the predicted Su(H)-bound cis-elements in vivo. Transgenic flies that bear P-elements that have any of the three cis-element clusters in tandem with lacZ could be generated. β-GAL expression from these flies would be compared with the scalacZ enhancer-trap. Notchact, Notchdn, and spen mutant analyses can then be performed with these enhancer-reporter lines as described above. Additionally, mutation of the Su(H) binding sites within the enhancer can be performed to determine the sites’ requirement for appropriate reporter expression. Alongside the in vivo experiments, in vitro electrophoretic mobility shift assays (EMSA) can also be performed to determine whether Su(H) interacts with these ciselements directly. In summary, the experiments proposed here may elucidate Spen’s role in antagonizing a positive feedback-loop that involves Notch pathway activation of sca and Sca potentiation of Notch’s signaling capability. 225 Spen opposition to Notch’s antagonism of the EGFR pathway An attractive model for how Spen modulates both EGFR and Notch signaling in the developing eye derives from the mutual antagonism between these two pathways (Doroquez and Rebay, 2006; Sundaram, 2005). Specifically, by antagonizing Notch signaling, which in turn antagonizes the EGFR pathway, spen may function positively with respect to EGFR signaling. In our experiments, we found that Ato expression was lost at the MF in spen mutant clones likely as a result of E(spl)-mediated repression. Ato is required for expression of dpERK (Chen and Chien, 1999) and we correspondingly observed loss of dpERK in our spen mutant clones. The current model suggests that Ato is responsible for the expression of Rhomboid (Rho) (Baonza et al., 2001), which encodes a protease that cleaves pro-protein ligands for EGFR activation (Lee et al., 2001; Urban et al., 2001). This model is based on experiments that show induction of β-GAL reporter from the RholacZ enhancer-trap upon ectopic expression of Ato in eye disc R7 photoreceptors (Baonza et al., 2001). (Experiments to test the Ato/Rho relationship at the MF proper were not performed.) We examined spen mutant discs and were surprised that Rho protein levels were elevated posterior to the MF (see Appendix II). Assuming that the Rho proteins whose levels are elevated in spen clones are functional, this observation appears to contradict our conclusion that spen potentiates EGFR signaling. Alternatively, the precise nature of how Ato regulates Rho and the EGFR pathway may not be completely understood and we may not be able draw firm conclusions from these observations. The Rho observations above suggest that Spen may potentiate EGFR signaling through a different node of the pathway. Our genetic analyses suggested that Spen acts closely with components downstream of Ras (Chen and Rebay, 2000) and specifically we see loss of dpERK levels in spen mutant clones. Therefore, one hypothesis is that Spen may regulate the phosphorylation state of kinases within the MAPK module by opposing the activity of phosphatases. Although it is formally possible that Spen may affect post-translational modification of such phosphatases, an attractive hypothesis is that Spen regulates the MAPKs by 226 antagonizing phosphatases that are transcriptionally activated by the Notch pathway. Precedence for Notch’s responsibility in activating MAP kinase phosphatases is found in C. elegans where LIN-12/Notch activates the MAPK phosphatase LIP-1 which in turn antagonizes MKP-1/MAPK during vulval development (Berset et al., 2001). Cis-element analysis, as performed with sca above, indicates that the characterized Drosophila MAP kinase phosphatase genes, Drosophila MAPK phosphatase 3 (dMkp3) and Protein Tyrosine Phosphatase-ERK/enhancer of Ras1 (PTPER), have one and two putative Su(H) binding sites, respectively (Fig. 5-2D,E). Therefore, these phosphatases may be direct targets of Notch signaling. However, when we tested dMkp3 and PTP-ER to see if these were transcriptionally affected in spen mutant discs, we did not observe significant change in relative mRNA levels suggesting that Spen may not transcriptionally regulate these MAP kinase phosphatases (see Appendix II). Are there other MAPK phosphatases that may be more likely to be regulated by Notch/ Su(H) activation? There are 98 phosphatases annotated in FlyBase. BLAST-comparisons identified seven of these phosphatases to be similar to dMkp3: Puc, CG115528, CG7378, Slingshot, CG14211, CG34099, and CG10089. Based on the observations of known Notch/ Su(H) regulated genes (Bailey and Posakony, 1995; Lecourtois and Schweisguth, 1995), we hypothesized that a MAPK phosphatase candidate would likely have a cluster of Su(H) cis elements Of these seven, the dual-specificity phosphatase CG10089 is annotated in FlyBase as a negative regulator of the MAPK cascade and had a cluster of nine predicted Su(H) binding sites upstream of its transcriptional start site (Fig. 5-2F). In order to pursue CG10089, it must be established first that CG10089 is a MAPK phosphatase (Karim and Rubin, 1999; Kim et al., 2002). CG10089 can be tested for general phosphatase activity using para-nitrophenyl phosphate, a general chromogenic substrate for most phosphatases. Co-immunoprecipitation experiments and GST pull-down assays may be performed to determine whether CG10089 binds to MAPK. To determine whether CG10089 is a phosphatase for MAPK, phosphatase assays may be performed by expressing activated RasV12 and epitope-tagged ERK in Drosophila S2 cultured cells with or without CG10089. Activated 227 ERK levels may be determined with the anti-dpERK monoclonal antibody or gel-shift assay detecting epitope-tagged ERK. Specificity of phosphatase activity may be determined by testing CG10089 ability to de-phosphorylate JNK. In vitro assays may affect the specificity of CG10089 phosphatase activity. In order to test in vivo relevance, a mutant in CG10089 can be tested to determine if elevated dpERK levels are measured relative to wild-type. Alternatively, if CG10089 is endogenously expressed in Drosophila S2 cell culture, knock-down of CG10089 can be performed to determine dpERK levels. CG10089 can also be over-expressed in S2 cells to determine if dpERK levels are reduced relative to control S2 cells. Once MAPK phosphatase activity has been established for CG10089, it would be interesting to determine if Spen antagonizes Notch’s ability to activate CG10089 transcription. It would be necessary to determine whether CG10089 is expressed at the appropriate time and place for this activity. RT-PCR and in situ hybridization would be performed to identify whether the transcript is expressed in Notch-active tissues, especially in eye imaginal discs. qRTPCR analysis and enhancer-trap/enhancer-lacZ analysis may be performed in these tissues to determine if over-expressed activated or dominant-negative Notch leads to elevated or reduced CG10089/β-GAL levels, respectively. To determine spen’s requirement for the regulation of this phosphatase, CG10089 phosphatase levels would be measured in spen mutant discs. Elevated transcript levels are predicted if Spen represses CG10089 levels. In order to determine if Spen regulates CG10089 via Notch, phosphatase transcript levels may be determined in spen mutant discs in a Notchts or Notch-/+ background. Reduced Notch levels are predicted to suppress upregulation of CG10089 likely to be found in spen clones. Notch-independent role for Spen in EGFR signaling Our analyses in the eye indicate that Spen acts positively with EGFR signaling and synergizes with downstream components of the pathway. Can the spen phenotypes we identified be solely attributed to Spen’s antagonism of Notch signaling? If not, then Spen may have a Notch-independent role during eye development that is more closely associated with the EGFR 228 Fig. 5-3. Spen may function as a transcription factor regulated by EGFR signaling. Phosphorylation by activation MAPK may regulate Spen’s ability to function as a transcriptional repressor or activator along with co-repressors (CoR) and co-activators (Co-Act), respectively. pathway. Thus, it is a major challenge to elucidate the mechanism of Spen function during EGFR signaling. One hypothesis is that Spen may function as a DNA-binding transcription factor induced by EGFR signaling. The following questions therefore arise: does Spen act as a DNA-binding transcription factor? Is Spen a target of phosphorylation by dpERK? Does the transcriptional activity of Spen change between active versus inactive EGFR signaling states? In addition to its role as a transcriptional co-repressor bound to Notch, the murine Spen229 ortholog Mint functions as a scaffold on the DNA that organizes and mediates transcriptional responses. In the active state, Mint binds to activating transcription factors to promote a transcriptional ‘on’ state. However, in the repressor state, Mint maintains DNA-binding and associates with repressors to facilitate a transcriptional ‘off’ state (Sierra et al., 2004). Does Spen act in an analogous manner (Fig. 5-3)? In order to address this we can determine whether Spen interacts with DNA via electrophoretic mobility shift assay (EMSA) of GT-rich and other DNA elements that are bound by Mint. Since Spen is such a large protein, an epitope-tagged DNA fragment containing the RRMs can be tested for the ability of the RRMs to bind DNA. The caveat of this experiment is that the fragment may not be in the native state for the portion of the Spen protein, which may require auxiliary sequence and structure to associate properly with DNA. Is Spen a phosphorylation target of MAPK? Spen has two predicted high-stringency MAPK phosphorylation sites at T2760 and T4328 in the middle of the protein as well as three putative MAPK binding domains (Scansite/MIT). (There are also lower stringency sites: 17 sites, 10 binding domains. It may be important to note that the Spen paralog Nito lacks high- stringency MAPK phosphorylation sites. Nito has a lower stringency phosophorylation site at S27 and three MAPK binding domains.) In order to determine if these sites are indeed phosphorylated by MAPK, kinase assays could be performed by incubating Spen fragments containing these sites with purified dpERK by determining if the peptides are phosphorylated by gel-shift assay. Phosphatase treatment and site-directed mutational analysis of the phosphoacceptor residues will confirm that these sites are responsible for any mobility difference observed in the gel-shift assays. If Spen was shown to bind DNA biochemically and to be phosphorylated by MAPK then Spen may have differential binding specificities in the presence or absence of EGFR signaling. dpERK phosphorylation of Spen may induce Spen binding or release from DNA. An ideal test for this would be to identify where Spen may be located along the genome with chromatin immunoprecipitation microarrays (ChIP-chip)(reviewed in Kim and Ren, 2006) by expressing 230 epitope-tagged Spen in vivo or in cultured cells and comparing its location in the presence or absence of activated RasV12. Because we do not have full-length constructs that express Spen, epitope-tagged Spen fragments would be required for this experiment. Another possibility is that Spen is constitutively associated with DNA, acting like a scaffold that toggles between activation and repression states dependent on EGFR signaling activities. This would be analogous to Mint in the Fibroblast Growth Factor (FGF)-mediated regulation of osteocalcin (Sierra et al., 2004). In this case, ChIP-chip experiments utilizing the epitope-tagged Spen-RRMs fragment may be informative. Will the Spen ChIP-chip results show overlap between Spen- and EGFR-regulated genes? The above experiments identify which gene regions Spen may bind, but they do not identify how these genes are affected by Spen. Are these genes regulated by Spen? Are Spen-regulated genes activated or repressed with EGFR signaling? In order to address these questions we could identify which genes have differential expression in the presence or absence of Spen. Gene expression analysis using microarrays can be performed to compare transcript levels between wild-type and spen mutant eye discs (reviewed in Stoughton, 2005)). The genes identified from this microarray analysis may relate to spen’s role in either the Notch or EGFR pathways, but we could focus on those genes that are differentially regulated both by EGFR signaling and Spen. One way to do this is to compare the microarray data from the spen microarrays with data from imaginal discs that over-express activated RasV12. These techniques could identify genes transcriptionally regulated by Spen that are induced/repressed by EGFR signaling. In these microarray experiments, how will the binding sites relate to sites bound by Ets-transcriptional effectors of EGFR signaling? Does Spen physically interact with Ets effectors like Yan or Pnt at these DNA sites? Although prevailing data overwhelmingly supports a model in which Spen functions as a transcription factor, other molecular functions are possible, and will need to be considered in the future. An in-depth proposal of these specific experiments is beyond the scope of our current discussion. 231 Concluding remarks In this thesis, we have described steps to elucidate the function of Spen during eye development. Much exciting work is yet to be done to fully understand Spen’s role in regulating the EGFR and Notch signaling pathways. It is our hope that observations made in the Drosophila may be a useful foundation for studies of Spen-like proteins in other organismal systems. References Bailey, A. M., and Posakony, J. W. (1995). Suppressor of hairless directly activates transcription of enhancer of split complex genes in response to Notch receptor activity. Genes Dev 9, 2609-2622. Baker, N. E., and Yu, S. Y. (1997). Proneural function of neurogenic genes in the developing Drosophila eye. Curr Biol 7, 122-132. Baker, N. E., and Zitron, A. E. (1995). Drosophila eye development: Notch and Delta amplify a neurogenic pattern conferred on the morphogenetic furrow by scabrous. Mech Dev 49, 173189. Baonza, A., Casci, T., and Freeman, M. (2001). A primary role for the epidermal growth factor receptor in ommatidial spacing in the Drosophila eye. Curr Biol 11, 396-404. Berset, T., Hoier, E. F., Battu, G., Canevascini, S., and Hajinal, A. (2001). Notch inhibition of RAS signaling through MAP kinase phosphatase LIP-1 during C. elegans vulval development. Science 291, 1055-1058. Chen, C. K., and Chien, C. T. (1999). Negative regulation of atonal in proneural cluster formation of Drosophila R8 photoreceptors. Proc Natl Acad Sci U S A 96, 5055-5060. Chen, F., and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10, 943-946. Dietrich, U., and Campos-Ortega, J. A. (1984). The expression of neurogenic loci in imaginal epidermal cells of Drosophila melanogaster. J Neurogenet 1, 315-332. Doroquez, D. B., and Rebay, I. (2006). Signal Integration During Development: Mechanisms of EGFR and Notch Pathway Function and Cross-Talk. Crit Rev Biochem Mol Biol 41, 339385. Frith, M. C., Hansen, U., and Weng, Z. (2001). Detection of cis-element clusters in higher eukaryotic DNA. Bioinformatics 17, 878-889. Karim, F. D., and Rubin, G. M. (1999). PTP-ER, a novel tyrosine phosphatase, functions downstream of Ras1 to downregulate MAP kinase during Drosophila eye development. Mol Cell 3, 741-750. Kim, S. H., Kwon, H. B., Kim, Y. S., Ryu, J. H., Kim, K. S., Ahn, Y., Lee, W. J., and Choi, K. Y. (2002). Isolation and characterization of a Drosophila homologue of mitogen-activated protein kinase phosphatase-3 which has a high substrate specificity towards extracellularsignal-regulated kinase. Biochem J 361, 143-151. Kim, T. H., and Ren, B. (2006). Genome-Wide Analysis of Protein-DNA Interactions. Annu Rev 232 Genomics Hum Genet 7, 81-102. Kuroda, K., Han, H., Tani, S., Tanigaki, K., Tun, T., Furukawa, T., Taniguchi, Y., Kurooka, H., Hamada, Y., Toyokuni, S., and Honjo, T. (2003). Regulation of Marginal Zone B Cell Development by MINT, a Suppressor of Notch/RBP-J Signaling Pathway. Immunity 18, 301312. Lecourtois, M., and Schweisguth, F. (1995). The neurogenic suppressor of hairless DNAbinding protein mediates the transcriptional activation of the enhancer of split complex genes triggered by Notch signaling. Genes Dev 9, 2598-2608. Lee, J. R., Urban, S., Garvey, C. F., and Freeman, M. (2001). Regulated intracellular ligand transport and proteolysis control EGF signal activation in Drosophila. Cell 107, 161-171. Li, J., Wang, J., Yang, X., Qin, H., Dong, X., Zhu, Y., Liang, L., Liang, Y., and Han, H. (2006). The Spen Homolog Msx2-Interacting Nuclear Target Protein Interacts with the E2 UbiquitinConjugating Enzyme UbcH8. Mol Cell Biochem 288, 151-157. Ma, X., Renda, M. J., Wang, L., Cheng, E. C., Niu, C., Morris, S. W., Chi, A. S., and Krause, D. S. (2007). Rbm15 Modulates Notch-induced Transcriptional Activation and Affects Myeloid Differentiation. Mol Cell Biol. Marinescu, V. D., Kohane, I. S., and Riva, A. (2005). MAPPER: a search engine for the computational identification of putative transcription factor binding sites in multiple genomes. BMC Bioinformatics 6, 79. Mlodzik, M., Baker, N. E., and Rubin, G. M. (1990). Isolation and expression of scabrous, a gene regulating neurogenesis in Drosophila. Genes Dev 4, 1848-1861. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W., et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J 21, 5417-5426. Rebay, I., Fehon, R. G., and Artavanis-Tsakonas, S. (1993). Specific truncations of Drosophila Notch define dominant activated and dominant negative forms of the receptor. Cell 74, 319329. Shellenbarger, D. L., and Mohler, J. D. (1975). Temperature-sensitive mutations of the notch locus in Drosophila melanogaster. Genetics 81, 143-162. Sierra, O. L., Cheng, S. L., Loewy, A. P., Charlton-Kachigian, N., and Towler, D. A. (2004). MINT, the Msx2 interacting nuclear matrix target, enhances Runx2-dependent activation of the osteocalcin fibroblast growth factor response element. J Biol Chem 279, 32913-32923. Stoughton, R. B. (2005). Applications of DNA microarrays in biology. Annu Rev Biochem 74, 53-82. Sundaram, M. V. (2005). The love-hate relationship between Ras and Notch. Genes Dev 19, 1825-1839. Urban, S., Lee, J. R., and Freeman, M. (2001). Drosophila rhomboid-1 defines a family of putative intramembrane serine proteases. Cell 107, 173-182. Wingender, E., Dietze, P., Karas, H., and Knuppel, R. (1996). TRANSFAC: a database on transcription factors and their DNA binding sites. Nucleic Acids Res 24, 238-241. 233 234 Appendices 235 236 Appendix I. Split ends is a tissue/promoter specific regulator of Wingless signaling Hua V. Lin1, David B. Doroquez2, Soochin Cho1, Fangli Chen2, Ilaria Rebay2 and Ken M. Cadigan1 Development. 2003. 130(14): 3125-3135 Department of Molecular, Cellular and Developmental Biology, University of Michigan, Natural Science Building, Ann Arbor, MI 48109, USA 2 Whitehead Institute for Biomedical Research, Massachusetts Institute of Technology, Cambridge, MA 02142, USA 1 H.V. Lin performed the primary work. D.B. Doroquez characterized SpenDN as a dominantnegative in the eye; this work appears in Chapter 4. 