Photoreceptor and cone cells in the Drosophila eye are recruited following activation of the epidermal growth factor receptor (EGFR) pathway. We have identified echinoid (ed) as a novel putative cell adhesion molecule that negatively regulates EGFR signaling. The ed mutant phenotype is associated with extra photoreceptor and cone cells. Conversely, ectopic expression of ed in the eye leads to a reduction in the number of photoreceptor cells. ed expression is independent of EGFR signaling and ED is localized to the plasma membrane of every cells throughout the eye disc. We present evidence that ed acts nonautonomously to generate extra R7 cells by a mechanism that is sina-independent but upstream of Tramtrack (TTK88). Together, our results support a model whereby ED defines an independent pathway that antagonizes EGFR signaling by regulating the activity, but not the level, of the TTK88 transcriptional repressor.

The Drosophila compound eye is composed of approximately 800 ommatidia, each of which contains eight photoreceptor cells (R1-R8), four non-neuronal cone cells, and eight accessory cells arranged in a highly ordered pattern. In third instar larvae, patterning of the ommatidial field begins at the posterior margin of the eye imaginal disc, with the morphogenetic furrow sweeping across the disc epithelium in a posterior to anterior direction. Behind the furrow, different cell types are recruited sequentially. The R8 photoreceptor is the first cell to differentiate, followed by R2/R5, R3/R4, R1/R6 and R7. Addition of cone and accessory cells to the photoreceptor cluster produces the final ommatidial unit (reviewed in Wolff and Ready, 1993). The Drosophila epidermal growth factor receptor (EGFR) is required for differentiation of all the cell types, with the exception of R8 (Freeman, 1996; Dominguez et al., 1998). A model describing the reiterative activation of the EGFR by the opposing action of the EGFR ligand, Spitz (SPI) and an antagonist, Argos (AOS), has been proposed to explain the successive recruitment of each cell type in the developing eye (Freeman, 1997).

The Drosophila EGFR signaling pathway is subject to modulation at multiple levels by various positive and negative mechanisms (reviewed in Perrimon and Perkins, 1997; Wasserman and Freeman, 1997; Schweitzer and Shilo, 1997). There are multiple EGFR ligands, SPI, Vein and Gurken that activate the receptor tyrosine kinase (RTK) at different stages of development. In addition, Rhomboid (RHO), a multiple transmembrane domain protein, can potentiate EGFR signaling by regulating SPI processing and presentation (reviewed in Wasserman and Freeman, 1997; Bang and Kintner, 2000). The activated EGFR triggers a conserved signal transduction cascade that includes DRK, SOS, RAS1, KSR, RAF, MAPK (reviewed by Perrimon and Perkins, 1997). While SOS activates RAS1 by promoting the exchange of GDP for GTP, the GTP-activating protein, GAP1, inactivates RAS1 by stimulating its intrinsic GTPase activity (Gaul et al., 1992). Furthermore, activated MAPK is thought to propagate the RAS cascade signal into the nucleus by phosphorylating two members of the ETS family of transcription factors, YAN (AOP – FlyBase) and Pointed (PNT) (Brunner et al., 1994a; Brunner et al., 1994b; O’Neill et al., 1994). PTP-ER, a cytoplasmic tyrosine phosphatase, binds to and downregulates activated MAPK (Karim and Rubin, 1999). Therefore, both GAP1 and PTP-ER act in the cytoplasm to negatively regulate EGFR signaling. YAN is a transcriptional repressor that inhibits the production of both photoreceptor and cone cells (Lai and Rubin, 1992). Upon phosphorylation, YAN moves to the cytoplasm for degradation, leading to the differentiation of both cell types (Rebay and Rubin, 1995). In addition, tramtrack (ttk), which encodes two alternatively spliced, zinc finger proteins, TTK69 and TTK88 (Read and Manley, 1992; Xiong and Montell, 1993), plays a central role in photoreceptor cell differentiation. TTK88 functions as a transcriptional repressor to inhibit photoreceptor but not cone cell differentiation (Lai et al., 1996; Yamamoto et al., 1996). Upon EGFR activation, two other nuclear proteins, Phyllopod (PHYL), which is transcriptionally regulated by the RAS/MAPK pathway, and Seven In Abstentia (SINA), form a complex that binds to TTK88. This association leads to the targeting of TTK88 for degradation which results in the differentiation of photoreceptor cells (Tang et al., 1997; Li et al., 1997).

In addition to the cytoplasmic (GAP1, PTP-ER) and nuclear (YAN and TTK88) repressors, EGFR signaling can also be down-regulated by the production of AOS, Kekkon 1 (KEK1), and Sprouty (STY), via negative feedback loops following EGFR activation. AOS, a secreted molecule functions nonautonomously to repress EGFR signaling by blocking both the receptor dimerization and the ability of SPI to bind the EGFR (Freeman et al., 1992; Schweitzer et al., 1995; Jin et al., 2000). KEK1, an adhesion molecule protein, might execute its inhibitory effect through direct association of its extracellular and transmembrane domain with the EGFR (Ghiglione et al., 1999). Finally, the intracellular but inner membrane-associated STY has been proposed to act through its direct binding to DRK and GAP1, thereby blocking the activation of RAS1 (Casci et al., 1999).

Studies of the EGFR have also revealed that heterologous pathways can also modulate the activity of this RTK (reviewed in Moghal and Sternberg, 1999; Tan and Kim, 1999). For example, when growth hormone binds to its receptor, it activates JAK2, which promotes tyrosine phosphorylation of a GRB2 binding site on the EGFR, thereby leading to the activation of RAS and MAPK (Yamauchi et al., 1997). In addition, occupied and aggregated integrins can collaborate with growth factors by triggering tyrosine phosphorylation of EGFR (Miyamoto et al., 1996; Moro et al., 1998). In contrast, extracellular matrix proteins such as collagen, can promote retinoid-induced differentiation of normal human bronchial epithelial cells, by reducing the level of EGFR-dependent MAPK activation (Moghal and Neel, 1998). Thus, molecules involved in cell adhesion can act either positively or negatively to regulate EGFR signaling.

In this report, we describe the identification of the gene echinoid (ed) as a novel negative regulator of EGFR signaling. Ommatidia of ed mutant flies contain extra photoreceptor and cone cells. In contrast, ectopic expression of ed in the eye leads to a reduction in the number of photoreceptors. We show that ed is not transcriptionally regulated by EGFR signaling and that it encodes a putative cell adhesion protein which is primarily localized to the plasma membrane of every cells throughout the eye disc. Our genetic analyses demonstrate that ed acts nonautonomously to generate extra R7 cells by a mechanism that is sina-independent but upstream of TTK (TTK88). Together, our results support a model whereby ED defines an independent pathway that antagonizes EGFR signaling by regulating the activity of the TTK88 transcriptional repressor.

Genetics

The stocks used were edslH8, edslA12, edlF20 (de Belle et al., 1993); l(2)k01102 (Torok et al., 1993); ElpB1 (Baker and Rubin, 1989); Gap1B2 (Chou et al., 1993); sose4G (Simon et al., 1991); hs-aos (Freeman, 1994); rlSEM (Brunner et al., 1994); sevd2 (Simon et al., 1991); sev-RasV12 (Fortini et al., 1992); sev-RasN17 (Karim et al., 1996); sev-tor4021Egfr (Reichman-Fried et al., 1994); sev-tor4021Raf (Dickson et al., 1992); pnt1277, pntΔ88 (Brunner et al., 1994); sevyanACT (Rebay and Rubin, 1995); sina2, sina3 (Carthew and Rubin, 1990); ttk0219 (Lai et al., 1996); GMR-ttk88 (Tang et al., 1997); sproutyS88 (Casci et al., 1999); UAS-kek1 (Ghiglione et al., 1999); hsrho (Dominguez et al., 1998); GMR-GAL4 (Freeman, 1996); and MS1096-GAL4 (Capdevila and Guerrero, 1994).

Production of mosaic clones

Mitotic clones in the eye were induced by using ey-FLP (kindly provided by B. Dickson). The genotype of these flies was w,sevd2/Y; edslA12,FRT40A/ P[w+]30C, FRT40A; ey-FLP/+.