237 Abstract Wingless directs many developmental processes in Drosophila by regulating expression of specific target genes through a conserved signaling pathway. Although many nuclear factors have been implicated in mediating Wingless-induced transcription, the mechanism of how Wingless regulates different targets in different tissues remains poorly understood. We report here that the split ends gene is required for Wingless signaling in the eye, wing and leg imaginal discs. Expression of a dominant-negative version of split ends resulted in more dramatic reductions in Wingless signaling than split ends-null alleles, suggesting that it may have a redundant partner. However, removal of split ends or expression of the dominant-negative had no effect on several Wingless signaling readouts in the embryo. The expression pattern of Split ends cannot explain this tissue-specific requirement, as the protein is predominantly nuclear and present throughout embryogenesis and larval tissues. Consistent with its nuclear location, the split ends dominantnegative acts downstream of Armadillo stabilization. Our data indicate that Split ends is an important positive regulator of Wingless signaling in larval tissues. However, it has no detectable role in the embryonic Wingless pathway, suggesting that it is a tissue or promoter-specific factor. Introduction The Drosophila Wingless (Wg) protein is a founding member of the Wnt family, which consists of secreted glycoproteins that are conserved throughout the animal kingdom. Wnts play vital roles in a wide range of events during development in worms, flies, amphibians and mice (Cadigan and Nusse, 1997). Inappropriate activation of the Wnt signaling pathway has also been implicated in several forms of human cancers (Polakis, 2000). Cells respond to Wg via a highly conserved signaling cascade that centers on Armadillo 238 (Arm). In unstimulated cells, Arm is constitutively expressed but the cytosolic pool is phosphorylated by a degradation complex containing two kinases, Shaggy/Zeste white 3 (Sgg/ Zw3) and Casein Kinase I (CKI ). Phosphorylated Arm is then rapidly degraded through the ubiquitination/proteosome pathway (Yanagawa et al., 2002). Wg signaling, through a membrane receptor complex, activates the cytoplasmic protein Disheveled, which in turn inhibits the function of the degradation complex, resulting in the stabilization and accumulation of Arm (Cadigan and Nusse, 1997; Polakis, 2000). The consensus view of downstream events is that the stabilized Arm translocates to the nucleus where it forms a complex with the DNA-binding protein TCF (Pangolin–FlyBase) and two other proteins, Legless and Pygopus (Pygo) (Belenkaya et al., 2002; Kramps et al., 2002; Parker et al., 2002; Thompson et al., 2002). How this nuclear complex mediates the regulation of Wg transcriptional targets is not understood, though several other factors have also been implicated in the process (for a review, see Hurlstone and Clevers, 2002). In the absence of Arm, TCF is thought to transcriptionally repress Wg target genes by interacting with the transcriptional co-repressor Groucho (Cavallo et al., 1998). In addition, the ARID domain protein Osa has been shown to repress Wg target genes by acting in a chromatin remodeling complex that contains the bromodomain protein Brahma (Collins et al., 1999; Collins and Treisman, 2000). Binding of Arm to TCF somehow blocks the functions of these factors and converts TCF into an activator (Hurlstone and Clevers, 2002). The Wg signaling pathway is used repeatedly throughout fly development where it exerts differential regulation on many genes in various tissues and cell types (Klingensmith and Nusse, 1994; Zecca et al., 1996). The molecular basis for this specificity is not well understood. Some of these differential responses are due to combinatorial inputs of multiple signaling cascades (Campbell et al., 1993; Lockwood and Bodmer, 2002). In other instances, there is evidence suggesting that other co-factors may be involved in regulating the activity of Arm/TCF in specific tissues or stages. Such examples include the zinc-finger protein Teashirt (Tsh) and transcriptional 239 repressor Brinker in the embryonic ventral epidermis and midgut (Gallet et al., 1998; Waltzer et al., 2001; Saller et al., 2002) and the nuclear protein Lines (Lin) in the dorsal epidermis of the embryo (Hatini et al., 2000). This report describes the role of the split ends (spen) gene in Wg signaling. Spen is a predominantly nuclear protein containing three RNA recognition motifs (RRM) and a SPOC domain at the C terminus (Kuang et al., 2000; Rebay et al., 2000; Wiellette et al., 1999). Spen has previously been implicated in neuronal cell fate, survival and axonal guidance (Chen and Rebay, 2000; Kuang et al., 2000), cell cycle regulation (Lane et al., 2000) and repression of head identity in the embryonic trunk (Wiellette et al., 1999). At the genetic level, spen has been suggested to act with the Hox gene Deformed (Wiellette et al., 1999) and the EGF/Ras signaling pathway (Chen and Rebay, 2000; Rebay et al., 2000). The human homolog SHARP has been shown to act as a transcriptional co-repressor for steroid hormone receptors (Shi et al., 2001) and RBP-J , which mediates Notch signaling (Oswald et al., 2002). The mouse homolog MINT has also been shown to bind to specific DNA sequences through its RRM domains (Newberry et al., 1999). We demonstrate that Spen is a positive regulator of Wg signaling in the larval eye, wing and leg imaginal discs. Consistent with its nuclear location, Spen acts downstream of stabilized Arm. Interestingly, we could find no requirement for spen in embryonic Wg signaling, indicating that it is a tissue or target promoter specific regulator of the pathway. Materials and Methods Drosophila strains The P[sev-wgts] transgene has been described previously (Cadigan et al., 2002); the P[sevwg] stock was from Konrad Basler (Brunner et al., 1997); the P[GMR-arm*]F36 stock (Freeman and Bienz, 2001) was kindly provided by Mariann Bienz; P[GMR-hid]10 and P[GMR-rpr] were from Bruce Hay; P[GMR-hid]IM was from from John Abrams and Antony Rodriguez; P[UASEGFRDN] was from Matt Freeman (Freeman, 1996); and P[UAS-lacZ] was from the Bloomington 240 stock center, as were the Gal4 drivers P[GMR-Gal4], P[armadillo-Gal4] (arm-Gal4) and P[paired-Gal4] (prd-Gal4). P[engrailed-Gal4] (en-Gal4) and P[patched-Gal4] (ptc-Gal4) were from Ron Johnson; P[daughterless-Gal4] (da-Gal4) was from Andreas Wodarz; P[twist-Gal4] (twi-Gal4) and P[twi-Gal4]; P[24B-Gal4] were from Rolf Bodmer; P[UAS-TCFDN] was from M. Piefer; P[dpp-lacZ] (Blackman et al., 1991) was from Laurel Raftery; and dppblk was from Jessica Treisman. The spenk07721 and spenk13624 lines were obtained from the Berkeley Drosophila Genome Project (http://www.fruitfly.org). spen3, P[FRT, hs-neo]40A (Kuang et al., 2000) and spenXIE1796, P[FRT, hs-neo]40A (P. Kolodziej, personal communication) were provided by Peter Kolodziej. spen3 is a point mutation that results in a truncation after the RRMs (at amino acids 964) (Kuang et al., 2000). spenXIE1796 is a 10 bp deletion causing a frameshift that affects the C-terminal 99 residues of Spen, removing more than half of the SPOC domain (P. Kolodziej, personal communication). For mosaics, spen9C7 and spen14C2 were recombined onto a P[FRT, hs-neo]40A chromosome as described previously (Xu and Rubin, 1993). The P[FRT, hs-neo]82B, pygo10 stock was described previously (Parker et al., 2002). Clonal markers were P[arm-lacZ]3R (from D. J. Pan) and a P[Ubi-GFPnls] on 2L (Davis et al., 1995). Mitotic clones were induced with P[hs-flp]1 (Golic and Lindquist, 1989) in the wing (60-90 minute heatshock 48-72 hours after egg laying), and P[eyeless-flp]T11 (eye-flp) or P[eye-flp]T12 (Newsome et al., 2000) in the eye. Embryonic balancer chromosome markers were CyO P[larB208] (Grossniklaus et al., 1992) and TM3 eve-lacZ. P[UAS-spenDN] lines were constructed by PCR amplifying the C-terminal 2.8 kb of the spen-coding region (amino acids 4540-5476) using the oligos 5’GGAAGATCTATGCCGAAGAAGAAGCGCAAGGTGGTTGCCGCCAGTCATTTG GCACC3’ and 5’CCGCTCGAGTTAGACAGTAGCGATGACAATCAG3’, digesting with BglII and XhoI, ligating into pUAST and injecting into w1118 embryos to obtain transgenics. The first primer contains a nuclear localization signal (PKKKRKV). Two lines were used, P[UAS-spenDN]II and P[UAS-spenDN]III, the latter of which gave significantly stronger phenotypes. These transgenes cause midline glia defects similar to those previously observed in spen mutants (Chen and Rebay, 241 2000). In addition, the severity of this phenotype is enhanced in spen heterozygotes and a spenDN rough eye phenotype is suppressed in animals carrying a duplication of the spen locus (D.B.D. and I.R., unpublished). These results suggest that the SpenDN protein is acting to antagonize endogenous spen activity. Excisions were generated from spenk07721 and spenk13624 using the ∆2-3, Sb chromosome (Robertson et al., 1988); homozygous viable revertant lines were isolated. Fly crosses were maintained at 25°C unless otherwise noted. Isolation of new spen alleles The spen alleles were generated using the mutagen ethyl methane sulfonate, and identified in a screen for modifiers of the P[sev-wgts] interommatidial bristle phenotype. The screen was performed at 17.6°C as described previously (Cadigan et al., 2002). Two suppressors belonged to a single lethal complementation group and were subsequently found to be allelic to spen (see Results). The molecular nature of these alleles was not determined. Whole-mount staining and microscopy Immunostaining was as described previously (Cadigan and Nusse, 1996). Rat anti-Spen (1:1000) was from P. Kolodziej, affinity-purified rabbit anti-Wg antisera (1:50) and mouse antiDfz2 (1:50) were from R. Nusse, mouse monoclonal anti-Ac (1:20) was from the Developmental Hybridoma Bank, rat anti-Elav (1:100) was from G. Rubin, and guinea pig anti-Slp1 (1:100) was from S. Small. Guinea pig anti-Sens (1:1000) was from H. Bellen, rabbit anti-Eve (1:100) was from Z. Han and R. Bodmer, mouse monoclonal anti-En supernatant (1:2) was from the University of Iowa Hybridoma Bank, mouse and rabbit anti-ß-galactosidase (1:500) were from Sigma and Cappel, respectively. Cy3- and Cy5-conjugated secondary antibodies were from Jackson Immunochemicals and Alexa Flour 488-conjugated secondaries were from Molecular Probes. All fluorescent pictures were obtained with a Zeiss Axiophot coupled to a Zeiss LSM510 242 confocal apparatus. All images were processed as Adobe Photoshop files. Cuticles were prepared and photographed as previously described (Bhanot et al., 1999). Flies were prepared for scanning electron microscopy (SEM) as described (Cadigan et al., 2002). The samples were viewed with a scanning electron microscope and photographed using Polapan 400 film (Kodak). Results Mutations in spen dominantly suppress a sensitized Wg phenotype in the eye Misexpression of wg in the eye via the sevenless promoter (P[sev-wg]) blocks interommatidial bristle formation, owing to Wg-mediated repression of proneural gene expression (Cadigan et al., 2002; Cadigan and Nusse, 1996). P[sev-wgts] flies express a temperature-sensitive form of Wg (Wgts). At 17.6°C, the Wgts protein is partially active, and these animals have 150-200 bristles/eye (compared to 600/eye in wild type) (Cadigan et al., 2002) (Fig. Fig. 1. spen dominantly suppresses a P[sev-wgts] eye bristle phenotype. Micrographs are SEMs of adult fly heads. (A,B) Flies contain the P[sev-wgts] transgene and are heterozygous for either the parental chromosome iso5A1 (A) or spen9C7 (B), both reared at 17.6°C. The spen heterozygotes display significant suppression of the P[sev-wgts] partial loss of interommatidial bristles. A similar suppression was also seen with spen14C2 (data not shown). 243 1A). This temperature was chosen as a sensitized background in which a screen for dominant modifiers was performed in order to identify genes that interact with wg. Two suppressors of the P[sev-wgts] phenotype increase the number of bristles to ~400/eye (Fig. 1B, data not shown). They form one lethal complementation group, which was mapped to the tip of chromosome 2L between 21B4-B6. Several lines of evidence demonstrate that these suppressors are alleles of spen, and they will hereby be referred to as spen9C7 and spen14C2. First, both spen9C7 and spen14C2 fail to complement alleles of spen. Second, immunostaining using Spen-specific antiserum (Kuang et al., 2000) showed that Spen protein is dramatically reduced in clones of spen14C2 (Fig. 2, clones marked by lack of GFP), and is present at higher than wildtype levels in spen9C7 clones (data not shown). Third, spen9C7 and spen14C2 germline clone embryos have head sclerite defects similar to those seen with other spen alleles (data not shown) (Wiellette et al., 1999). Finally, identical effects on several Wg readouts in the eye and the wing were observed in clones of the suppressors and known alleles of spen (see below). We found Spen protein to be predominantly nuclear in eye imaginal discs (Fig. 2), consistent with previous reports in the embryo (Kuang et al., 2000 ; Wiellette et al., 1999). Spen is ubiquitously expressed throughout embryogenesis as well as in the larval eye, wing and leg imaginal discs (data not shown). Fig. 2. Spen protein is severely reduced in spen14C2 clones. (A-C) Confocal images of a third instar eye imaginal disc containing spen14C2 clones generated by ey-FLP. Clones were marked by the absence of nuclear GFP (A) and stained for Spen (B); the merged image is shown (C). Spen signal is predominantly nuclear (C) and is greatly reduced in spen14C2 clones. 244 Spen potentiates Wg signaling in the eye The suppression of the P[sev-wgts] phenotype suggests that Spen is a positive effector of Wg signaling in the eye. To examine this in more detail, we used clonal analysis with spen mutant alleles and spen hypomorphic combinations to explore its requirement on Wg-dependent inhibition of bristle formation and morphogenetic furrow (MF) initiation. Ectopic Wg in P[sev-wg] flies represses the expression of a proneural protein, Acheate (Ac) (Cadigan and Nusse, 1996). In clones of spen in a P[sev-wg] background, Ac is significantly derepressed (Fig. 3B), while the expression of Wg is not affected (Fig. 3C). This strongly suggests that spen alleles do not suppress the P[sev-wg] phenotypes by reducing Wg expression. Ac is not derepressed in all areas of the spen clones (Fig. 3B, arrows), indicating that a significant level of Wg-dependent Ac repression still occurs in the absence of spen. Ac derepression is occasionally seen in cells adjacent to the spen clones (Fig. 3B, arrowhead), which may be caused by defects in Ras-dependent activation of Delta (Dl) expression (see Discussion). The conclusion that spen is partially required for Wg-dependent bristle inhibition is complicated by the fact that loss of spen causes an increase in Ac expression in wild-type eyes as well (Fig. 3E,F). As Wg signaling is thought to play no physiological role in proneural gene expression in the interior of the eye (Cadigan et al., 2002; Cadigan and Nusse, 1996), these data indicate that removal of spen is affecting a Wg-independent process that could account for the effects of spen on P[sev-wg]-mediated Ac inhibition. To determine if spen was required for other Wg readouts in the eye, we examined its effect on Wg-mediated inhibition of the MF. The MF is a coordinated wave of apical constriction of the columnar epithelial cells that triggers differentiation of the fly eye. The MF starts at the early third larval instar and sweeps across the eye imaginal disc, from the posterior to the anterior (Wolff and Ready, 1993). Clusters of photoreceptors develop behind the MF (Fig. 4G,J, marked by the neuronal protein Elav). Wg is expressed at the dorsal and ventral edges of the eye disc (Fig. 4D,J), where it inhibits MF initiation (Ma and Moses, 1995; Treisman and Rubin, 1995). 245 Fig. 3. spen is required for maximal Wg-dependent repression of proneural genes in the eye. (A-D) Confocal images of a P[sev-wg] pupal eye (3-6 hours after pupal formation–APF) containing spen3 clones generated by ey-FLP. Clones of spen3 were marked by the absence of GFP (A; clonal boundaries shown by the white lines in B-D) and were stained for Ac (B) and Wg (C); the merged image is shown (D). P[sev-wg] eyes have high levels of Wg behind the morphogenetic furrow (C) and low levels of Ac (B) outside the spen3 clones. Ac is derepressed in much of the clone (B; note that cells in and ahead of the MF are not competent to express Ac) but significant Ac repression still occurs inside the clone (arrows in B). Transgenic Wg levels are unaffected in spen3 clones (C). Similar Ac derepression and P[sev-wg] expression are also seen in spen14C2 clones (data not shown). Occasional nonautonomous derepression of Ac is observed adjacent to spen clones (B, arrowhead). (E,F) Confocal images of a pupal eye (3-6 hours APF) containing spen14C2 clones generated by eyFLP. Clones of spen14C2 were marked by the absence of GFP (E) and were stained for Ac (F). Ac expression is elevated in spen14C2 clones (arrows in F; note that the laser intensity is lower than in B). 