Molecular biology

A 8 kb EcoRI genomic DNA flanking the l(2)k01102 P-element insertion was isolated by plasmid rescue and used to screen an eye disc cDNA library (provided by Dr A. Cowman). Two classes of cDNAs were isolated and sequenced. One is 5.1 kb and encodes an open reading frame (ORF) of 1332 amino acids. We refer to it as ed cDNA. The other cDNA is 4.5 kb and encodes a noncoding RNA: the longest open reading frame (ORF) would encode a polypeptide of 102 amino acids but its AUG codon is in a poor position for translation initiation (Cavener, 1987). The ORF region of the ed cDNA was generated by PCR and inserted into the pCaSpeR-hs and pUAST transformation vectors (Brand and Perrimon, 1993) to create hs-ed and UAS-ed. UAS-edΔintra was made by inserting by PCR a stop codon 31 amino acids after the transmembrane domain of ED. All the constructs were sequenced. Transgenic lines were generated by P-element-mediated transformation (Spradling and Rubin 1982). hs-ed was tested for its ability to rescue the lethality of edslH8 by heat shock in a 37° water bath for 20 minutes every 12 hours throughout development.

ed-specific RT-PCR was performed as follows. Five late third instar larval eye discs of different genotypes were dissected and subjected to cDNA synthesis with a Cells-to-cDNA system (Ambion). 20 cycles of PCR amplifications were carried out with primer pairs from the seventh (CGATGCCCGGAAATGAATGG) and ninth exon (GCGTATGACGCGACGGTTT) of ed genomic DNA. 18S rRNA primers (Ambion) were used as internal controls.

Histology

Fixation, embedding and sectioning of adult retina were performed as described by Wolff and Ready (Wolff and Ready, 1991). Scanning electron micrographs were prepared as described by Kimmel et al. (Kimmel et al., 1990). Cobalt sulphide staining of pupal retinas was performed as described by Wolff and Ready (Wolff and Ready, 1991). Immunohistochemical staining of imaginal discs was performed as described in Xu and Rubin (Xu and Rubin, 1993). Polyclonal rabbit α-ED antibodies were generated against a synthetic peptide, corresponding to the N-terminal region of ED (MRRKTVTKGTAIVNSRSARRAATTI) and were used at a dilution of 1: 200. α-ELAV (rat, 1: 250, Developmental Studies Hybridoma Bank); αi-Cut (mouse, 1: 5, Developmental Studies Hybridoma Bank); α-β-galactosidase (rabbit, 1: 1000, Cappel); α-TTK88 (mouse, 1:100); and α-Boss (mouse, 1:1000); Cy3-, Cy5-, FITC-conjugated secondary IgGs are from Jackson Immunological Laboratories. Confocal microscopy was performed using a Zeiss Model 310.

GenBank Accession Number

The accession number for the ed sequence reported in this paper is AF275903.

Identification of a negative regulator of EGFR signaling pathway

ElpB1 is a gain-of-function allele of the Egfr (Baker and Rubin, 1989). We carried out a genetic modifier screen for components of the EGFR pathway that dominantly enhance or suppress the rough eye phenotype caused by ElpB1 (Fig. 1B). 1×5 was isolated as an EMS induced mutation which strongly enhances the rough eye phenotype associated with ElpB1 (Fig. 1C). The dominant enhancer activity of 1×5 is similar to the effect of Gap1 (Fig. 1D) or yan (data not shown) mutations, two known negative regulators of the EGFR signaling pathway.

Fig. 1.

1×5 genetically interacts with mutations in the EGFR pathway during eye development. A wild-type eye possesses around 750 ommatidia arranged in a highly ordered pattern (A). ElpB1/+ eyes are rough (B) and this phenotype is enhanced when heterozygous for 1×5 (C), and Gap1B2 (D). The eye phenotype of sev-tor4021Egfr/+ (E) is enhanced when heterozygous for 1×5 (F). The rough eye phenotype associated with hs-rho/+ (G) is enhanced when heterozygous for 1×5 (H). Overexpression of aos under the heat shock promoter (hs-aos/+) causes a weak rough eye (I), and this phenotype is suppressed when heterozygous for 1×5 (J).

Fig. 1.

1×5 genetically interacts with mutations in the EGFR pathway during eye development. A wild-type eye possesses around 750 ommatidia arranged in a highly ordered pattern (A). ElpB1/+ eyes are rough (B) and this phenotype is enhanced when heterozygous for 1×5 (C), and Gap1B2 (D). The eye phenotype of sev-tor4021Egfr/+ (E) is enhanced when heterozygous for 1×5 (F). The rough eye phenotype associated with hs-rho/+ (G) is enhanced when heterozygous for 1×5 (H). Overexpression of aos under the heat shock promoter (hs-aos/+) causes a weak rough eye (I), and this phenotype is suppressed when heterozygous for 1×5 (J).

Consistent with the genetic interaction with ElpB1, 1×5 also enhances the eye phenotype caused by sev-tor4021Egfr (Fig. 1E,F), another constitutively active form of the EGFR (Reichman-Fried et al., 1994). To define further the role of 1×5 in the EGFR signaling pathway, we examined the genetic interactions between 1×5 and rho, a specific activator of EGFR pathway, and aos, a specific EGFR inhibitor. Interestingly, we found that 1×5 enhances the rough eye phenotype caused by ectopic expression of rho (Fig. 1G,H), and suppresses the rough eye phenotype caused by misexpression of aos (Fig. 1I,J).

Further genetic interactions between the EGFR pathway and 1×5 were also detected in the wing. 1×5 enhances the extra wing-vein phenotype caused by the overactive ElpB1 mutation, as well as rlSEM, a constitutively active MAPK (Brunner et al., 1994; data not shown). In addition, flies heterozygous for both 1×5 and Gap1, or both 1×5 and styS88 (data not shown), exhibit extra vein materials, although heterozygosity for either mutation alone causes no phenotype. Therefore, the genetic interactions observed between 1×5 and several components of the EGFR pathway suggest that 1×5 is a negative regulator of the EGFR signaling pathway during eye and wing vein development.

1×5 is allelic to echinoid and is required for the formation of photoreceptor and cone cells

1×5 was mapped to 24D3-4 using three overlapping deficiencies: Df(2L)ed1 (24A3-4;24D3-4), Df(2L)ed-dp (24C3-5;25A2-3) and Df(2L)M24F11 (24D3-4;25A2-3). This region contains the ed gene and we found that edlF20 (de Belle et al., 1993) fails to complement 1×5 and enhances the ElpB1 rough eye phenotype, as well as the extra wing vein phenotype of rlSEM (data not shown). Thus 1×5 is allelic to ed, and we refer to it as such below. All ed mutations are pupal lethal in homozygotes with the exception of edslH8, which is a weaker allele. Homozygous edslH8, as well as edslH8 in combination with all other ed alleles, including Df(2L)ed-dp, are semi-lethal. Emerging adults have rough eyes (Fig. 2B) and extra wing veins (see Fig. 5J). When sectioned, 33% of ommatidia contain extra R7-like cells with small and centrally positioned rhabdomeres (Fig. 2D). To exclude that these extra cells with small rhabdomeres are R8, third instar larval imaginal discs of ed1×5/edslH8 transheterozygotes were stained with anti-Boss, an R8-specific antibody. Single R8 cell was seen in each mature ommatidium (Fig. 2E), confirming that the extra photoreceptor cells are indeed R7. In addition, 26% of ommatidia exhibit extra outer-photoreceptor cells while 6% of the ommatidia show reduced outer-photoreceptor cells. Further, edslH8 hemizygotes animals have more R7 cells than ed1×5/edslH8 transheterozygote animals indicating that the ed alleles are loss of function. edslH8 hemizygotes have 1.68 R7 cells in average (n=64), compared with 1.34 (n=164) in ed1×5/edslH8. To determine the origins of the extra photoreceptor cells, ed1×5/edslH8 transheterozygote discs were stained with the anti-Elav neural marker (Fig. 2F). Extra Elav-positive cells were first detected in rows 2 and 3, where R8/R2/R5 are located. However, these four-cell clusters contain only single R8. In addition, one or two extra Elav-positive mystery cells were detected adjacent to R3 and R4 cells four row of cells behind the furrow. Mystery cells will normally leave the five-cell precluster and disappear; however, as in sty or yan mutants (Casci et al., 1999; Lai et al., 1992), they are transformed into neuronal photoreceptor cells in the ed mutant discs.

Fig. 2.

ed mutant eyes contain extra cone and photoreceptor cells. Scanning electron micrographs (A,B) of adult eyes; apical sections through adult retinas (C,D); third instar eye imaginal discs stained for the R8 specific marker, Boss (green), and neuronal marker, Elav (red) (E,F); and pupal retina stained with cobalt sulphide (G,H). (A,C,G) Wild type; (B,D-F,H) are from ed1×5/edslH8 flies. ed1×5/edslH8 transheterozygote animals have large, rough eyes (B). Thin section analysis of these eyes reveals that 26% of the ommatidia contain extra R1-R6 photoreceptor cells, 33% contain extra R7 cells (D) and 6% have a decreased number of photoreceptor cells. The arrowhead indicates an ommatidium that contains two R7 and either six or seven (right) outer photoreceptors. There is only a single R8 cell in each mature ommatidium (E). In some ommatidia, one extra Elav-positive cell (small arrow in F) was first detected in row 2-3 where R8/R2/R5 are located. Inset in (F) shows the four-cell cluster only contains a single R8 cell. In addition, one Elav-positive mystery cell (circle) was found in row 4 (large arrow in F). Together, these may contribute to the formation of the supernumerary photoreceptors that are seen in ed1×5/edslH8 animals. The morphogenetic furrow is to the left. Cobalt-sulfide staining of the ed1×5/edslH8 pupal eye imaginal discs reveals that 69% of the ommatidia have five to six (H), instead of four cone cells (G).