246 Decapentaplegic (Dpp) signaling at the posterior edge represses Wg expression. In the eyespecific dppblk mutant, Wg expression in early third instar discs expands posteriorly (Royet and Finkelstein, 1997) (compare Fig. 4J with 4K), causing a partial inhibition of the MF, reduced photoreceptor differentiation (Fig. 4H) and resulting in an adult small eye phenotype (Fig. 4B) (Chanut and Heberlein, 1997). Posterior expansion and upregulation of Wg is also observed in late third instar dppblk eyes (Fig. 4E). The dppblk small eye phenotype is greatly suppressed by removal of wg activity (Treisman and Rubin, 1995). To examine the effect of spen on this Wg readout, we reduced spen dosage in the dppblk background using transheterozygotes of spenk07721 and spenk13624, two P-element insertions in the spen region. These alleles fail to complement spen9C7, spen14C2 and several other spen alleles, and they have reduced viability when transheterozygous (Wielette et al., 1999) (data not shown). Reduction of spen gene activity significantly rescued photoreceptor formation and the adult small eye phenotype of dppblk eyes (compare Fig. 4I with 4H and 4C with 4B). We could detect no change in Wg expression in early third instar dppblk eyes with spen reduction (compare Fig. 4L with 4K). spen reduction does lead to a decrease in posterior Wg expression at a later stage (compare Fig. 4F with 4E), yet the dorsal and ventral Wg expression remains higher than wild type (Fig. 4D). We believe this late reduction in Wg is unlikely to influence the initiation and progression of the MF, although we cannot rule out this possibility. With this caveat, the data are consistent with spen being required for Wg signaling downstream of Wg in MF inhibition. In an otherwise wild-type eye, loss of Wg at the lateral edges allows ectopic MF initiation and inward progression (Ma and Moses, 1995; Treisman and Rubin, 1995). Therefore, removal of Wg signaling components at the lateral edge of the eye should result in an ectopic MF. This is indeed observed in mutant clones of pygo (Fig. 4M, arrow), in which Wg signaling is blocked (Belenkaya et al., 2002; Kramps et al., 2002; Parker et al., 2002; Thompson et al., 2002). To establish the role of spen in Wg-mediated MF inhibition in the wild-type situation, we looked at spen clones at similar positions. In contrast to pygo, clones of spen never give rise to ectopic 247 248 Fig. 4. spen potentiates Wg-dependent repression of the morphogenetic furrow. (A-C) SEM images of adult eyes of wild-type (A), dppblk (B) and spenk07721, dppblk/spenk13624, dppblk (C) flies. (D-L) Confocal images of eye imaginal discs with the following genotypes: spenk07721-rev/spenk13624-rev (D,G,J), spenk07721-rev, dppblk/spenk13624-rev, dppblk (E,H,K) and spenk07721, dppblk/spenk13624, dppblk (F,I,L). The k07721-rev and k13624-rev alleles are homozygous viable revertants of the spen Pelement alleles k07721 and k13624 resulting from precise excisions. Late third instar eyes (D-I) and early third instar eyes (J-L) are stained for Wg (D-F,J-L) and Elav (G-L). Ectopic Wg in dppblk eyes inhibits the furrow, thereby reducing Elav-positive photoreceptors (compare G with H) and resulting in a small eye phenotype (compare A with B). Reduced spen dosage in transheterozygotes of hypomorphic alleles increases the number of photoreceptors (I) and the adult eye size (C); the higher than normal Wg level in dppblk eyes (compare D with E and J with K) is unaffected in the dppblk, spen mutants at early third instar (L) and slightly reduced at late third instar (F). (M,N) Confocal images of third instar eye imaginal discs showing Elav (red) and either lacZ (M, green) or GFP (N, green). Clones of pygo10 (M) and spen14C2 (N) (marked by the absence of lacZ or GFP) were generated by ey-FLP. pygo clones at the edge of the eye block Wgdependent furrow inhibition and produce ectopic photoreceptors (arrow in M). By contrast, ectopic photoreceptors do not form in spen clones at similar positions (arrow in N; spen14C2: n=16; spen3: n=11). photoreceptors (Fig. 4N, arrow), suggesting that spen is not essential for Wg signaling in this context. spen is required for Wg signaling in the wing and leg To determine whether spen plays a role in Wg signaling in other tissues, we examined its effects on the developing wing. In the third instar wing imaginal disc, wg is expressed in a narrow stripe along the dorsoventral (DV) border from which it emanates to form a morphogen gradient that regulates the expression of many genes (Neumann and Cohen, 1997; Zecca et al., 1996). The zinc-finger nuclear protein Senseless (Sens) is activated by Wg signaling in the proneural clusters on either side of the DV border, immediately adjacent to the wg-expressing stripe (Nolo et al., 2000; Parker et al., 2002) (Fig. 5A). In addition, Wg signaling refines the distribution of Wg protein by auto-repression of wg expression (Rulifson et al., 1996) and downregulation of the Wg receptor Fz2 (Cadigan et al., 1998). Thus, loss of Wg signaling in this tissue would lead to loss of Sens expression, expansion of the Wg stripe, and derepression of Fz2. Loss of spen function in clones of several spen alleles leads to reduction or complete blockage of Sens expression (Fig. 5A) with incomplete penetrance. Wg levels are either normal or occasionally slightly derepressed (Fig. 5B); thus, the loss of Sens is not due to reduced Wg expression. spen clones are typically significantly smaller than their wild-type twin spots, suggesting that spen is required for optimal growth. The small size of the clones raises the possibility that trace amounts of Spen protein may remain in each cell, resulting in the incomplete penetrance of Wg defects. We used earlier heat shocks to induce hs-FLP to try to obtain larger spen clones at the DV boundary, with no success. Immunostaining showed that Spen protein levels were not detectable inside the spen3 clones in the wing (data not shown). Thus, to the best of our knowledge, spen is needed for maximal Wg signaling in the wing disc, but is not essential. The effect of a putative dominant-negative spen transgene (spenDN) on Wg targets in the presumptive wing provides evidence that the clonal analysis of spen underestimates the 249 Fig. 5. spen is required for Wg signaling in the wing imaginal disc. (A-C) Confocal images of third instar wing imaginal discs containing randomly generated spenXIE1796 clones (marked by the absence of GFP, boundary shown by white lines). Samples were stained for Sens (A) and Wg (B); the merged image is shown (C). Reduction or total lack of Sens (A) was observed in some spen clones (spen9C7: 23% and 26%, respectively, n=31; spen14C2: 50% and 6.7%, n=30; spen3: 27% and 0%, n=30; spenXIE1796: 30% and 8.6%, n=70), and Wg level is unaffected or slightly derepressed (B), consistent with attenuated Wg signaling in cells that lack spen. (D-G) Confocal images of third instar wing imaginal discs containing the transgenes P[en-Gal4] and P[UAS-spenDN]II stained for Sens (D,F), Wg (E) and Fz2 (DFz2 in figure) (G). Flies were reared at 18°C. Expression from P[UAS-spenDN] in the posterior compartment of the wing reduces (D) or eliminates (F) Sens expression (the posterior compartment is towards the right of the arrows in F and G; 44% are similar to D, 27% are more severe but some Sens remains and 29% are as in F, n=68) and derepresses Wg (E, 97%, n=31). By contrast, expression of Fz2 is either unaffected (compare arrowheads; G, 73%, n=45) or reduced throughout the posterior compartment (data not shown, 27%). 250 contribution of spen to Wg signaling. The spenDN construct contains the C-terminal 936 amino acids of spen, including a nuclear localization signal and the conserved SPOC domain. en-Gal4 was used to express spenDN throughout the posterior compartment of the wing disc, while the anterior compartment remains wild type. With this Gal4 driver, spenDN caused reduced (Fig. 5D) or complete absence (Fig. 5F) of Sens expression. The penetrance of loss of Sens expression was higher than seen in spen clones. More strikingly, a significant Wg depression was always observed, even when Sens is only moderately affected (Fig. 5E). By contrast, no effect on Wgdependent Fz2 inhibition was observed (Fig. 5G). Thus, the spenDN construct indicates an absolute requirement for spen in Wg stripe refinement and an important role in Wg-mediated regulation of Sens, but it has no discernable role in Wg-dependent Fz2 repression. We believe that the spenDN experiments may also underestimate the importance of spen in Wg signaling. Gal4 is known to be cold sensitive in flies and we observe a strict temperature dependence in our en-Gal4/spenDN experiments. The discs shown in Fig. 5 were reared at 18°C. At higher temperatures (e.g. 20°C or 25°C) where Gal4 is more active, we observe either gross deformities of the wing disc or organismal lethality before the 3rd instar larval stage (data not shown). Thus, we cannot assay the effect of spenDN on Wg signaling when expressed at higher levels than the ones shown in Fig. 5. However, at those levels, spenDN expression results in morphologically normal wing discs with strong Wg signaling defects. In the leg imaginal discs, we see a similar situation as has just been described in the developing wing (Fig. 6). In the leg, Wg signaling inhibits dpp expression in the ventral portion of the discs (Brook and Cohen, 1996; Heslip et al., 1997; Jiang and Struhl, 1996) (Fig. 6A-C). Expression of spenDN with ptc-Gal4, which is active in a stripe overlapping both the dpp and wg expression domains, causes a complete breakdown of disc morphology (data not shown) at 25°C. At lower temperatures (18-22°C), small leg discs are observed with either normal restriction of dpp-lacZ expression to the dorsal half (Fig. 6D-F) or derepression in the ventral region (Fig. 6GI). This derepression is consistent with a block in Wg signaling. 251 Fig. 6. spen is required for Wg signaling in the leg imaginal disc. (A-I) Confocal images of third instar leg imaginal discs of wild type (A-C) and animals containing transgenes P[ptcGal4], P[dpp-lacZ] and P[UAS-spenDN]II (D-I) reared at 21°C. Samples were stained for Wg (A,D,G) and lacZ (B,E,H); merged images are shown (C,F,I). All panels are shown with the same magnification. Ventral (v) expression of Wg (A) restricts lacZ expression to the dorsal (d) half (B) in the wild type. Expression from P[UAS-spenDN] along a stripe overlapping both the wg and dpp domains either has no effect on the restriction of lacZ expression (D-F) or leads to derepression of lacZ in the ventral region (G-I; note the overlap of Wg and lacZ in I; 9% are similar to I, 45% have significant but incomplete derepression of lacZ, and 46% are similar to F, n=35). This derepression is consistent with a block in Wg signaling. 252 spen acts downstream of Arm stabilization To address the question of where spen acts in the Wg signaling pathway, we carried out epistasis analysis using spenDN. In the absence of Wnt signaling, Arm is phosphorylated at serine and threonine residues at its N terminus by CKI and Sgg/Zw3, and then degraded (Peifer et al., 1994; Yanagawa et al., 2002). Mutations in these residues render Arm resistant to degradation (Freeman and Bienz, 2001; Pai et al., 1997). Expression of these Arm mutants (Arm*) in flies activates Wg signaling independent of upstream components (Pai et al., 1997). When expressed under the control of the eye-specific GMR enhancer, Arm* reduces eye size dramatically (Freeman and Bienz, 2001) (Fig. 7A). Co-expression of spenDN using the GMRGal4 driver severely suppresses this phenotype (Fig. 7B), indicating that spen potentiates Wg signaling downstream of Arm stabilization. Expression of spenDN could also suppress a small eye phenotype caused by GMR-driven expression of a dominant-negative version of the EGF receptor (EGFRDN) but not of the pro-apoptotic factors head involution defective or reaper (data not shown). This specificity indicates that spenDN does not affect GMR enhancer-mediated expression. Fig. 7. spenDN acts downstream of Arm stabilization. Micrographs are SEMs of adult fly heads from flies containing P[GMR-Gal4], P[GMR-arm*]F36 and either P[UAS-lacZ] (A) or P[UASspenDN]II (B) reared at 25°C. Expression of an activated form of Arm activates Wg signaling and produces an eye that is severely reduced in size and is suppressed by P[UASspenDN]. 253 Fig. 8. spen is not required for Wg signaling in the embryo. (A,D) Micrographs of cuticles of wild type (A) and embryos containing transgenes P[da-Gal4] and P[UAS-spenDN]III (D) reared at 25°C. Expression from P[UAS-spenDN] has variable effects on cuticle patterning, ranging from wild type (P[UAS-spenDN]II at 25°C, data not shown), moderate reduction of denticles (D) to complete disruption of cuticle formation (P[da-Gal4], P[UAS-spenDN]III containing embryos reared at 29°C, data not shown). (B,C,E,F) Confocal images of stage 11 embryos containing P[prd-Gal4] and either P[UAS-TCFDN] (25°C; B,C) or P[UAS-spenDN]III (29°C; E,F). Samples were stained for En (B,E) or Slp1 (C,F). En and Slp1 stripes remain wild type in spenDNexpressing embryos. (G-I) Confocal images of stage 13 wild-type embryos (G) and embryos containing transgenes P[twi-Gal4] and P[UAS-spenDN]II (H,I) reared at 25°C. Expression from P[UAS-spenDN] has variable effects on the Eve pericardial expression, ranging from wild type (H) to an increase in Eve-positive cells (I). Some embryos containing transgenes P[twi-Gal4], P[24B-Gal4] and P[UAS-spenDN]II reared at 25°C or 29°C also display disorganization of Eveexpressing cells or occasional gaps missing Eve expression, in addition to an overall increase in Eve-positive cells (data not shown). These effects are qualitatively different from a blockage of Wg signaling [ectopic denticles, loss of En (B) and Slp-1 (C) stripes and loss of Eve in pericardial cells]. 254 spen is not required for Wg signaling in the embryo Embryos completely lacking spen gene activity have head defects and ectopic sclerite formation (Wiellette et al., 1999), as well as axonal path-finding and midline glial cell defects (Chen and Rebay, 2000; Kuang et al., 2000). No defects in Wg-dependent developmental decisions were observed. We have confirmed this by examining spen germline clones for several Wg readouts (described in detail below), all of which were normal (data not shown). As the spenDN construct gave stronger Wg defects in the imaginal discs than loss of spen, we also examined the effects of spenDN expression on Wg targets in the embryo. Wild-type embryos have a distinctive denticle patterning on their ventral cuticles with trapezoidal arrays of denticle belts intermittent with naked cuticles (Fig. 8A). Wg signaling is required for naked cuticle formation; and wg mutants form ectopic denticles in place of naked cuticle (Nusslein-Volhard and Wieschaus, 1980). The cuticles of embryos carrying UASspenDN and the ubiquitous driver da-Gal4 were examined. At 25°C, these embryos show cuticle phenotypes ranging from wild type (data not shown) to moderate reduction of denticle formation (Fig. 8D), inconsistent with reduced Wg signaling. At 29°C, spenDN expression causes complete disruption of cuticle formation (data not shown). At moderate expression levels, spenDN caused cuticular defects, similar to those reported for spen mutants (Wiellette et al., 1999), including defects in head structures (e.g. reduced or missing dorsal bridge) and apparent ectopic sclerites (data not shown). At higher expression levels, there was an almost complete block of head involution and head cuticle formation. Thus, the lack of Wg signaling defects in the cuticle is unlikely to be due to insufficient expression of spenDN. Additionally, prd-Gal4, which is expressed in alternating segments throughout the embryo, was also used to express spenDN, and no ventral cuticle defects were observed, even at 29°C. The effects of spenDN expression were further characterized using molecular markers. En is normally expressed in epidermal stripes of single segment periodicity. Wg is required for the maintenance of En expression (DiNardo et al., 1988); expression of a dominant-negative TCF 255 (TCFDN), which blocks Wg signaling in the nucleus, causes En to fade from the epidermis by full germband extension (Fig. 8B). Expression of spenDN using prd-Gal4 (Fig. 8E), or the ubiquitous drivers arm-Gal4 and da-Gal4 (data not shown) does not affect En expression at full germband extension, indicating that spen is not needed for En maintenance. Similarly, expression of another Wg target in the embryonic ectoderm, Sloppy-paired 1 (Slp1) (Lee and Frasch, 2000), is markedly reduced by TCFDN expression (Fig. 8C), but is not affected by spenDN expression under the control of prdGal4 (Fig. 8F), armGal4 or daGal4 (data not shown). Wg signaling in the mesoderm is required for the expression of Even-skipped (Eve) in a subset of pericardial cells (Wu et al., 1995) (Fig. 8G). At 25°C or 29°C, Eve expression in embryos expressing spenDN throughout the mesoderm using twi-Gal4 ranges from wild-type (Fig. 8H) to an increased number of Eve-positive cells (Fig. 8I). Some embryos expressing spenDN via two mesodermal drivers (twi-Gal4 and 24B-Gal4) simultaneously exhibit general disorganization of Eve-positive pericardial cells, with occasional segmental gaps (one or two per embryo) missing Eve expression, and an overall increased number of Eve-positive cells (data not shown). As Eve expression is always present in these embryos, and segments missing Eve are always concurrent with those with more Eve expression in the same embryo, we conclude that spen is not required for Wg signaling in this readout, and that functions of spen in other pathways or the non-specificity of spenDN may be the culprit for the defects in Eve expression. Discussion Spen is required for normal Wg signaling in imaginal discs In this study, a total of seven distinct readouts of Wg signaling were examined in imaginal discs. They are: inhibition of interommatidial bristle formation (Figs 1, 3); MF initiation/progression (Fig. 4); repression of Wg and DFz2, and activation of Sens expression at the presumptive wing margin (Fig. 5); inhibition of dpp expression in the dorsal leg (Fig. 6); 256 and reduction of eye size (Fig. 7). Wg regulation of six of these readouts is significantly blocked by partial or complete removal of spen and/or the expression of spenDN. These results provide a strong genetic argument that spen is required for Wg signaling in these tissues. Interpreting spen phenotypes is complicated by the fact that spen has been implicated in several other pathways. Can these functions explain the apparent loss of Wg signaling phenotypes we observed? spen has been found to act with Deformed to suppress head identity in the embryonic trunk (Wiellette et al., 1999) and spen genetically interacts with cell cycle mutants (Lane et al., 2000). We think it unlikely that these spen functions can account for the phenotypes observed. However, Spen has also been shown to be involved with the Ras and Notch signaling pathways, which do affect the readouts we employed for studying Wg signaling. Therefore, it is possible that some of the spen phenotypes we have documented are due to disruption of these signaling cascades, though we argue below that this is unlikely. spen mutations affect some Ras targets in a way that suggests it acts positively in Ras signaling (Chen and Rebay, 2000; Rebay et al., 2000). This may be the explanation for the nonautonomous derepression of Ac expression adjacent to spen clones in P[sev-wg] eyes (Fig. 3B, arrowhead), as Dl expression is activated by the EGF/Ras pathway in the eye (Tsuda et al., 2002). Ras signaling plays a positive role in MF progression (Greenwood and Struhl, 1999; Kumar and Moses, 2001) and elevated Ras signaling can suppress a Wg or Arm induced small eye phenotype (Freeman and Bienz, 2001) (K.M.C., unpublished). Therefore, a reduction in Ras signaling caused by loss of spen cannot explain our observations. Ras signaling has no effect on wing margin formation (Diaz-Benjumea and Hafen, 1994; Nagaraj et al., 1999) and acts downstream of Wg/Dpp crossregulation in the leg (Campbell, 2002; Galindo et al., 2002), again arguing that the role of Spen in Ras signaling cannot account for the apparent Wg signaling defects we observed. Expression of Suppressor of Hairless [Su(H)], a transcription factor required for Notch signaling, is significantly reduced in spen mutant embryos (Kuang et al., 2000). Can a reduction of Notch signaling explain our results? Notch signaling is required for interommatidial bristle 257 inhibition (Cagan and Ready, 1989) so this could explain the requirement of spen for Wgdependent Ac inhibition (Fig. 3). However, Notch is absolutely required for Wg expression at the DV stripe in the wing (Rulifson and Blair, 1995) and plays a positive role in MF progression (Kumar and Moses, 2001). Thus, reducing Notch activity by loss of spen or spenDN cannot explain the wider Wg stripe (Fig. 5) and suppression of the dppblk MF defect (Fig. 4) that we observed. Though no evidence for elevated Notch signaling in spen mutants has been reported in Drosophila, a recent report has suggested that SHARP, a human Spen homolog, functions as a transcriptional co-repressor for RBP-J /CBL, the ortholog of Su(H) (Oswald et al., 2002). In addition, the fly homolog of human SMRT, which binds to SHARP (Shi et al., 2001), has been shown to act as a negative regulator of Notch signaling (Tsuda et al., 2002). This could mean that loss of spen activity in flies results in higher expression of Notch/Su(H) targets, owing to derepression. Although this could conceivably contribute to the MF and wing phenotypes we found, such derepression could not account for the suppression of Wg-dependent reduction of eye size and bristle inhibition (Figs 1, 3) or the derepression of dpp expression in the leg (Fig. 6). In summary, the only explanation consistent with all the spen (or spenDN) imaginal disc phenotypes discussed above is a loss of Wg signaling. Spen is not required for embryonic Wg signaling In contrast to the data in the imaginal discs, we could find no evidence for the involvement of spen in Wg signaling in the embryo (Fig. 8), either by removing spen gene activity or expressing spenDN. Thus, it appears that Spen may be a tissue-specific regulator of Wg signaling. Spen is a predominately nuclear protein expressed ubiquitously in embryos and imaginal discs (Kuang et al., 2000; Wiellette et al., 1999) (Fig. 2; data not shown). It could be that a Spen co-factor is not expressed in embryos, or that Spen is post-translationally modified in a tissue-specific way. Alternatively, the specificity could lie in the promoters of the targets that were tested. This appears to be the case in the wing, where Wg and Sens regulation by Wg signaling is spen dependent (Fig. 5A-F), while that of Fz2 is not (Fig. 5G). 258 The negative results we obtained in the embryo cannot be viewed as definitive. Embryos that lack maternal and zygotic spen activity could be normal for Wg signaling because of redundancy (see below). Likewise, even though expression of spenDN in the imaginal discs caused strong Wg loss of function phenotypes (Figs 5, 6, 7), and caused spen-like phenotypes under mild expression conditions in the embryo (data not shown), it is possible that we did not supply adequate amounts of spenDN in our embryonic assays. To address this issue, we used several Gal4 drivers at 29°C (to ensure optimal Gal4 activity; see Results for details). We did observe phenotypes with spenDN not previously reported that are Wg independent. For example, arm-Gal4- and da-Gal4-driven spenDN expression causes reduced denticle formation to varying degrees in the embryonic ventral cuticle (Fig. 8). A possible explanation is reduced DER/Ras signaling, which promotes the denticle fate by activating shavenbaby (Payre et al., 1999). In addition, spenDN expression also causes a variable increase in the number of Eve-expressing cells in the embryonic dorsal mesoderm. This could be explained by a reduction in Su(H) levels, as impairment of Notch signaling causes an increase in Eve-positive pericardial cells (Carmena et al., 2002). Under conditions where spenDN blocked other pathways, we could observe no reduction in Wg signaling. Spen may have a redundant partner Our experiments with loss of function spen alleles indicate that spen is not absolutely required for Wg signaling in the wing and eye. Although reduction of spen activity could suppress a dppblk MF defect (Fig. 4C), which can be explained by a reduction in Wg signaling, complete removal of spen did not cause an ectopic MF (Fig. 4N). Because removal of Wg signaling is known to induce an ectopic MF (Ma and Moses, 1995; Treisman and Rubin, 1995) (Fig. 4M), this indicates that sufficient Wg signaling still occurs in the spen clones. In the wing, spen clones affect Wg readouts, but with incomplete penetrance (Fig. 5A-C), again indicating a partial reduction in Wg signaling in the absence of spen. Our experiments with spenDN suggest that the partial loss of Wg signaling in spen mutants 259 may be due to redundancy. Expressing spenDN causes more severe phenotypes and much higher penetrance in disruption of Sens and expansion of Wg in the wing than complete removal of spen (Fig. 5D-F). A likely explanation is that the SpenDN protein also inhibits the function of another gene that has roles in the Wg pathway redundant to spen. Although many genes exist in the fly genome that encode proteins containing RRMs, only one other besides Spen is predicted to encode a protein with both RRMs and a SPOC domain. This factor has been called short Spen-like protein (SSLP or DmSSp) (Kuang et al., 2000; Wiellette et al., 1999) and is referred to as CG2910 in the annotated genome. No genetic or molecular characterization of SSLP has been reported and we are pursuing its possible redundancy with spen. Mechanism of Spen action Where does Spen act in the Wg pathway? Our epitasis experiments in the eye (Fig. 7) indicate that SpenDN blocks Wg signaling downstream of Arm stabilization. Thus, Spen could act in Arm nuclear import, or in mediating TCF/Arm transcriptional regulation. Consistent with a role in Wg target gene transcription, Spen is predominantly nuclear in imaginal tissues (Fig. 2; data not shown). In addition, the mouse and human homologs of Spen have been implicated as transcription factors (see below). Studies on the vertebrate homologs of Spen have provided functions for the RRM and SPOC domain that these proteins share with Spen. Spen has three predicted RRMs near its N terminus. The role of RRMs in specific RNA binding is well established (Burd and Dreyfuss, 1994) and the RRM domains in the human Spen homolog SHARP has been shown to bind to the steroid receptor RNA co-activator SRA (Shi et al., 2001). By contrast, the RRM domain of the mouse Spen homolog, MINT, has been shown to bind to specific double-stranded DNA, including the proximal promoter of the osteocalcin gene (Newberry et al., 1999). SHARP also binds to the nuclear receptor co-repressor SMRT and acts as a transcription corepressor by recruiting histone deacetylases (HDACs) through its SPOC domain (Shi et al., 2001). A similar 260 co-repressor function for SHARP with the DNA-binding protein RBP-Jκ/CBL has also been reported (Oswald et al., 2002). Finally, MINT was also found to interact with Msx2, a known transcriptional repressor (Newberry et al., 1999). These studies on the vertebrate homologs suggest that Spen may bind DNA or RNA at its N terminus, and may regulate the Wg pathway as a transcription corepressor. Why is spen required for only a subset of Wg targets? Based on studies with its vertebrate homologs, could spen only regulate the Wg targets that are transcriptionally repressed by TCF/ Arm? Wg-dependent transcriptional inhibition through TCF has been shown for the stripe gene in the embryo (Piepenburg et al., 2000) and has been suggested for bristle inhibition in the eye (Cadigan et al., 2002). However, no direct targets of Wg signaling in the imaginal discs have been determined and our attempts to determine whether stripe repression in the embryo requires spen have been inconclusive (H.V.L. and K.M.C., unpublished). It is interesting to note that two embryonic targets tested which were spen independant, eve and slp1, are both directly activated by TCF/Arm (Halfon et al., 2000; Knirr and Frasch, 2001; Lee and Frasch, 2000; Han et al., 2002). Identification of spen-dependent direct targets of Wg signaling will be necessary to explore this model. Two factors have previously been reported that are tissue/promoter-specific regulators of Wg signaling. tsh has been shown to be required for Wg-mediated inhibition of denticle formation in the ventral embryonic epidermis (Gallet et al., 1998) and lin, which is needed for Wg signaling only in the dorsal epidermis (Hatini et al., 2000). We report a third factor, Spen, which is only needed for imaginal disc regulation of Wg targets. The existence of these specific factors begs the question: what is the difference between the various Wg targets that requires such specificity? Acknowledgements The authors thank K. Matthews, K. Cook and the Bloomington Stock Center, and everyone mentioned in Materials and Methods for providing fly stocks and antisera. They 261 also thank Bruce Donohue of the Univeristy of Michigan Cell Biology Laboratory for expert operation of the SEM, R. Bodmer for helpful discussions, and D. S. Parker for preparing samples of pygo clones and careful reading of the manuscript. H.V.L. was supported by the Cancer Biology Training Program at the University of Michigan Comprehensive Cancer Center (UMCCC). This work was supported by a Rackham Grant and a Pheonix Memorial Grant and from the University of Michigan, a Beginning Investigator Grant from the UMCCC and NIH grant RO1 GM59846 to K.M.C. References Belenkaya, T. Y., Han, C., Standley, H. J., Lin, X., Houston, D. W. and Heasman, J. (2002). pygopus Encodes a nuclear protein essential for wingless/Wnt signaling. Development 129,4089 -4101. Bhanot, P., Fish, M., Jemison, J. A., Nusse, R., Nathans, J. and Cadigan, K. M. (1999). Frizzled and Dfrizzled-2 function as redundant receptors for Wingless during Drosophila embryonic development. Development 126,4175 -4186. Blackman, R. K., Sanicola, M., Raftery, L. A., Gillevet, T. and Gelbart, W. M. (1991). An extensive 3’ cis-regulatory region directs the imaginal disk expression of decapentaplegic, a member of the TGF-beta family in Drosophila. Development 111,657 -666. Brook, W. J. and Cohen, S. M. (1996). Antagonistic interactions between wingless and decapentaplegic responsible for dorsal-ventral pattern in the Drosophila Leg. Science 273,1373 -1377. Brunner, E., Peter, O., Schweizer, L. and Basler, K. (1997). pangolin encodes a Lef-1 homologue that acts downstream of Armadillo to transduce the Wingless signal in Drosophila. Nature 385,829 -833. Burd, C. G. and Dreyfuss, G. (1994). Conserved structures and diversity of functions of RNAbinding proteins. Science 265,615 -621. Cadigan, K. M. and Nusse, R. (1996). wingless signaling in the Drosophila eye and embryonic epidermis. Development 122,2801 -2812. Cadigan, K. M. and Nusse, R. (1997). Wnt signaling: a common theme in animal development. Genes Dev. 11,3286 -3305. Cadigan, K. M., Fish, M. P., Rulifson, E. J. and Nusse, R. (1998). Wingless repression of Drosophila frizzled 2 expression shapes the Wingless morphogen gradient in the wing. Cell 93,767 -777. Cadigan, K. M., Jou, A. D. and Nusse, R. (2002). Wingless blocks bristle formation and morphogenetic furrow progression in the eye through repression of Daughterless. Development 129,3393 -3402. Cagan, R. L. and Ready, D. F. (1989). Notch is required for successive cell decisions in the developing Drosophila retina. Genes Dev. 3,1099 -1112. Campbell, G. (2002). Distalization of the Drosophila leg by graded EGF-receptor activity. Nature 262 418,781 -785. Campbell, G., Weaver, T. and Tomlinson, A. (1993). Axis specification in the developing Drosophila appendage: the role of wingless, decapentaplegic, and the homeobox gene aristaless. Cell 74,1113 -1123. Carmena, A., Buff, E., Halfon, M. S., Gisselbrecht, S., Jimenez, F., Baylies, M. K. and Michelson, A. M. (2002). Reciprocal regulatory interactions between the Notch and Ras signaling pathways in the Drosophila embryonic mesoderm. Dev. Biol. 244,226 -242. Cavallo, R. A., Cox, R. T., Moline, M. M., Roose, J., Polevoy, G. A., Clevers, H., Peifer, M. and Bejsovec, A. (1998). Drosophila Tcf and Groucho interact to repress Wingless signalling activity. Nature 395,604 -608. Chanut, F. and Heberlein, U. (1997). Role of decapentaplegic in initiation and progression of the morphogenetic furrow in the developing Drosophila retina. Development 124,559 -567. Chen, F. and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10,943 -946. Collins, R. T. and Treisman, J. E. (2000). Osa-containing Brahma chromatin remodeling complexes are required for the repression of wingless target genes. Genes Dev. 14,3140 3152. Collins, R. T., Furukawa, T., Tanese, N. and Treisman, J. E. (1999). Osa associates with the Brahma chromatin remodeling complex and promotes the activation of some target genes. EMBO J. 18,7029 -7040. Davis, I., Girdham, C. H. and O’Farrell, P. H. (1995). A nuclear GFP that marks nuclei in living Drosophila embryos; maternal supply overcomes a delay in the appearance of zygotic fluorescence. Dev. Biol. 170,726 -729. Diaz-Benjumea, F. J. and Hafen, E. (1994). The sevenless signalling cassette mediates Drosophila EGF receptor function during epidermal development. Development 120,569 578. DiNardo, S., Sher, E., Heemskerk-Jongens, J., Kassis, J. A. and O’Farrell, P. H. (1988). Twotiered regulation of spatially patterned engrailed gene expression during Drosophila embryogenesis. Nature 332,604 -609. Freeman, M. (1996). Reiterative use of the EGF receptor triggers differentiation of all cell types in the Drosophila eye. Cell 87,651 -660. Freeman, M. and Bienz, M. (2001). EGF receptor/Rolled MAP kinase signalling protects cells against activated Armadillo in the Drosophila eye. EMBO Rep. 2, 157-162. Galindo, M. I., Bishop, S. A., Greig, S. and Couso, J. P. (2002). Leg patterning driven by proximal-distal interactions and EGFR signaling. Science 297,256 -259. Gallet, A., Erkner, A., Charroux, B., Fasano, L. and Kerridge, S. (1998). Trunk-specific modulation of wingless signalling in Drosophila by teashirt binding to armadillo. Curr. Biol. 8,893 -902. Golic, K. G. and Lindquist, S. (1989). The FLP recombinase of yeast catalyzes site-specific recombination in the Drosophila genome. Cell 59,499 -509. Greenwood, S. and Struhl, G. (1999). Progression of the morphogenetic furrow in the Drosophila eye: the roles of Hedgehog, Decapentaplegic and the Raf pathway. Development 126,5795 -5808. Grossniklaus, U., Pearson, R. K. and Gehring, W. J. (1992). The Drosophila sloppy paired locus encodes two proteins involved in segmentation that show homology to mammalian 263 transcription factors. Genes Dev. 6,1030 -1051. Halfon, M. S., Carmena, A., Gisselbrecht, S., Sackerson, C. M., Jimenez, F., Baylies, M. K. and Michelson, A. M. (2000). Ras pathway specificity is determined by the integration of multiple signal-activated and tissue-restricted transcription factors. Cell 103,63 -74. Han, Z., Fujioka. M., Su, M., Liu, M., Jaynes, J. B. and Bodmer, R. (2002). Transcriptional Integration of Competence Modulated by Mutual Repression Generates Cell-Type Specificity within the Cardiogenic Mesoderm. Dev. Biol. 252,225 -240. Hatini, V., Bokor, P., Goto-Mandeville, R. and DiNardo, S. (2000). Tissue- and stage-specific modulation of Wingless signaling by the segment polarity gene lines. Genes Dev. 14,1364 1376. Heslip, T. R., Theisen, H., Walker, H. and Marsh, J. L. (1997). Shaggy and dishevelled exert opposite effects on Wingless and Decapentaplegic expression and on positional identity in imaginal discs. Development 124,1069 -1078. Hurlstone, A. and Clevers, H. (2002). T-cell factors: turn-ons and turn-offs. EMBO J. 21,2303 -2311. Jiang, J. and Struhl, G. (1996). Complementary and mutually exclusive activities of decapentaplegic and wingless organize axial patterning during Drosophila leg development. Cell 86,401 -409. Klingensmith, J. and Nusse, R. (1994). Signaling by wingless in Drosophila. Dev. Biol. 166,396 -414. Knirr, S. and Frasch, M. (2001). Molecular integration of inductive and mesoderm-intrinsic inputs governs even-skipped enhancer activity in a subset of pericardial and dorsal muscle progenitors. Dev. Biol. 238,13 -26. Kramps, T., Peter, O., Brunner, E., Nellen, D., Froesch, B., Chatterjee, S., Murone, M., Zullig, S. and Basler, K. (2002). Wnt/wingless signaling requires BCL9/legless-mediated recruitment of pygopus to the nuclear beta-catenin-TCF complex. Cell 109,47 -60. Kuang, B., Wu, S. C., Shin, Y., Luo, L. and Kolodziej, P. (2000). split ends encodes large nuclear proteins that regulate neuronal cell fate and axon extension in the Drosophila embryo. Development 127,1517 -1529. Kumar, J. P. and Moses, K. (2001). The EGF receptor and notch signaling pathways control the initiation of the morphogenetic furrow during Drosophila eye development. Development 128,2689 -2697. Lane, M. E., Elend, M., Heidmann, D., Herr, A., Marzodko, S., Herzig, A. and Lehner, C. F. (2000). A screen for modifiers of cyclin E function in Drosophila melanogaster identifies Cdk2 mutations, revealing the insignificance of putative phosphorylation sites in Cdk2. Genetics 155,233 -244. Lee, H. H. and Frasch, M. (2000). Wingless effects mesoderm patterning and ectoderm segmentation events via induction of its downstream target sloppy paired. Development 127,5497 -5508. Lockwood, W. and Bodmer, R. (2002). The patterns of wingless, decapentaplegic, and tinman position the Drosophila heart. Mech. Dev. 114,13 . Ma, C. and Moses, K. (1995). Wingless and patched are negative regulators of the morphogenetic furrow and can affect tissue polarity in the developing Drosophila compound eye. Development 121,2279 -2289. Nagaraj, R., Pickup, A. T., Howes, R., Moses, K., Freeman, M. and Banerjee, U. (1999). Role 264 of the EGF receptor pathway in growth and patterning of the Drosophila wing through the regulation of vestigial. Development 126,975 -985. Neumann, C. J. and Cohen, S. M. (1997). Long-range action of Wingless organizes the dorsalventral axis of the Drosophila wing. Development 124,871 -880. Newberry, E. P., Latifi, T. and Towler, D. A. (1999). The RRM domain of MINT, a novel Msx2 binding protein, recognizes and regulates the rat osteocalcin promoter. Biochemistry 38,10678 -10690. Newsome, T. P., Schmidt, S., Dietzl, G., Keleman, K., Asling, B., Debant, A. and Dickson, B. J. (2000). Trio combines with dock to regulate Pak activity during photoreceptor axon pathfinding in Drosophila. Cell 101,283 -294. Nolo, R., Abbott, L. A. and Bellen, H. J. (2000). Senseless, a Zn finger transcription factor, is necessary and sufficient for sensory organ development in Drosophila. Cell 102,349 -362. Nusslein-Volhard, C. and Wieschaus, E. (1980). Mutations affecting segment number and polarity in Drosophila. Nature 287,795 -801. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W. et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J. 21,5417 -5426. Pai, L. M., Orsulic, S., Bejsovec, A. and Peifer, M. (1997). Negative regulation of Armadillo, a Wingless effector in Drosophila. Development 124,2255 -2266. Parker, D. S., Jemison, J. and Cadigan, K. M. (2002). Pygopus, a nuclear PHD-finger protein required for Wingless signaling in Drosophila. Development 129,2565 -2576. Payre, F., Vincent, A. and Carreno, S. (1999). ovo/svb integrates Wingless and DER pathways to control epidermis differentiation. Nature 400,271 -275. Peifer, M., Sweeton, D., Casey, M. and Wieschaus, E. (1994). wingless signal and Zeste-white 3 kinase trigger opposing changes in the intracellular distribution of Armadillo. Development 120,369 -380. Piepenburg, O., Vorbruggen, G. and Jackle, H. (2000). Drosophila segment borders result from unilateral repression of hedgehog activity by wingless signaling. Mol. Cell 6,203 -209. Polakis, P. (2000). Wnt signaling and cancer. Genes Dev. 14,1837 -1851. Rebay, I., Chen, F., Hsiao, F., Kolodziej, P. A., Kuang, B. H., Laverty, T., Suh, C., Voas, M., Williams, A. and Rubin, G. M. (2000). A genetic screen for novel components of the Ras/ Mitogen-activated protein kinase signaling pathway that interact with the yan gene of Drosophila identifies split ends, a new RNA recognition motif- containing protein. Genetics 154,695 -712. Robertson, H. M., Preston, C. R., Phillis, R. W., Johnson-Schlitz, D. M., Benz, W. K. and Engels, W. R. (1988). A stable genomic source of P element transposase in Drosophila melanogaster. Genetics 118,461 -470. Royet, J. and Finkelstein, R. (1997). Establishing primordia in the Drosophila eye-antennal imaginal disc: the roles of decapentaplegic, wingless and hedgehog. Development 124,4793 -4800. Rulifson, E. J. and Blair, S. S. (1995). Notch regulates wingless expression and is not required for reception of the paracrine wingless signal during wing margin neurogenesis in Drosophila. Development 121,2813 -2824. Rulifson, E. J., Micchelli, C. A., Axelrod, J. D., Perrimon, N. and Blair, S. S. (1996). wingless refines its own expression domain on the Drosophila wing margin. Nature 384,72 -74. 265 Saller, E., Kelley, A. and Bienz, M. (2002) The transcriptional repressor Brinker antagonizes Wingless signaling. Genes Dev. 16,1828 -1838. Shi, Y., Downes, M., Xie, W., Kao, H. Y., Ordentlich, P., Tsai, C. C., Hon, M. and Evans, R. M. (2001). Sharp, an inducible cofactor that integrates nuclear receptor repression and activation. Genes Dev. 15,1140 -1151. Thompson, B., Townsley, F., Rosin-Arbesfeld, R., Musisi, H. and Bienz, M. (2002). A new nuclear component of the Wnt signalling pathway. Nat. Cell Biol. 22, 22. Treisman, J. E. and Rubin, G. M. (1995). wingless inhibits morphogenetic furrow movement in the Drosophila eye disc. Development 121,3519 -3527. Tsuda, L., Nagaraj, R., Zipursky, S. and Banerjee, U. (2002). An EGFR/Ebi/Sno pathway promotes Delta expression by inactivating Su(H)/SMRTER repression during inductive Notch signaling. Cell 110,625 . Waltzer, L., Vandel, L. and Bienz, M. (2001). Teashirt is required for transcriptional repression mediated by high Wingless levels. EMBO J. 20,137 -145. Wiellette, E. L., Harding, K. W., Mace, K. A., Ronshaugen, M. R., Wang, F. Y. and McGinnis, W. (1999). spen encodes an RNP motif protein that interacts with Hox pathways to repress the development of head-like sclerites in the Drosophila trunk. Development 126,5373 -5385. Wolff, T. and Ready, D. F. (1993). Pattern Formation in the Drosophila Retina. In The Development of Drosophila melanogaster, Vol. II (ed. A. Bate and M. Martinez Arias), pp. 1277-1326. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. Wu, X., Golden, K. and Bodmer, R. (1995). Heart development in Drosophila requires the segment polarity gene wingless. Dev. Biol. 169,619 -628. Xu, T. and Rubin, G. M. (1993). Analysis of genetic mosaics in developing and adult Drosophila tissues. Development 117,1223 -1237. Yanagawa, S., Matsuda, Y., Lee, J. S., Matsubayashi, H., Sese, S., Kadowaki, T. and Ishimoto, A. (2002). Casein kinase I phosphorylates the Armadillo protein and induces its degradation in Drosophila. EMBO J. 21,1733 -1742. Zecca, M., Basler, K. and Struhl, G. (1996). Direct and long-range action of a wingless morphogen gradient. Cell 87,833 -844. 266 Appendix II. Further insights into Spen’s requirement during Drosophila eye development David B. Doroquez and Ilaria Rebay 267 Results The Notch and Epidermal Growth Factor (EGFR) signaling pathways interact cooperatively and antagonistically to regulate many aspects of Drosophila development. In Chapter 2, we reported that Split ends (Spen) functions to antagonize Notch and to promote EGFR signaling during retinal cell differentiation. Here, we report additional insights to the function of spen during Drosophila eye development. spen mutant clones lack Spen protein The spenAH393 allele was isolated as a dominant modifier of the rough-eye associated with eye-specific expression of constitutive yan repressor (sev-yanACT) (Rebay et al., 2000). The molecular lesion of this allele has not been described but spenAH393 represents a strong, or ‘nulllike’, allele of spen. In order to determine if Spen is produced in spenAH393 mutant tissue, we looked for Spen protein expression in spenAH393 mutant clones with an anti-Spen monoclonal antibody (IR25). In spen mutant clones, the ubiquitous, nuclear expression of Spen observed in wild-type is undetectable in mutant tissue (Fig. AII-1). The peripheral signal observed between cells is likely background. These results that Spen protein is not expressed in spenAH393 mutant clones and therefore spen clones may be used in spen loss-of-function analysis. spen is required for proper adult eye patterning We were interested in how loss of spen affected the adult eye. Therefore, we first looked at the gross morphology of the adult eye under the dissecting scope and the scanning electron microscope (SEM). Wild-type eyes have regular patterning of ommatidia (Fig. AII-1C). However, spen mutant eyes are rough and lack the highly regular pattern of ommatidia, normal hexagonal morphology, and regular spacing of ommatidial bristles observed in wild-type. The eyes are smaller and have a partial transformation of eye tissue into head cuticle, observed by the increased area of hair and cuticle along ventral-posterior region of the eye. Furthermore, 268 Fig. AII-1. spen mutation in the eye leads to loss of Spen protein and a rough eye phenotype. Confocal micrograph of third larval instar eye imaginal disc from spenAH393 mutant eye clones stained (in red) for Spen (B). spen mutant tissue lacks GFP (A, green). Disc is orientated anterior to the left. (C,D) Scanning electron micrographs (SEMs) of adult eyes for (C) wild-type and (D) spen mutant (EGUF hid/cell lethal). ommatidial bristles appear to be longer and thinner in spen mutant eyes compared to wild-type (Fig. AII-1D). 269 Fig. AII-2. Loss of spen disrupts morphology of the adult eye. Tangential sections of adult compound eyes from (A) wild-type (Canton S) and (B,C) spen clones. Photoreceptor loss and ommatidial disorganization is evident in spen mutant tissue which is marked by lack of the white gene product. At this section level, seven of the eight rhabdomeres are visible in a trapezoidal pattern (see schematic on right). Dorsal-plane ommatidia have R3s that are pointed up; ventralplane ommatidia have R3s that are pointed down. 270 We also analyzed tangential sections in wild-type and spen mutant clones. In these sections, only seven of the eight photoreceptor R cells are visible. These R cells are marked by their darkly staining rhabdomeres that are arrayed in a trapezoidal pattern (Fig. AII2A). Perturbation of adult ommatidia observed in spen mutants is significantly penetrant. Sixty-six percent of spen ommatidia have a reduced complement of rhabdomeres (Table 1). Approximately five percent of the spen ommatidia had a greater number of rhabdomeres than are typically observed at a given plane of section (Table 1), which may be due in part to splitting of rhabdomeres with an R cell or to a failure of rhabdomere positioning in R8 and/or R7 (data not shown). Although 29 percent of spen ommatidia had a normal complement of seven rhadomeres for tangential sections, 91 percent of these seven-rhabdomere-containing ommatidia had defects in normal rhabdomere trapezoid pattern and/or directional orientation (Table 2). The orientation defects observed in the eye are consistent with observations from the wing that spen may play a role in tissue planar polarity (Mace and Tugores, 2004). However, from these analyses, it is unclear if the spen orientation defect is a secondary effect from the loss of cell types in 271 neighboring ommatidia. In toto, spen is required for the proper complement and pattern of ommatidial cell types, consistent with earlier descriptions of spen mutant eyes (Dickson et al., 1996). Loss of spen leads to reduced dpERK and elevated Yan but not yan transcript Consistent with the loss of Ato expression signaling in the intermediate group (IG) cells due to hyper-activated Notch at the Morphogenetic Furrow (MF), we observed dpERK expression in the IGs to be drastically reduced in spen mutant clones. We did not observe change of activated mitogen activated protein kinase (actMAPK)/dual-phosphorylated Extracellular signal-Regulated Kinase (dpERK) posterior to the MF (see Chapter 3). Western blot analysis probing for dpERK expression in eye discs predominantly mutant for spen (see Materials and Methods) indicates close to a two-fold reduction in dpERK levels relative to wild-type discs (Fig. AII-3A). It is unclear whether this reduction represents the loss of dpERK solely at the MF or if it also reflects a loss of dpERK posterior to the MF that is below the detection threshold of the antibody in situ. Regardless, the reduced dpERK levels that result from loss of spen suggest Spen may potentiate EGFR signaling via antagonism of the Notch pathway. We found that output from the EGFR signaling pathway is compromised in spen mutant clones. In spen clones, Yan expression is elevated in cells both at and posterior to the MF. Western blot analysis of eye discs mutant for spen confirms that Yan protein levels are elevated relative to wild-type (Fig. AII-3B). We found that Spen regulated Yan levels post-translationally via regulation of dpERK, but it was also feasible that yan is regulated transcriptionally by Spen. This possibility seems reasonable because Notch signaling has been previously shown to transcriptionally activate yan expression (Rohrbaugh et al., 2002); and second, Spen family proteins are implicated as transcriptional repressors of Notch pathway target genes (Kuroda et al., 2003; Ma et al., 2007; Oswald et al., 2002). To address this question, we asked whether loss of spen affects the expression of yan transcript in situ. yan is expressed strongly at the MF in wild-type tissue (Fig. AII-4). Posterior to the MF, yan is expressed at low levels (Lai and Rubin, 272 Fig. AII-3. Loss of spen leads to increased reduced dpERK and Yan levels. (A) Protein immunoblot probed simultaneously for (A) dpERK or (B) Yan and the loading reference βtubulin. Relative dpERK or Yan levels from wild-type and spen/cell lethal eye discs were calculated amongst three samples (mean ± st. dev.). 1992) but mRNA levels are undetectable via in situ hybridization (Fig. AII-4). In spen mutant clones, yan transcript is maintained at the MF, but we did not see elevated yan mRNA levels posterior to the MF. The yan minimal eye enhancer, yanO (Rohrbaugh et al., 2002), contains three Su(H) DNA-binding sites and a single Ets (Pointed (Pnt)/Yan) binding site. The expression from the yanO-lacZ reporter is similar to the β-GAL reporter expression of yanP (Ramos et al., 2003). At the time we were performing the above yan transcript experiments, we wanted to determine if Spen regulated yanO-lacZ output in vivo. The prediction was that if spen repressed yan transcriptionally, the levels of β-GAL from the lacZ reporter would be elevated. Surprisingly, we observed reduced β-GAL expression from the yanO-lacZ reporter in spen mutant clones (Fig. AII-5A), suggesting that Spen negatively regulated this transcriptional element. We reasoned that this result was observed because high Yan levels in the spen clones may regulate a negative feedback loop. Elevated Yan could repress yan transcription via binding to the Ets sites within the minimal enhancer. In order to test this hypothesis, we generated spen yan double-mutant clones and asked if yanO-lacZ was still reduced in these clones. We still observed reduced 273 Fig. AII-4. yan transcript levels are not elevated in spen mutant clones. (A,B) In situ hybridization of spen clones in eye discs to detect yan mRNA in third instar eye discs. spen mutant tissue lacks GFP. yan mRNA is readily detectable at but not posterior to the MF in wildtype and spen mutant tissue. yanO-lacZ expression in the spen yan clones (Fig. AII-5B). In addition, we surprisingly found that the pntP1lacZ enhancer trap was upregulated in spen mutant clones (Fig. AII-5C). The pntP1 isoform of pnt is an EGFR pathway target, activated by the PntP2 isoform (Frankfort and Mardon, 2004). We predicted that any change in pntP1 regulation in spen clones would lead to upregulation of pntP1 transcript because its transcriptional activation may be antagonized by elevated Yan levels. We cannot currently explain why we had observed these results, but we hypothesize that these effects do not reflect the normal function of Spen on its endogenous targets. 274 Fig. AII-5. yanO-lacZ and pnt-P1 enhancer-trap in spen clones is reduced and elevated, respectively. Confocal (A) and (B,C) epifluorescence micrographs of third instar eye discs with (A) spen clones in yanO-lacZ background, (B) spen yan clones in yanO-lacZ background, and (C) spen clones in pnt1277(lacZ) background. (A,C) spen or (B) spen yan mutant tissue lack GFP. 275 Fig. AII-6. Delta and Rhomboid levels are elevated with loss of spen. Confocal micrographs of third instar eye imaginal discs in spen mutant clones. Discs, oriented anterior to the left, were stained (red) for (A,B) Delta or (C,D) Rhomboid. spen mutant tissue is marked by lack of GFP (green). 276 Elevated Delta and Rhomboid observed in spen mutant clones We identified Spen as a negative of Notch signaling. Specifically we saw elevated expression of Sca, Notch, and E(spl)-bHLHs in spen mutant clones. We were interested in identifying other components of the Notch pathway that may be regulated by Spen. Therefore, we asked whether the expression of the Notch ligand Delta (Dl) was perturbed with spen loss. In spen mutant clones, we observed mild upregulation of Dl at and posterior to the MF (Fig. AII-6A,B). Interestingly, cis­-element cluster analysis predicts that Dl is a transcriptional target of both the EGFR and Notch signaling pathways (Fig. AII-7A)(Tsuda et al., 2006; Tsuda et al., 2002). Both Ets- and Su(H)-binding site clusters are found upstream to the Dl transcriptional start site. In the current model, Notch antagonizes the EGFR signaling pathway by restricting the number of cells that express the proneural protein Atonal (Ato). Ato activates rhomboid (rho) expression and Rho in turn promotes the activational processing of Transforming Growth Factor (TGF)-α-like ligands for active EGFR signaling (Baonza et al., 2001; Chen and Chien, 1999). Since we saw loss of Ato expression (see Chapter 3), we hypothesized that Rho protein levels would be reduced. However, we find that Rho levels were elevated in spen mutant clones (Fig. AII-6C,D). However, these elevated levels do not reflect an increase in EGFR pathway output suggesting that another EGFR signaling component may be limiting to allow pathway activation. Intriguingly, we also identified predicted Ets- and Su(H)-binding site clusters upstream of the rho transcriptional start site (Fig. AII-7). These data indicate that regulation of both Dl and Rho is complicated by possible Notch and EGFR pathway interactions. Spen may normally limit Dl and rho expression by antagonizing Notch/Su(H) activation of these predicted transcriptional targets. However, it is unclear how this would affect EGFR pathway regulation of these targets. Transcript level analysis in spen mutants: Notch, ato, rho, and MAPK phosphatases We saw protein levels affected in spen clones related to Notch and EGFR pathway components (see Chapter 3). We thus wanted to determine whether corresponding transcripts 277 were affected as well. In spen mutant clones, Notch protein was significantly upregulated, however, we see that the levels of Notch are not greatly affected suggesting that the mechanism of Notch activation of Sca may play the predominant role in Notch’s activation (Fig. AII-8A). We also observed loss of Ato protein at the MF in spen mutant clones, but, in spen mutant discs ato transcript is unchanged (Fig. AII-8D). The result that we did not see a change in ato transcript may be due to the observation that Ato expression appears to recover and Ato-expressing do not represent a large proportion of cells that are tested for transcript quantitation. We also observed a slight decrease in rho transcript (Fig. AII-8D) which does not follow with its elevated protein expression pattern observed in spen mutant clones (Fig. AII-6C,D). The slight change that we observed here is not statistically significant. In spen mutant tissue, we observed loss of dpERK (see Chapter 3). One plausible mechanism for Spen function is that it might mediate transcriptional repression of an inhibitor such as a MAPK phosphatase (reviewed in Farooq and Zhou, 2004). qRT-PCR analysis of spen mutant discs do not indicate a role for Spen in regulating the expression of two characterized MAPK phosphatases—Drosophila MAPK phosphatase 3 (dMkp3) and Protein Tyrosine Phosphatase-ERK/enhancer of Ras1 (PTP-ER)(Fig. AII-8F,G). However, it is formally possible that Spen may regulate the expression of other MAPK phosphatases. Increased cell adhesion observed with loss of spen Prior to photoreceptor differentiation significant cell shape rearrangements occur within and posterior to the MF. Groups of cells are clustered into arcs or rosettes that are the early glimpses of each ommatidium (reviewed in Wolff and Ready, 1993). The molecular nature how these shape changes occur is unclear. Studies are beginning to elucidate morphology changes at the MF. Recently, the EGFR signaling pathway and Ato was identified as regulators of cell shape at this stage. Both Ato and EGFR are required for the formation of ommatidial pre-clusters, arcs, and rosettes. EGFR regulates adherens junctions indirectly at a post-transcriptional level, 278 Fig. AII-7. Dl and rho are predicted Ets and Notch pathway transcriptional targets. Cister cis-element cluster analysis to identify Ets (red) and Su(H) (blue) binding sites in the (A) Dl or (B) rho gene regions. Cartoons depict gene and mRNA/CDS. 279 280 Fig. AII-8. Quantitative RT-PCR. qRT-PCR was performed to measure relative (A) Notch (p= 0.02), (B) E(spl)-mδ (p=0.78), (C) E(spl)-mγ (p=0.34), (D) ato (p=0.72), (E) rho (p=0.37), (F) dMkp3 (p=0.56), (G) PTP-ER (p=0.69) mRNA levels in wild-type and spen/cell lethal eye discs (three samples, mean + st. dev.). Fig. AII-9. Elevated cell