Fig. 2.

ed mutant eyes contain extra cone and photoreceptor cells. Scanning electron micrographs (A,B) of adult eyes; apical sections through adult retinas (C,D); third instar eye imaginal discs stained for the R8 specific marker, Boss (green), and neuronal marker, Elav (red) (E,F); and pupal retina stained with cobalt sulphide (G,H). (A,C,G) Wild type; (B,D-F,H) are from ed1×5/edslH8 flies. ed1×5/edslH8 transheterozygote animals have large, rough eyes (B). Thin section analysis of these eyes reveals that 26% of the ommatidia contain extra R1-R6 photoreceptor cells, 33% contain extra R7 cells (D) and 6% have a decreased number of photoreceptor cells. The arrowhead indicates an ommatidium that contains two R7 and either six or seven (right) outer photoreceptors. There is only a single R8 cell in each mature ommatidium (E). In some ommatidia, one extra Elav-positive cell (small arrow in F) was first detected in row 2-3 where R8/R2/R5 are located. Inset in (F) shows the four-cell cluster only contains a single R8 cell. In addition, one Elav-positive mystery cell (circle) was found in row 4 (large arrow in F). Together, these may contribute to the formation of the supernumerary photoreceptors that are seen in ed1×5/edslH8 animals. The morphogenetic furrow is to the left. Cobalt-sulfide staining of the ed1×5/edslH8 pupal eye imaginal discs reveals that 69% of the ommatidia have five to six (H), instead of four cone cells (G).

Fig. 3.

Molecular characterization of the ed locus. (A) Genomic organization of the ed gene. The ed locus is encompassed by two P1 phage, DS04766 and DS00705. The ed cDNA and the direction of transcription is indicated. The black boxes represent the coding region of ed, whereas the white boxes represent the untranslated regions. ed is composed of nine exons and eight introns. The size of introns from 1 to 8 are 866 bp, 35 kb, 4.79 kb, 62 bp, 17.7 kb, 583 bp, 6.3 kb and 5.8 kb, respectively. The position of the P element l(2)k01102 was mapped to the first intron. The other cDNA shares the first two exons but encodes a noncoding RNA, as the longest open reading frame (ORF) would encode a polypeptide of 102 amino acids but its AUG codon is in an poor context for translation initiation (Cavener, 1987). (B) Amino acid sequence of ED. ed encodes a putative transmembrane protein of 1332 amino acids. The extracellular domain contains six immunoglobulin (Ig) C2 type domains (unbroken lines) and one fibronectin type III domain (double lines). The transmembrane (TM) domain is boxed. (C) Alignment of Ig domains. The consensus sequence of Ig C2 type domain (Williams and Barclay, 1988) is shown. (D) Alignment of the fibronectin type III domain. The consensus sequence of fibronectin type III domain (Hynes, 1986) is shown.

Fig. 3.

Molecular characterization of the ed locus. (A) Genomic organization of the ed gene. The ed locus is encompassed by two P1 phage, DS04766 and DS00705. The ed cDNA and the direction of transcription is indicated. The black boxes represent the coding region of ed, whereas the white boxes represent the untranslated regions. ed is composed of nine exons and eight introns. The size of introns from 1 to 8 are 866 bp, 35 kb, 4.79 kb, 62 bp, 17.7 kb, 583 bp, 6.3 kb and 5.8 kb, respectively. The position of the P element l(2)k01102 was mapped to the first intron. The other cDNA shares the first two exons but encodes a noncoding RNA, as the longest open reading frame (ORF) would encode a polypeptide of 102 amino acids but its AUG codon is in an poor context for translation initiation (Cavener, 1987). (B) Amino acid sequence of ED. ed encodes a putative transmembrane protein of 1332 amino acids. The extracellular domain contains six immunoglobulin (Ig) C2 type domains (unbroken lines) and one fibronectin type III domain (double lines). The transmembrane (TM) domain is boxed. (C) Alignment of Ig domains. The consensus sequence of Ig C2 type domain (Williams and Barclay, 1988) is shown. (D) Alignment of the fibronectin type III domain. The consensus sequence of fibronectin type III domain (Hynes, 1986) is shown.