adhesion in basal MF cells in spen mutant clones. Confocal micrographs of third instar eye imaginal discs in spen mutant clones. Discs, oriented anterior to the left, were stained (red) for (A) DE-cad or (B) Arm. (A,C) Apica and (B,D) Basal projections are shown. spen mutant tissue lacks GFP. 281 Fig. AII-10. spen loss does not affect the cell cycle. Confocal micrographs of third instar eye imaginal discs in spen mutant clones. Discs, oriented anterior to the left were stained for (A,B) GFP (green) and anti-BrdU (red) to detect cells in S phase or (C,D) Phospho-histone H3 to detect cells in mitosis. spen mutant tissue lacks GFP. 282 whereas Ato controls the transcription of DE-cadherin/shotgun (DE-cad/shg) which encodes a cell-adhesion molecule (Brown et al., 2006). We were interested to do determine whether Spen regulated cell adhesion at the MF so we analyzed the expression of DE-Cad and the adherens junction molecule Armadillo/β-catenin (Arm/β-Cat). ato and Egfr mutants do not have cluster formation as shown at the apical surface of the eye imaginal disc (Brown et al., 2006). We find that cluster formation in spen clones is retained at the apical surface (Fig. AII-9A,C), suggesting that Spen does not play a role in EGFR and Ato’s role in cluster development. On the other hand, spen mutant tissue has elevated DECad (Fig. AII-9B) and Arm (Fig. AII-9D) at the basal level of the eye imaginal disc (Fig. AII9D). Increased cell adhesion at the MF may prevent undifferentiated cells from changing shape properly, leading to a failure in differentiation. Spen does not affect cell cycle regulation The cell cycle is highly regulated during Drosophila eye development (reviewed in Doroquez and Rebay, 2006)). Anterior to the MF, cells undergo asynchronous replication and mitosis. As the MF passes through the disc cells arrest at G1 phase and begin to differentiate. A few cell rows behind the MF, cells that have not begun differentiation undergo a final round of mitosis called the ‘second mitotic wave’ and then re-arrest at G1. spen was isolated as an enhancer of the rough eye phenotype associated with eye-specific expression of Drosophila E2F and DP (GMR-dE2F,dDP) (Staehling-Hampton et al., 1999), suggesting a role for Spen in the cell cycle. Does Spen have a role in the cell cycle? To determine if Spen acted to regulate cell division, we analyzed eye discs with cell cycle markers. S phase/replication analysis by bromodeoxyuridine (BrdU) labeling shows that the BrdU foci pattern corresponding to DNA replication is similar between wild-type and spen mutant clones (Fig. AII-10A,B). We also observe that phospho-histone H3 (PH3) immunostaining to detect cells undergoing mitosis is not 283 Fig. AII-11. Cell death during pupal and larval eye development. (A,B) Confocal micrographs of 30 hr APF pupal eye discs in spen mutant clones. Discs were stained (red) for activated, cleaved Caspase-3 (Casp-3*). (C,D) spen/cell lethal third instar eye imaginal discs stained with acridine orange have high levels of apoptosis. 284 different between wild-type and spen mutant clones (Fig. AII-10C,D). Therefore, our analyses indicate that Spen is not likely to have a significant role in regulating the cell cycle. However, if Spen does indeed play a role in the cell cycle, it may be redundant in its activity to other proteins. Spen and Programmed Cell Death Cells that have not differentiated in the developing eye undergo apoptosis or programmed cell death (PCD) during pupal stages (Wolff and Ready, 1991). Our analyses of spen mutants suggest that there may be elevated PCD in spen mutant tissue because fewer cell types are observed in pupal eye discs and adult eye retinas (see Chapter 3). The identification of PCD genes as modifiers in our dominant-negative spen screen (see Chapter 4) suggest that spen regulates PCD. Therefore, we wanted to confirm this role by determining if markers for PCD were perturbed in spen mutants. At 30 hr APF, when activate caspases are active (Yu et al., 2002), we immunostained pupal discs for activated, cleaved Caspase-3 (Casp-3*), but did not observe a significant change in Casp-3* staining between wild-type and spen mutant tissue (Fig. AII-11A,B). However, in larval discs mostly mutant for spen, we observed elevated cell death as marked by acridine orange staining (Fig. AII-11C,D), but it is unclear if the cell death observed here is due to the cell-lethal allele associated in generating these spen mutant discs (see Materials and Methods). Thus, it is unclear if Spen is required to regulate cell death in the developing eye. Materials and Methods Fly Strains and Genetics Fly stocks were maintained at 25ºC and obtained from the Bloomington Stock Center. w1118 was used as the control strain, unless otherwise indicated. For the generation of mitotic clones in the eye imaginal disc, the following genotype was analyzed: P[ey-FLP] ; spenAH393 285 P[FRT]40A / P[w+, ubi-GFPnls] P[FRT]40A. For tangential sections, adult eyes were fixed, embedded in plastic, sectioned, and mounted as described (Wolff, 2000). Mutant tissue was recognizable by the absence of red eye pigment (Xu and Rubin, 1993). The following are the lacZ/enhancer-trap line genotypes analyzed: P[ey-FLP]/+; spenAH393 P[FRT]40A/P[ubi-GFPnls] P[FRT]40A; P[yanO-lacZ]/+, P[ey-FLP]/+; spenAH393 yanE433/P[ubi-GFPnls] P[FRT]40A; P[yanO-lacZ]/+ P[ey-FLP]/+; spenAH393 P[FRT]40A/P[ubi-GFPnls] P[FRT]40A ; pnt1277 P[lacZ]/+ To generate eye discs predominantly mutant for spen, the EGUF/hid cell lethal technique (Stowers and Schwarz, 1999) was employed to produce the genotype: spenAH393 P[FRT]40A / P[GMR-hid] l(2)cl-L31 P[FRT]40A ; P[ey-Gal4] P[UAS-FLP]/+, or P[ey-FLP] ; spenAH393 P[FRT]40A / l(2)cl-L31 P[FRT]40A to analyze cell death. Immunohistochemistry and Immunofluorescence Eye imaginal discs were dissected from wandering third instar larvae, fixed with 4% paraformaldehyde in PBT (0.1% Triton X-100 in Phosphate-Buffered Saline [PBS]) for 15 min, washed three times in PBT (5 min each), incubated in primary antibody overnight in PBT with 5% normal goat serum (PNT) at 4ºC, and then washed three times in PBT (5 min each). Discs were incubated in secondary antibody for 2 hr in PNT at room temperature, washed three times in PBT (5 min each), and mounted in Vectashield mountant (Vector Laboratories). Microscopy and image acquisition was performed on a Zeiss LSM510 confocal microsope. For immunofluorescence, the following primary antibodies were used: 1:50 mouse anti- Spen (mAbIR25, I. Rebay), 1:1000 guinea pig anti-GFP (a gift of M.L. Pardue), 1:20,000 rabbit anti-β-Galactosidase (Cappel Laboratories), 1:1000 mouse anti-Delta (C594.9B, Developmental Studies Hybridoma Bank [DSHB]), 1:500 rabbit anti-Rhomboid (Sturtevant et al., 1996), 1:50 mouse anti-DE-Cadherin (DCAD2, DSHB), 1:200 mouse anti-Armadillo (N2 7A1, DSHB), 1:100 rabbit anti-BrdU (Becton Dickinson), 1:1000 rabbit anti-phospho-Histone H3 (Upstate Biotech), 1:100 rabbit anti-Cleaved Caspase-3 (Asp175, Cell Signaling #9661S). Secondary 286 antibodies were Alexa Fluor 568-conjugated goat anti-mouse IgG (Molecular Probes), or appropriate Rhodamine Red-X- and Cy5-conjugated antibodies (Jackson ImmunoResearch). BrdU labeling was performed as described previously (Sullivan et al., 2000). For acridine orange staining, eye imaginal discs from wandering third instar larvae were dissected in PBS and stained in 1.6 µM acridine orange (Sigma) in PBS for 5 min. After three brief washes in PBS, the discs were mounted in PBS and viewed immediately by epifluorescence microscopy. Western Blot Analysis Dissected tissue from forty eye imaginal disc pairs (control and spen/cell lethal clones) was lysed in NP-40 Lysis Buffer (50 mM Tris-HCl [pH 8.0], 150 mM NaCl, 1 mM DTT, 1 mM PMSF, 1% NP-40, 0.25% Protease Inhibitor Cocktail [Sigma]) mixed with 2× SDS sample buffer and boiled. Protein samples were resolved on 8% SDS-PAGE gels and transferred by semi-dry blot to a PVDF membrane (Millipore Immobilon P). Membranes were blocked and probed with 1:200 mouse anti-Yan antibody, 1:10,000 mouse anti-dpERK, 1:400 rat anti-β-tubulin (YL1/2, YL1/34, Harlan Seralab) in PBS with 5% non-fat milk (Block Buffer) overnight at 4˚C. Washes were performed in TBS with 0.15% Tween-20 (TTBS) (3×10 min each). Membranes were incubated with HRP-conjugated secondary antibodies (Jackson ImmunoResearch) in Block Buffer for 2 hr at room temperature. Washes were performed as above, in TTBS with a 5 min TBS final wash. Chemiluminescent detection was done using the ECL Plus Western Blotting Detection System (Amersham Biosciences) according to the manufacturer’s instructions. In Situ Hybridization In situ DNA probes were prepared with probe labeling reaction [cDNA, Hexanucleotide Mix (Boehringer Mannheim), DIG DNA Labeling Mix (Boehringer Mannheim), DNA polymerase Klenow fragment] followed with purification LiCl/ethanol precipitation. Eye imaginal discs from wandering third instar larvae were dissected in Schneider Insect Cell Medium (Sigma) and fixed in 9:1 PP* (4% paraformaldehyde, 0.1% Tween-20, 0.1% Triton X287 100 in PBS):DMSO at room temperature and washed in PBT. Discs were proteinase K, treated with glycine, and post-fixed in 9:1 PP*/DMSO. Discs were pre-hybridized at 80°C, cooled, and hybridized with in situ probes overnight at 50°C in Hybe Solution (50% formamide, 5× SSC, 100 µg/mL salmon sperm DNA, 100 µg/mL tRNA, 50 µg/mL heparin, 0.1% Tween-20). Following hybridization, discs were washed with Hybe/PBT and incubated with pre-absorbed 1:2000 AP-conjugated sheep anti-digoxigenin (anti-DIG-AP) in PBT overnight at 4°C. For detection, colorimetric reaction was performed with nitro blue tetrazolium (NBT), 5-bromo-4-chloro3-indolyl phosphate (BCIP) in AP buffer (100 mM Tris, 100 mM NaCl, 50 mM MgCl2, 0.1% Tween-20). Discs were then immunostained to detect GFP as described above and mounted for brightfield and epifluorescence microscopy. Quantitative RT-PCR Forty pairs of eye imaginal discs were dissected from control and spen/cell lethal clone third instar larvae. Total RNA was isolated using Trizol Reagent (Invitrogen) according to the manufacturer’s protocol, treated with DNase I (Invitrogen), and used as template to generate cDNA using the random or oligo-dT primers provided in the Reverse Transcription System (Promega). Quantitative/real-time PCR from 1 µL cDNA template was performed using 2× ABI SYBR Green Master Mix (Applied Biosystems), 80 nM forward primer, and 80 nM reverse primer in 25 µL reactions for 35 thermocycling reactions. Each experimental reaction was performed in triplicate using the relative quantitation method (ABI) or alongside four ten-fold dilutions of standard [wild-type eye disc cDNA] and no-template control reactions (in triplicate) as previously described in (Claycomb et al., 2002). For each sample, relative fluorescence was measured in comparison to standard curves. Mean and standard deviations of the triplicate reactions were calculated (ABI Prism 7300 software). Fold-expression of the experimental transcript was normalized to RpS17 expression for each sample. Analysis to determine statistical significance was performed with paired, two-tailed Student’s t-tests. 288 PCR primers are as follows: RpS17 forward 5’-GTACGAACCAAGACGGTGAAGA-3’ RpS17 reverse 5’-GGCGAGTGTAGTACTTCTCGATG-3’ Notch forward 5’-CCACGAAAAAGACAGTAACCAG-3’ Notch reverse 5’-CGTAGATCGTCGAGTGTCCTTAT-3’ atonal forward 5’-CCCAGTCCCAAGTACTATTCTCA-3’ atonal reverse 5’-AGCTACAACATTTGGCAAGCTC-3’ rhomboid forward 5’-GGCCGCAATTACTACGAATGAA-3’ rhomboid reverse 5’-GATACAGTTTGCTGATGCTTCG-3’ dMkp3 forward 5’-GACTCGGAAGCGTTGAAAAAGT-3’ dMkp3 reverse 5’-TGGCAAATCTGGTGTCACATTC-3’ PTP-ER forward 5’-TAGCGAGGACTTGCCCTACATTA-3’ PTP-ER reverse 5’-CACATAGCACTTCGACACGTAG-3’ Acknowledgements We would like to thank R.G. Fehon, P.A. Garrity, J.C. Jemc, P. Vivekanand, J.A. Lees and the Fehon, Orr-Weaver, and Rebay labs for stimulating discussion. We are grateful to A. Mukherjee, S. Rashkova, and M.L. Pardue and the Artavanis-Tsakonas and Garrity labs for reagents, and to N. Watson and E. Batchelder (WIBR/Keck Microscopy Facility) for microscopy assistance. D.B.D. was supported by a National Science Foundation Graduate Fellowship. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. 289 References Baonza, A., Casci, T., and Freeman, M. (2001). A primary role for the epidermal growth factor receptor in ommatidial spacing in the Drosophila eye. Curr Biol 11, 396-404. Brown, K. E., Baonza, A., and Freeman, M. (2006). Epithelial cell adhesion in the developing Drosophila retina is regulated by Atonal and the EGF receptor pathway. Dev Biol. Chen, C. K., and Chien, C. T. (1999). Negative regulation of atonal in proneural cluster formation of Drosophila R8 photoreceptors. Proc Natl Acad Sci U S A 96, 5055-5060. Claycomb, J. M., MacAlpine, D. M., Evans, J. G., Bell, S. P., and Orr-Weaver, T. L. (2002). Visualization of replication initiation and elongation in Drosophila. J Cell Biol 159, 225-236. Dickson, B. J., van der Straten, A., Dominguez, M., and Hafen, E. (1996). Mutations Modulating Raf signaling in Drosophila eye development. Genetics 142, 163-171. Doroquez, D. B., and Rebay, I. (2006). Signal Integration During Development: Mechanisms of EGFR and Notch Pathway Function and Cross-Talk. Crit Rev Biochem Mol Biol 41, 339385. Farooq, A., and Zhou, M. M. (2004). Structure and regulation of MAPK phosphatases. Cell Signal 16, 769-779. Frankfort, B. J., and Mardon, G. (2004). Senseless represses nuclear transduction of Egfr pathway activation. Development 131, 563-570. Kuroda, K., Han, H., Tani, S., Tanigaki, K., Tun, T., Furukawa, T., Taniguchi, Y., Kurooka, H., Hamada, Y., Toyokuni, S., and Honjo, T. (2003). Regulation of Marginal Zone B Cell Development by MINT, a Suppressor of Notch/RBP-J Signaling Pathway. Immunity 18, 301312. Lai, Z. C., and Rubin, G. M. (1992). Negative control of photoreceptor development in Drosophila by the product of the yan gene, an ETS domain protein. Cell 70, 609-620. Ma, X., Renda, M. J., Wang, L., Cheng, E. C., Niu, C., Morris, S. W., Chi, A. S., and Krause, D. S. (2007). Rbm15 Modulates Notch-induced Transcriptional Activation and Affects Myeloid Differentiation. Mol Cell Biol. Mace, K., and Tugores, A. (2004). The product of the split ends gene is required for the maintenance of positional information during Drosophila development. BMC Dev Biol 4, 15. Oswald, F., Kostezka, U., Astrahantseff, K., Bourteele, S., Dillinger, K., Zechner, U., Ludwig, L., Wilda, M., Hameister, H., Knochel, W., et al. (2002). SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J 21, 5417-5426. Ramos, E., Price, M., Rohrbaugh, M., and Lai, Z. C. (2003). Identifying functional cis-acting regulatory modules of the yan gene in Drosophila melanogaster. Dev Genes Evol 213, 83-89. Rebay, I., Chen, F., Hsiao, F., Kolodziej, P. A., Kuang, B. H., Laverty, T., Suh, C., Voas, M., Williams, A., and Rubin, G. M. (2000). A genetic screen for novel components of the Ras/Mitogen-activated protein kinase signaling pathway that interact with the yan gene of Drosophila identifies split ends, a new RNA recognition motif- containing protein. Genetics 154, 695-712. Rohrbaugh, M., Ramos, E., Nguyen, D., Price, M., Wen, Y., and Lai, Z. C. (2002). Notch activation of yan expression is antagonized by RTK/pointed signaling in the Drosophila eye. Curr Biol 12, 576-581. Staehling-Hampton, K., Ciampa, P. J., Brook, A., and Dyson, N. (1999). A genetic screen for modifiers of E2F in Drosophila melanogaster. Genetics 153, 275-287. 290 Stowers, R. S., and Schwarz, T. L. (1999). A genetic method for generating Drosophila eyes composed exclusively of mitotic clones of a single genotype. Genetics 152, 1631-1639. Sturtevant, M. A., Roark, M., O’Neill, J. W., Biehs, B., Colley, N., and Bier, E. (1996). The Drosophila rhomboid protein is concentrated in patches at the apical cell surface. Dev Biol 174, 298-309. Sullivan, W., Ashburner, M., and Hawley, R. S. (2000). Drosophila Protocols (Cold Spring Harbor, NY, Cold Spring Harbor Laboratory Press). Tsuda, L., Kaido, M., Lim, Y. M., Kato, K., Aigaki, T., and Hayashi, S. (2006). An NRSF/RESTlike repressor downstream of Ebi/SMRTER/Su(H) regulates eye development in Drosophila. Embo J. Tsuda, L., Nagaraj, R., Zipursky, S., and Banerjee, U. (2002). An EGFR/Ebi/Sno Pathway Promotes Delta Expression by Inactivating Su(H)/SMRTER Repression during Inductive Notch Signaling. Cell 110, 625. Wolff, T. (2000). Histological techniques for the Drosophila eye part I: larva and pupa. In Drosophila Protocols, W. Sullivan, M. Ashburner, and R. S. Hawley, eds. (Cold Spring Harbor, NY, Cold Spring Harbor Laboratory Press), pp. 201-227. Wolff, T., and Ready, D. F. (1991). Cell death in normal and rough eye mutants of Drosophila. Development 113, 825-839. Wolff, T., and Ready, D. F. (1993). Pattern formation in the Drosophila retina. In The Development of Drosophila melanogaster, M. Bate, and A. Martinez Arias, eds. (Cold Spring Harbor, N.Y., Cold Spring Harbor Laboratory Press), pp. 1277-1325. Xu, T., and Rubin, G. M. (1993). Analysis of genetic mosaics in developing and adult Drosophila tissues. Development 117, 1223-1237. Yu, S. Y., Yoo, S. J., Yang, L., Zapata, C., Srinivasan, A., Hay, B. A., and Baker, N. E. (2002). A pathway of signals regulating effector and initiator caspases in the developing Drosophila eye. Development 129, 3269-3278. 291 292 Appendix III. Characterization of dominant-negative Spen: in vivo and in vitro David B. Doroquez and Ilaria Rebay 293 Results Over-expression of the Spen C-terminus in the eye (sev-spenDN) acts as a dominant-negative with respect to endogenous Spen function. We also performed genetic screen to identify genetic interactors of spen. We identified genes from the EGFR and Notch signaling pathways as well as genes implicated in chromatin remodeling and programmed cell death. Here, we provide additional insight by identifying genetic interactors of spen through a candidate approach. We also describe our experiments of SpenDN in Drosophila cultured cells. First, we show that spenDN also acts as a dominant-negative during embryonic central nervous system (CNS) development. spenDN expression in CNS midline glial cells acts dominant-negatively spen is required for the proper migration and survival of midline glial cells (MGCs) during Drosophila CNS development (Chen and Rebay, 2000). Expression of Spen that lacks its C-terminus (Spen∆C) acts as a dominant-negative. We asked if the SV40 Large T Antigen Nuclear Localization Signal (NLS)-tagged Spen C-terminus (SpenDN) also functioned as a dominant-negative by analyzing the organization and pattern of longitudinal axons detected by immunostaining for Fascicilin II (Fig. AIII-1). In wild-type stage 16 embryos, two sets of three longitudinal axons run parallel and adjacent to the midline where the MGCs are located (Fig. AIII-1A-C). However, in heterozygous spen mutants that over-express SpenDN driven by the MGC-specific slit-GAL4 driver, we observed aberrant crossing over of the longitudinal axons. (Fig. AIII-1D-F). This phenotype is similar to maternal-zygotic loss-of-function spen embryos and to slit>spen∆C. These data indicate that SpenDN functions as a dominant-negative during CNS development. Although we did not test slit>spenDN in a wild-type spen background, we hypothesize that this genotype would also have CNS defects, but maybe at lower penetrance as was the case for slit>spen∆C. 294 Fig. AIII-1. Over-expression of the Spen C-terminus disrupts embryonic CNS development. FasII immunostaining of embryonic CNS longitudinal axon tracts in (A) wild-type and (B) slit>spenDN, spen-/+ stage 16 embryos. A candidate approach to identify dominant modifiers of the sev-spenDN rough-eye phenotype We used a candidate approach to identify modifiers of the rough eye associated with spenDN. It is important to note in the crosses described here that we have not confirmed for 295 296 table Aiii-1 Dominant modifiers of dominant-negative Spen: Candidate Approach gene/Allele Role sev2B>spenDn sev-spenDn sev-yanACt EGFR ligand mild supp? mild supp? spiP1 P2 EGFR ligand mild supp? mild supp spi EGFR inhibitor mild enh mild/mod supp aos'7 W11 EGFR inhibitor mod/strong enh mild/mod supp aos EGFR inhibitor mild/mod supp aos05845 5352 EGFR MAPK strong enh? rl EGFR MAPK strong enh? rlXS324 A0455 EGFR TF strong enh pnt EGFR TF+ mod/strong enh pntAF397 EGFR TFyan833 ER433 EGFR TFmild enh yan MAPK (P)-ase mild enh dMkp3P1 F02707 MAPK (P)-ase mild enh PTP-ER MAPK (P)-ase mild enh PTP-ERXE-2776 93i Notch/EGFR mild enh sno Notch/EGFR strong supp? ebiE90 l(3)ttkE-Gal TF very strong enh mild supp ttk TF mild supp ttk1 BP2 Notch activator mod supp sca Notch ligand mild/mod supp mod enh? DlB2 Notch receptor mild supp N54l9 Notch TF mild enh Su(H) Df(3R)Espl3 Notch target mild/mod enh E(spl) Notch co-activ. mod supp mamC8 Notch co-repress H1 08269 txn co-repressor mild supp Sin3A txn co-repressor MBDLEY04582 04556 HDAC Rpd3 cell cycle TF Mild/mod supp dpa1 cell cycle TF dpa3 cell cycle TF mild supp dpa4 sev-rasv12 297 table Aiii-1 Dominant modifiers of dominant-negative Spen: Candidate Approach gene/Allele Role sev2B>spenDn sev-spenDn sev-yanACt cell cycle TF mild/mod enh dpa5 i1 cell cycle TF mild/mod enh e2f cell cycle TF mild/mod enh e2fi2 9 cell cycle TF mild/mod supp e2f cyclin mild/mod supp cycB32 2 cyclin mild supp cycB cyclin strong enh cycE1672 4454 CDK inhibitor dap cell cycle TF Rbf14 cell cycle TF Rbf120a 1 RD network TF mod/strong enh dac RD network TF dac3 1-8 Wg ligand mild supp mod supp wg Wg ligand mod supp wg1-12 1 Wg receptor mod enh fz Wg component mild enh dsh6 CT1 TF very mild enh rno TF strong enh rnoCV1 TF mod/strong enh rnoFS1 TF mod enh rnoCM3 nuclear receptor mild/mod enh mild supp svp07842 EY0666 nuclear importin Fs(2)Ket nuclear importin Fs(2)KetKX3 I53 TAF0 mod supp SY3-2 TAF0 mod supp SY3-2AG456 0102 TAF0 mod supp SY3-2 mod enh sev-rasv12 promoter-specific interactions. We first started off by crossing either GAL4 driven or the direct enhancer/promoter driven spenDN (sev2B>spenDN or sev-spenDN, respectively. Interestingly, we found that spi alleles mild suppressed spenDN; we predicted an enhancement based on the observation that spen functions positively with respect to EGFR signaling (Table AIII-1). Other mutants were consistent with our predictions: positive EGFR components would enhance and antagonists of EGFR would suppress. It is unclear, however, how yan alleles interact in this assay; one yan allele had enhanced the spenDN phenotype. We also found that alleles of the Mitogen activated protein kinase (MAPK) phosphatases Drosophila MAPK Phosphatase 3 (dMkp3) and Protein Tyrosine Phosphatase-ERK/enhancer of Ras1 (PTP-ER) showed enhancement; again, we predicted suppression. We also observed genetic interaction with components of the Notch signaling pathway. sca and Notch alleles suppressed sev-spenDN moderately and mildly, respectively (Table AIII-1). This is consistent with Spen as an antagonist of Notch signaling. We observed that a Suppressor of Hairless [Su(H)] loss-of-function allele and an Enhancer of Split [E(spl)] deficiency each enhanced the SpenDN rough-eye. Su(H) is a bi-functional transcription factor so it would have been predicted to interact either way. The E(spl) deficiency used here may also delete other genes required for Spen’s function. Specific E(spl) alleles should be used in this case. We tested spenDN with other genes that may be related to Spen function. For example, spen was implicated in regulation of the cell cycle (Staehling-Hampton et al., 1999) so we tested alleles of cell cycle genes and found a variety of genetic interactors including dE2F and dDP (Table AIII-1). It may be interesting in the future to confirm these genetic interactors to study how spen is involved in the cell cycle or other E2F/DP-required processes. We also tested how over-expression of specific transgenes affected the sev2B>spenDN rough-eye phenotype (Fig. AIII-2B). We do not have a full-length transgene that can over-express Spen because of its size and difficulty in maintaining its propagation in E. coli. Therefore, we over-expressed a spen transgene (minigene) that has a middle portion of Spen deleted (I. Rebay, unpublished results). We find that transgenic expression of spen∆aa3162-4106 suppressed the spenDN 298 299 Fig. AIII-2. Genetic interactions with over-expression of SpenDN in the eye. Scanning electron micrographs of adult eye eyes for (A) Wild-type, (B) sev2B>spenDN, (C) sev2B>spen∆3162-4106, (D) sev2B>spenDN,>spen∆3162-4106, (E) sev2B>spenDN,>Su(H) (F) sev2B>spenDN,>PntP1, (G) sev2B>spenDN,>PntP1, (H) sev2B>spenDN,>PntP2. phenotype (Fig. AIII-2D), but we identified a mild rough-eye phenotype when spen∆aa3162-4106 was over-expressed alone (Fig. AIII-2C). We also observed that over-expression of Su(H) enhanced the spenDN roughe-eye phenotype (Fig. AIII-2E). Consistent with a role for Spen as an antagonist of Notch, we found that NotchDN suppressed spenDN (Fig. AIII-2F). Over-expression of PntP1 and PntP2 both suppressed the small rough-eye of spenDN to the extent neo-plastic growth was observed (Fig. AIII-2G,H). SpenDN may regulate Notch pathway targets in Drosophila cell culture Since spen acts to antagonize Notch signaling in vivo, we were interested to recapitulate this system in vitro using Drosophila Schneider 2 (S2) cultured cells. We first performed Fig. AIII-3. SpenDN regulates Notch pathway transcription targets in Drosophila cell culture. S2 cell transcription assays performed with (A) E(spl)-m8-luciferase and (B) yanO2×5-luciferase reporters. Relative Luciferase Activity (mean + st. dev.) is luciferase/β-gal. 300 transcription assays (Silver et al., 2003) that used enhancer sequences from E(spl)-m8 placed upstream of luciferase (Eastman et al., 1997). This reporter would be hypothesized to be repressed by Spen. We found that this reporter was mildly responsive to activation by Su(H) translationally fused to the Herpes Simplex Virus Protein 16 activator [Su(H)-VP16] and SpenDN alone. However, we find that Su(H)-VP16 and SpenDN synergize to activate transcription (Fig. AIII-3A). We also observe that SpenDN can also regulate a reporter that has two Su(H)-binding sites and a single Ets binding site that are a subset of sequences within the minimal yan enhancer (yanO) (Rohrbaugh et al., 2002). Addition of SpenDN leads to 20-fold expression from the reporter (Fig. AIII-3B). This would indicate that Spen may regulate yan transcriptionally but we have not observed this in vivo. We also have not observed this with actual yan transcript in vitro or Yan protein in vitro in S2 cells (data not shown). In summary, the experiments described above indicate that Spen may regulate Notch targets in cell culture. It may be important to note that there is no Notch expression in S2 cells (Fehon et al., 1990). It would be interesting to introduce activated Notch to see Spen or SpenDN interacts at these reporters. Materials and Methods Fly Strains and Genetics Fly stocks were maintained at 25°C and obtained from the Bloomington Stock Center. w1118 or Canton S were used as control strains, unless otherwise indicated. The following strains were analyzed: For CNS: spenXFM911/+; P[slit-GAL4]/P[UAS-spenDN]3851 301 For eye: P[sev2B-GAL4] P[UAS-spenDN]3851/+ P[UAS-spen∆aa3162-4106]/+ ; P[sev2B-GAL4] P[UAS-spen∆aa3162-4106]/+ ; P[sev2B-GAL4] P[UAS-spenDN]3851/+ P[UAS-Su(H)]3/+; P[sev2B-GAL4] P[UAS-spenDN]3851/+ P[sev2B-GAL4] P[UAS-spenDN]3851/P[UAS-Ndn]6 P[sev2B-GAL4] P[UAS-spenDN]3851/P[UAS-PntP1] P[sev2B-GAL4] P[UAS-spenDN]3851/P[UAS-PntP2] Immunohistochemistry and Histology Immunostaining for longitudinal axons in the CNS was performed with stage 16 embryos as described in Chen et al., 2000 with 1:10 anti-FasII (1D41H9). Flies were prepared for scanning electron microscopy (SEM) by fixation in 1% glutaraldehyde/1% paraformaldehyde in 0.1 M sodium phosphate buffer for 2 hr as described (Tootle et al., 2003). The fixed tissue was dehydrated through an ethanol series. Samples were critical point dried, sputter coated, and images were acquired on a scanning electron microscope (JEOL 5600LV). Fixation and tangential sections of adult eyes was performed as previously described (Tomlinson and Ready, 1987). Transcription Assays Transcription assays were performed as described in (Silver et al., 2003), but E(spl)- m8-luciferase (Eastman et al., 1997) and yanO2×5-luciferase reporters were used. yanO2×5luciferase contains five tandem of copies of a cassette that contains two of the required Su(H) binding sites and one Ets site found within the yanO minimal enhancer (Rohrbaugh et al., 2002). pMT-Su(H)-VP16 and pMT-spenDN were induced with CuSO4 as previously described (Silver et al., 2003). 302 Acknowledgements We would like to thank T.L. Orr-Weaver, R.G. Fehon, P.A. Garrity, J.C. Jemc, P. Vivekanand, S.J. Silver Brown, J.A. Lees and the Fehon, Orr-Weaver, and Rebay labs for stimulating discussion. We are grateful to A. Mukherjee, C. Delidakis, and the ArtavanisTsakonas and Garrity labs for reagents, and to N. Watson (WIBR/Keck Microscopy Facility) for microscopy assistance. D.B.D was supported by a National Science Foundation Graduate Fellowship. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. References Chen, F., and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10, 943-946. Eastman, D. S., Slee, R., Skoufos, E., Bangalore, L., Bray, S., and Delidakis, C. (1997). Synergy between suppressor of Hairless and Notch in regulation of Enhancer of split m gamma and m delta expression. Mol Cell Biol 17, 5620-5628. Fehon, R. G., Kooh, P. J., Rebay, I., Regan, C. L., Xu, T., Muskavitch, M. A., and ArtavanisTsakonas, S. (1990). Molecular interactions between the protein products of the neurogenic loci Notch and Delta, two EGF-homologous genes in Drosophila. Cell 61, 523-534. Rohrbaugh, M., Ramos, E., Nguyen, D., Price, M., Wen, Y., and Lai, Z. C. (2002). Notch activation of yan expression is antagonized by RTK/pointed signaling in the Drosophila eye. Curr Biol 12, 576-581. Silver, S. J., Davies, E. L., Doyon, L., and Rebay, I. (2003). Functional dissection of eyes absent reveals new modes of regulation within the retinal determination gene network. Mol Cell Biol 23, 5989-5999. Staehling-Hampton, K., Ciampa, P. J., Brook, A., and Dyson, N. (1999). A genetic screen for modifiers of E2F in Drosophila melanogaster. Genetics 153, 275-287. Tomlinson, A., and Ready, D. F. (1987). Neuronal differentiation in the Drosophila ommatidium. Developmental Biology 120, 366-376. Tootle, T. L., Lee, P. S., and Rebay, I. (2003). CRM1-mediated nuclear export and regulated activity of the Receptor Tyrosine Kinase antagonist YAN require specific interactions with MAE. Development 130, 845-857. 303 304 Appendix IV. Phenotypic analysis of Spen over-expression David B. Doroquez and Ilaria Rebay 305 Results Our analysis of Split ends (Spen) during Drosophila development has involved studying spen loss-of-function alleles or dominant-negative transgenes in order to imply roles for Spen function. Here, we describe how over-expression of Spen in the eye affects eye patterning. A spen EP line that allows for Spen over-expression Because large, full-length constructs of spen have not been successfully propagated in bacteria (I. Rebay, unpublished) we could not generate a transgene to over-express Spen. However, Enhancer-Promoter (EP) P-elements may provide a solution. EP elements are transposable elements that contain Upstream Activating Sequences (UAS) and promoters that when bound by the transcriptional activator GAL4 will drive ectopic expression of adjacent genes (Rorth, 1996). The bias for P-elements to insert at the 5’ end genes (reviewed in Spradling et al., 1999) makes the EP P-elements as an ideal tool for targeted gene over-expression. The Drosophila Gene Search Project (DGSP) organized by the Tokyo Metropolitan University has performed insertional mutagenesis to establish 20,000 EP lines. We screened several EP lines for their ability to over-express Spen when crossed to a tissue-specific GAL4 driver. We identified the spenGS2279 allele as an EP allele that allowed for Spen over-expression. The EP element GSV1 spenGS2279 is inserted -128 bp upstream of the transcriptional start site (Fig. AIV-1A) and is likely to permit ectopic expression for two of the three Spen isoforms. It is important to note that this GSV1 EP element has the ability to drive expression of genes flanking it in either direction so other genes may be affected. We crossed spenGS2279 to a GAL4 driver that drives expression of UAS target genes in the embryonic stripe pattern of the segment polarity gene patched (ptc) and immunostained for Spen protein. We found that Spen protein is expressed in stripes in ptc>spenGS2279 stage 16 embryos (Fig. AIV-1B). This suggests that this EP allows for Spen ectopic expression. 306 Fig. AIV-1. The EP allele spenGS2279 allows for Spen over-expression. (A) Schematic of the insertion point for the GSV1 EP P-element transposon with respect to the spen gene region. (B) Stage 16 embryo showing ectopic striped expression of Spen driven by ptc-Gal4 from spenGS2279 EP. Eye-specific expression of the spen EP disrupts eye development We were interested in understanding the consequences of Spen over-expression in the eye so we crossed spenGS2279 to a set of eye-specific drivers that have varying strengths and scope of targeted cells. The spenGS2279 allele is homozygous viable and does not appear to disrupt eye patterning (Fig. AIV-3B). However, when driven with GMR2-GAL4, which drives expression of 307 transgenes posterior to all cell posterior to the morphogenetic furrow (MF), we observed severe disruption of eye patterning causing the eye to be glassy in appearance (Fig. AIV-2B, 3B). We also saw a similar but graded effect when we tested spenGS2279 allele with a set of sev-GAL4 drivers that have strong to weak driving potential (Fig. AIV-2C,D; 3D-F). This suggests that spen over-expression during eye development perturbs eye patterning. We wanted to understand the nature of the spenGS2279 allele so we asked if spenGS2279 expression can suppress the rough-eye associated with dominant-negative spen (spenDN) (Fig. AIV-2E, 3G). We find that co-expression of spenGS2279 and spenDN with the sev2B-GAL4 driver leads to suppression of the spenDN rough-eye phenotype (Fig. AIV-2F, 3H). These results suggest that spenGS2279 may rescue loss of spen function. We then asked whether reduction in the endogenous levels of spen would suppress the activity of spenGS2279. However, heterozygous reduction of spen led to exacerbation of the rough-eye phenotype associated with sevK25>spenGS2279 (data not shown). This result suggests that spenGS2279 may function as a dominant-negative. Therefore, the analysis of the nature of spenGS2279 may be complicated because it suppressed a dominant-negative transgene but also acted dominant-negatively when spen levels were reduced. Spen over-expression leads to reduced Yan expression and ectopic BarH1 expression In spen mutant clones, Yan protein is elevated compared to wild-type tissue. We therefore asked whether over-expression of Spen would lead to reduced levels of Yan. In wildtype, Yan protein is expressed strongly at the MF and reduced basally posterior to the MF (Fig. AIV-4A). In sevK25>spenGS2279 eye discs Yan expression is lost more posterior to the MF (Fig. AIV4B), confirming Spen’s antagonism of Yan. We also find that Yan expression is mildly elevated around the cells where loss of Yan occurs in the posterior of the disc in sevK25>spenGS2279 eye discs. We currently do not understand the significance of this mild upregulation of Yan. We also wanted to see how spenGS2279 affected differentiation during eye development. There may be a subtle failure for photoreceptor R cells to differentiate in sevK25>spenGS2279 308 Fig. AIV-2. SpenGS2279 over-expression disrupts eye patterning but suppresses SpenDN (SEM). Scanning electron micrographs (SEM) of adult eyes for (A) wild-type, (B) GMR2>spenGS2279, (C) sevK25>spenGS2279, (D) sev2B>spenGS2279, (E) sev2B>spenDN, and (F) sev2B>spenDN,>spenGS2279. eye discs (Fig. AIV-4D) compared to wild-type (Fig. AIV-4C). We also looked at BarH1 a differentiation marker for photoreceptors R1 and R6 (Fig. AIV-4C) (Higashijima et al., 1992). In addition to this pattern, we observed that BarH1 is induced in interommatidial cells in sevK25>spenGS2279 eye discs (Fig. AIV-4D). BarH1 is known to be activated by Hedgehog (Hh) 309 310 Fig. AIV-3. SpenGS2279 over-expression disrupts eye patterning but suppresses SpenDN (Sections). Tangential sections of adult eyes for (A) wild-type, (B) spenGS2779, (C) GMR2>spenGS2779, (D) sevK25>spenGS2279, (E) sev1B>spenGS2279, (F) sev2B>spenGS2279, (G) sev2B>spenDN, and (H) sev2B>spenDN,>spenGS2279. and EGFR signaling in the eye where BarH1 acts as an anti-proneural transcription factor (Lim and Choi, 2003; Lim and Choi, 2004). It is a possibility that Spen over-expression may induce BarH1 anti-proneural function to limit differentiation of photoreceptors. Fig. AIV-4. SpenGS2279 over-expression in eye inhibits Yan but induces BarH1 expression. Third-instar eye imaginal discs staining for (A,B) Yan or (C,D) Elav (in blue) and BarH1 (in red) for (A,C) wild-type and (B,D) sevK25>spenGS2279. Discs are oriented anterior to the left. 311 Materials and Methods Fly Strains and Genetics Fly stocks were maintained at 25°C and obtained from the Bloomington Stock Center. w1118 or Canton S were used as control strains, unless otherwise indicated. The following strains were analyzed: For embryo: ptcP[GAL4]/ spenP[GS2279] For eye: P[GMR-GAL4]/spenP[GS2279] spenP[GS2279]/+; P[sevK25-GAL4]/+ spenP[GS2279]/+; P[sevK25-GAL4]/+ spenP[GS2279]/+; P[sev1B-GAL4]/+ spenP[GS2279]/+; P[sev2B-GAL4]/+ P[sevK25-GAL4]/P[sev-spenDN]3851 spenP[GS2279]/+; P[sevK25-GAL4]/P[sev-spenDN]3851 Histology, Immunohistochemistry and Immunofluorescence Flies were prepared for scanning electron microscopy (SEM) by fixation in 1% glutaraldehyde/1% paraformaldehyde in 0.1 M sodium phosphate buffer for 2 hr as described (Tootle et al., 2003). The fixed tissue was dehydrated through an ethanol series. Samples were critical point dried, sputter coated, and images were acquired on a scanning electron microscope (JEOL 5600LV). Fixation and tangential sections of adult eyes was performed as previously described (Tomlinson and Ready, 1987). Eye imaginal discs were dissected from wandering third instar larvae, fixed with 4% paraformaldehyde in PBT (0.1% Triton X-100 in Phosphate-Buffered Saline [PBS]) for 15 min, 312 washed three times in PBT (5 min each), incubated in primary antibody overnight in PBT with 5% normal goat serum (PNT) at 4ºC, and then washed three times in PBT (5 min each). Discs were incubated in secondary antibody for 2 hr in PNT at room temperature, washed three times in PBT (5 min each), and mounted in Vectashield mountant (Vector Laboratories). Microscopy and image acquisition was performed on a Zeiss LSM510 confocal microsope. For immunofluorescence, the following primary antibodies were used: 1:200 mouse anti- Yan (8B12H9, Developmental Studies Hybridoma Bank [DSHB]), 1:50 rat anti-Elav (7E8A10, DSHB), and 1:500 rabbit anti-BarH1 (Higashijima et al., 1992). Secondary antibodies were Alexa Fluor 568-conjugated goat anti-mouse IgG (Molecular Probes), or appropriate Cy3/ Rhodamine Red-X- and Cy5-conjugated antibodies (Jackson ImmunoResearch). Acknowledgements We would like to thank R.G. Fehon, P.A. Garrity, J.C. Jemc, P. Vivekanand, J.A. Lees and the Fehon, Orr-Weaver, and Rebay labs for stimulating discussion. We are grateful to A. Mukherjee, and the Artavanis-Tsakonas and Garrity labs for reagents, and to N. Watson and E. Batchelder (WIBR/Keck Microscopy Facility) for microscopy assistance. D.B.D was supported by a National Science Foundation Graduate Fellowship. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. References Higashijima, S., Kojima, T., Michiue, T., Ishimaru, S., Emori, Y., and Saigo, K. (1992). Dual Bar homeo box genes of Drosophila required in two photoreceptor cells, R1 and R6, and primary pigment cells for normal eye development. Genes Dev 6, 50-60. Lim, J., and Choi, K. W. (2003). Bar homeodomain proteins are anti-proneural in the Drosophila eye: transcriptional repression of atonal by Bar prevents ectopic retinal neurogenesis. Development 130, 5965-5974. 313 Lim, J., and Choi, K. W. (2004). Induction and autoregulation of the anti-proneural gene Bar during retinal neurogenesis in Drosophila. Development 131, 5573-5580. Rorth, P. (1996). A modular misexpression screen in Drosophila detecting tissue-specific phenotypes. Proc Natl Acad Sci U S A 93, 12418-12422. Spradling, A. C., Stern, D., Beaton, A., Rhem, E. J., Laverty, T., Mozden, N., Misra, S., and Rubin, G. M. (1999). The Berkeley Drosophila Genome Project gene disruption project: Single P-element insertions mutating 25% of vital Drosophila genes. Genetics 153, 135-177. Tomlinson, A., and Ready, D. F. (1987). Neuronal differentiation in the Drosophila ommatidium. Developmental Biology 120, 366-376. 314 Appendix V. Characterization of dominant-negative Spen∆C during Drosophila eye development and in Drosophila cultured cells David B. Doroquez and Ilaria Rebay 315 Results Over-expression of Spen that lacks its C-terminus functions as a dominant-negative during Drosophila central nervous system (CNS) development (Chen and Rebay, 2000). We were interested in identifying how Spen∆C behaved with respect to eye development and Epidermal Growth Factor Receptor (EGFR) signaling. Here, we report that Spen∆C overexpression has properties similar to spen loss-of-function in the eye. We also report that Drosophila cell culture analysis suggests that Spen∆C adversely affects EGFR/Pnt activation of transcription targets. Over-expression of Spen∆C results in elevated Yan protein levels Our examination of spen loss-of-function indicates that output of from the EGFR pathway is reduced in spen mutant clones. This is in part due to elevated levels of Yan protein in these clones. Over-expression of SV40 Large T Antigen Nuclear Localization Signal (NLS)tagged Spen C-terminus (SpenDN) genetically functions as a dominant-negative but we have not observed at the molecular level whether SpenDN leads to elevated Yan protein levels (data not shown). We decided to determine if Spen∆C had any affect on Yan protein. Over-expression of SpenDN under the GMR-GAL4 leads to a slightly rougher eye compared to the rough-eye associated with GMR-GAL4 alone (data not shown). We analyzed Yan expression for wild-type and GMR>spen∆C. In wild-type Yan expressed strongly in basal cells and reduced posterior (Fig. AV-1A). In GMR>spen∆C eye discs, we observed strong basal-cell expression of Yan (Fig. AV1C) and more Yan-expressing cells at the apical level of the eye disc (Fig. AV-1D). These results suggest that Spen∆C recapitulates the upregulated-Yan phenotype associated with spen loss-offunction. Spen∆C does not regulate Yan at the post-transcriptional level via the yan 3’UTR 316 Our investigations suggest that Spen inhibits Yan post-transcriptionally likely Fig. AV-1. Spen∆C leads to elevated Yan protein levels but not to a noticeable change in tubGFP-yanWT3’UTR reporter expression. (A,C) Basal and (B,D) apical confocal micrograph projections of (A,B) ‘wild-type’ control and (C,D) GMR>spen∆C third-instar eye imaginal discs with tub-GFP-yanWT3’UTR transgenic reporter in background. 317 by mediating activation of activated mitogen activated protein kinase (actMAPK)/dualphosphorylated Extracellular signal Regulated Kinase (dpERK) (see Chapter 3). dpERK in turn directly inhibits Yan repressor function (Rebay and Rubin, 1995). Another possibility is that Spen may function to regulate yan transcript by affecting the yan mRNA 3’UTR possibly by direct-binding or regulation of a micro-RNA such as miR-7 (Li and Carthew, 2005). In such a model, Spen would play a role in inhibiting translation from the yan transcript. Therefore, we tested a GFP-yan3’UTR reporter in discs that over-express Spen∆C and asked if Spen∆C induced expression of this reporter. In this heterologous system, if Spen regulates the yan 3’UTR, dominant-negative Spen∆C would lead to de-repression from the reporter. We did not observe a noticeable change in GFP levels (Fig. AV-1C,D). We also tested Spen∆C’s ability to regulate a luciferase reporter that includes the yan 3’UTR (Li and Carthew, 2005) in Drosophila S2 cultured cells. Addition of Spen∆C did not affect the luciferase reporter. miR-7 a miRNA that inhibits translation from yan transcript by binding to the yan 3’UTR (Li and Carthew, 2005), however, limits luciferase levels measured from the luci-yan3’UTR reporter (Fig. AV-2A). Ectopic over-expression of miR-7 driven by the GMRGAL4 driver leads to significant loss of Yan posterior to the morphogenetic furrow (MF) (Fig. AV-2B) (Li and Carthew, 2005). We find that spen suppresses miR-7’s ability to inhibit Yan levels (Fig. AV-2C). It is unclear if Spen then is required for miR-7 post-transcriptional repression of Yan. But, since Spen did not regulate yan post-transcriptionally at the 3’UTR level, it is likely that suppression effect on GMR>miR-7 is occurring from a parallel pathway. For example, the Yan protein that is generated posterior to the MF in GMR>miR-7 discs may not be repressed by dpERK in the GMR>spen∆C background. Spen affects activation of Ets targets Spen is a positive regulator of EGFR signaling (Chen and Rebay, 2000)(see Chapter 3). In this thesis, we reported down-regulation of EGFR pathway targets such as argos (aos) in spen mutant discs (see Chapter 3). We wanted to study this effect of Spen’s activity in vitro in 318 Fig. AV-2. Spen∆C does not directly affect GFP-yan3’UTR levels but suppresses ectopic miR7’s repression of endogenous Yan protein. (A) S2 cell transcription assays were performed with tub-GFP-yanWT3’UTR-luciferase reporter. Relative Luciferase Activity represents luciferase/β-gal activity (mean + st. dev.). (B,C) Confocal micrographs of third instar eye imaginal discs staining for Elav (blue) and Yan (red). (B) Over-expression of miR-7 under the control of GMR--GAL4 driver leads to loss of Yan expression posterior to the MF. (C) Coexpression of Spen∆C suppresses the loss of Yan observed in (B). 319 Fig. AV-3. Spen∆C modulates activity of PntP1 and may antagonize RasV12’s inhibition of Yan repression. S2 cell transcription assays were performed with 6×EBS-luciferase reporter. Relative Luciferase Activity represents luciferase/β-gal activity (mean + st. dev.). S2 cells to determine if any further insight can be made into Spen function. We used a luciferase that contains 6× Ets-binding sites (EBS) upstream of the luciferase transcriptional start site (O’Neill et al., 1994). Pointed-P1 (PntP1) strongly activates this reporter, but Spen suppresses its activation (Fig. AV-3). We have not determined how Yan protein levels are affected upon Spen∆C transfection. It is a possibility that Spen may lead to maintenance of endogenous protein levels. We also found that Spen∆C does not further enhance the repression by transfected Yan, but Spen may regulate RasV12’s ability to inhibit Yan’s transcriptional repression (Fig. AV-3). In summary, our results indicate that Spen positively regulates the activation of EGFR/Ets targets in Drosophila cultured cells. 320 table Av-1A yan nuclear export assays with Spen'C (S2 cells) Condition n %nuc %nuc+Cyt Yan 1312 71% 15% V2 Yan+Ras 653 2% 3% V2 856 1% 3% Yan+Ras +Spen'C %Cyt 15% 95% 96% table Av-1B yan nuclear export assays with Spen RnAi (S2 cells) Condition n %nuc %nuc+Cyt %Cyt Yan 574 90% 1% 9% Yan+RasV2 531 5% 6% 90% Yan+RasV2+Spen RNAi 657 4% 9% 87% table Av-2 time-course yan nuclear export assays with Spen'C (S2 cells) Condition time n %nuc %nuc+Cyt %Cyt Yan 0:00 316 97% 2% 1% Yan+RasV2 0:00 405 77% 6% 17% 0:00 369 81% 10% 9% Yan+RasV2+Spen'C Yan :0 2 98% 1% 1% Yan+RasV2 :0 7 76% 4% 20% :0 2 63% 5% 32% Yan+RasV2+Spen'C Yan 3:00 308 98% 1% 1% V2 Yan+Ras 3:00 462 12% 8% 80% V2 3:00 336 15% 10% 75% Yan+Ras +Spen'C Yan :0 9 99% 1% 1% V2 Yan+Ras :0 0 9% 2% 90% V2 :0 13% 8% 79% Yan+Ras +Spen'C Yan 6:00 342 99% 0% 1% V2 6:00 285 7% 7% 86% Yan+Ras V2 6:00 297 8% 3% 88% Yan+Ras +Spen'C Spen does not affect Yan nuclear export mediated by activated RasV12 in S2 cells Upon EGFR pathway stimulation and phosphorylation of Yan by dpERK, Yan is exported from the nucleus where it is normally degraded. In S2 cells, this degradation does not occur, thus Yan export from the nucleus may be measured (Rebay and Rubin, 1995). We were interested in determining whether Spen played a role in regulating Yan localization. If Spen is an antagonist of Yan nuclear localization, over-expression of Spen∆C may inhibit RasV12’s ability to promote 321 Yan nuclear export. We therefore performed Yan export assays (Rebay and Rubin, 1995; Tootle et al., 2003) to determine Spen’s role. We find that Yan nuclear export mediated by RasV12 is not affected with expression of Spen∆C or with spen RNA interference (Table AV-1A). We then asked whether cells that have RasV12 and Spen∆C have slower export than with just RasV12 alone by performing Yan nuclear export assays during a time-course. We did not observe any significant change in rate of Yan nuclear export (Table AV-2). These results suggest that Spen does not regulate Yan nuclear export in Drosophila S2 cells. spen RNAi knock-down does not affect Yan protein levels but dpERK levels are elevated spen mutant eye clones have elevated Yan and reduced dpERK levels (see Chapter 3). We were interested if a similar situation occurred in Drosophila S2 cultured cells. We knockeddown spen transcript levels via addition of spen dsRNA to S2 cells (Fig. AV-4). We did not observe a significant change in Yan protein levels normalized to β-tubulin (Fig. AV-4), suggesting Fig. AV-4. spen RNAi knockdown does not affect Yan but increases dpERK levels in cell culture. (A) Semi-quantitative RT-PCR of spen and RpS17 was performed for conditions with or without spen dsRNA-treatment of S2-R+ cells. (B,C) Protein immunoblot was performed to detect (B) Yan and (C) dpERK in conditions with or without spen dsRNA-treatment of S2-R+ cells. β-Tubulin is was blotted for loading control. 322 Spen does not regulate endogenous Yan in this system. Surprisingly, we observed upregulation of dpERK levels with spen knock-down which we compare with the reduction we observed in vivo (Fig. AV-4). These results suggest that there may tissue specific differences between in vivo eye imaginal discs and S2 cells that do not permit us to study Spen’s function with respect to Yan and dpERK. Alternatively, there may be defects in the localization of dpERK such that it is not properly imported into the nucleus where it phosphorylates targets and is quickly degraded in a spen knock-down background. It would be interesting to determine how importins such as Drosophila Importin 7 (DIM-7)/Moleskin (Msk) and Fs(2)Ketel are regulated with loss of spen. If spen regulates importins at the transcript level, Msk and Ketel transcripts would be predicted to be reduced and over-expression of Msk and Ketel would be predicted to rescue the dpERK elevated expression observed with spen knock-down. Preliminary data indicates that Msk transcript levels, however, are elevated in spen mutant eye discs (data not shown). Ketel transcript levels still need to be measured in spen mutant discs. Materials and Methods Fly Strains and Genetics Fly stocks were maintained at 25°C and obtained from the Bloomington Stock Center. w1118 or Canton S were used as control strains, unless otherwise indicated. The follwing strains were analyzed: P[tub-GFP-yanWT3’UTR]/P[GMR-GAL4] (Li and Carthew, 2005) P[tub-GFP-yanWT3’UTR]/P[GMR-GAL4]; P[UAS-spen∆C]DBD154/+ P[GMR-GAL4]/P[UAS-miR7] (Li and Carthew, 2005) P[GMR-GAL4]/P[UAS-miR7]; P[UAS-spen∆C]DBD154/+ 323 Luciferase-UTR/Transcription Assays Luciferase-UTR assays were performed as described in (Silver et al., 2003), but tub- luciferase-yanWT3’UTR (Li and Carthew, 2005) was used as a reporter and were co-transfected with pMT-spen∆C or ubi-GAL4/UAS-miR-7 (Li and Carthew, 2005). Transcription assays were performed similarly, but 6×EBS-luciferase (O’Neill et al., 1994) was used as a reporter and was co-transfected with pMT-PntP1, pMT-spen∆C, pMT-yan, and/or pMT-RasV12. Immunohistochemistry and Immunofluorescence Eye imaginal discs were dissected from wandering third instar larvae, fixed with 4% paraformaldehyde in PBT (0.1% Triton X-100 in Phosphate-Buffered Saline [PBS]) for 15 min, washed three times in PBT (5 min each), incubated in primary antibody overnight in PBT with 5% normal goat serum (PNT) at 4ºC, and then washed three times in PBT (5 min each). Discs were incubated in secondary antibody for 2 hr in PNT at room temperature, washed three times in PBT (5 min each), and mounted in Vectashield mountant (Vector Laboratories). Microscopy and image acquisition was performed on a Zeiss LSM510 confocal microsope. For immunofluorescence, the following primary antibodies were used: 1:200 mouse anti-Yan (8B12H9, Developmental Studies Hybridoma Bank [DSHB]) and 1:50 rat anti-Elav (7E8A10, DSHB). RNAi interference and semi-quantitative RT-PCR spen dsRNA was generated with primers with T7 RNA polymerase sites (in italics) at the 5’ end using the Ambion Megascript Kit. Primers are as follows: spen dsRNA Forward 5’-gaattaatacgactcactatagggagagtgttctggtgaaaggaaacgc-3’ spen dsRNA Reverse 5’-gaattaatacgactcactatagggagacgtgaggaggcatttgtatctgag-3’ RNAi soaking was performed by resuspending S2-R+ cells at 1×106 cells/mL in Schneider Insect Cell Medium (GibcoBRL) (-) FBS at plating in 6-well tissue culture plates. Cells were treated with 20 µg dsRNA, incubated with dsRNA at RT for 30 min , and then 3 mL 324 Medium (+) FBS was added to each well. Cells were incubated at RT cell culture incubator before harvest. For harvesting, total RNA was isolated with Trizol reagent (Invitrogen). Total RNA was isolated using Trizol Reagent (Invitrogen) according to the manufacturer’s protocol, treated with DNase I (Invitrogen), and used as template to generate cDNA using the Random Primers provided in the Reverse Transcription System (Promega). PCR was performed: initial denaturation 1 min 94°C, 30 PCR cycles (94°C 30 sec, 55°C 30 sec, 72°C 1 min), 5min final elongation at 72°C. PCR primers are as follows: spen Forward 5’-aagtccgtttatcggtcgtg-3’ spen Reverse 5’-atcctcagcagcggtgtatt-3’ Rps17 Forward 5’-ttggacttccacaccaacaa-3’ RpS17 Reverse 5’-gtcgacctcgatgatgtcct-3’ Yan Nuclear Export Assays Yan nuclear export assays were performed as described (Rebay and Rubin, 1995; Tootle et al., 2003). The transfection constructs included pMT-yan, pMT-RasV12, and pMT-spen∆C. Time-course assays were performed using pPAc-yan, pMT-RasV12, pUbi-spen∆C, and pMT-spen∆C transfection constructs. Western Blot Analysis Drosophila S2-R+ cells were mixed with 2× SDS sample buffer and boiled. Protein samples were resolved on 8% SDS-PAGE gels and transferred by semi-dry blot to a PVDF membrane (Millipore Immobilon P). Membranes were blocked and probed with 1:200 mouse anti-Yan antibody, 1:10,000 mouse anti-dpERK, 1:400 rat anti-β-tubulin (YL1/2, YL1/34, Harlan Seralab) in PBS with 5% non-fat milk (Block Buffer) overnight at 4˚C. Washes were performed in TBS with 0.15% Tween-20 (TTBS) (3×10 min each). 325 Acknowledgements We would like to thank T.L. Orr-Weaver, R.G. Fehon, L.A. Perkins, P.A. Garrity, J.C. Jemc, P. Vivekanand, S.J. Silver Brown, J.A. Lees and the Fehon, Orr-Weaver, and Rebay labs for stimulating discussion. We are grateful to A. Mukherjee, and the Artavanis-Tsakonas and Garrity labs for reagents, and to N. Watson (WIBR/Keck Microscopy Facility) for microscopy assistance. D.B.D was supported by a National Science Foundation Graduate Fellowship. This work was supported in part by National Institutes of Health grant R01 EY12549 to I.R. References Chen, F., and Rebay, I. (2000). split ends, a new component of the Drosophila EGF receptor pathway, regulates development of midline glial cells. Curr Biol 10, 943-946. Li, X., and Carthew, R. W. (2005). A microRNA Mediates EGF Receptor Signaling and Promotes Photoreceptor Differentiation in the Drosophila Eye. Cell 123, 1267-1277. O’Neill, E. M., Rebay, I., Tjian, R., and Rubin, G. M. (1994). The activities of two Ets-related transcription factors required for Drosophila eye development are modulated by the Ras/ MAPK pathway. Cell 78, 137-147. Rebay, I., and Rubin, G. M. (1995). Yan functions as a general inhibitor of differentiation and is negatively regulated by activation of the Ras1/MAPK pathway. Cell 81, 857-866. Silver, S. J., Davies, E. L., Doyon, L., and Rebay, I. (2003). Functional dissection of eyes absent reveals new modes of regulation within the retinal determination gene network. Mol Cell Biol 23, 5989-5999. Tootle, T. L., Lee, P. S., and Rebay, I. (2003). CRM1-mediated nuclear export and regulated activity of the Receptor Tyrosine Kinase antagonist YAN require specific interactions with MAE. Development 130, 845-857. 326 327 328