Fig. 4.

ed is not transcriptionally regulated by EGFR signaling. Embryos of cellular blastoderm stage were labeled with antibodies against ED. The ED immunostaining is uniformly distributed at the membrane of each cell (A). Expression of the enhancer trap reporter gene in edl(2)k01102/+ is detected in the photoreceptor cells of wild type (B) and GMR-aos (C) discs. The lacZ expression of l(2)k1102 mimics the ED expression pattern. The relative levels of ed mRNA from eye discs were measured by RT-PCR and the predicted 554 bp products were visualized on a 1.2% agarose gel (D).

Fig. 4.

ed is not transcriptionally regulated by EGFR signaling. Embryos of cellular blastoderm stage were labeled with antibodies against ED. The ED immunostaining is uniformly distributed at the membrane of each cell (A). Expression of the enhancer trap reporter gene in edl(2)k01102/+ is detected in the photoreceptor cells of wild type (B) and GMR-aos (C) discs. The lacZ expression of l(2)k1102 mimics the ED expression pattern. The relative levels of ed mRNA from eye discs were measured by RT-PCR and the predicted 554 bp products were visualized on a 1.2% agarose gel (D).

Fig. 5.

Overexpression of ed antagonizes the activity of the EGFR signaling pathway. GMR-GAL4/+; UAS-ed/+. eyes are rough and reduced in size (A), and the number of photoreceptors is reduced (B). Note that some ommatidia still contain seven photoreceptors while others only contain four or five photoreceptors (arrowhead). In GMR-GAL4/+; UAS-ed/UAS-ed third instar eye discs (C), there are only four or five Elav-positive cells per cluster. The morphogenetic furrow in this section is to the left. The eyes of GMR-GAL4; UAS-sty flies are rough (D). However, this phenotype is suppressed when heterozygous for edlF20 (E), but enhanced in the presence of a UAS-ed transgene (the genotype of the eye shown in F is GMR-GAL4; UAS-sty; UAS-ed). Similarly, the eye of GMR-GAL4; UAS-kek1 are rough (G), and this phenotype is enhanced in GMR-GAL4; UAS-kek1; UAS-ed flies (H). ed also antagonizes the activity of the EGFR signaling pathway during wing vein formation. In wild type, the veins are arranged in a stereotyped pattern (I), while ed1×5/edslH8 transheterozygote animals show an increased size of the wing and extra wing vein (arrowhead in J). Overexpression of UAS-ed by MS1096-GAL4 results in severe reduction in the vein material and size of wing, ranging from one quarter to one fifth of normal wing size (K).

Fig. 5.

Overexpression of ed antagonizes the activity of the EGFR signaling pathway. GMR-GAL4/+; UAS-ed/+. eyes are rough and reduced in size (A), and the number of photoreceptors is reduced (B). Note that some ommatidia still contain seven photoreceptors while others only contain four or five photoreceptors (arrowhead). In GMR-GAL4/+; UAS-ed/UAS-ed third instar eye discs (C), there are only four or five Elav-positive cells per cluster. The morphogenetic furrow in this section is to the left. The eyes of GMR-GAL4; UAS-sty flies are rough (D). However, this phenotype is suppressed when heterozygous for edlF20 (E), but enhanced in the presence of a UAS-ed transgene (the genotype of the eye shown in F is GMR-GAL4; UAS-sty; UAS-ed). Similarly, the eye of GMR-GAL4; UAS-kek1 are rough (G), and this phenotype is enhanced in GMR-GAL4; UAS-kek1; UAS-ed flies (H). ed also antagonizes the activity of the EGFR signaling pathway during wing vein formation. In wild type, the veins are arranged in a stereotyped pattern (I), while ed1×5/edslH8 transheterozygote animals show an increased size of the wing and extra wing vein (arrowhead in J). Overexpression of UAS-ed by MS1096-GAL4 results in severe reduction in the vein material and size of wing, ranging from one quarter to one fifth of normal wing size (K).

We also examined the ed mutant phenotype during pupariation. At this stage there are four cone cells and two primary pigment cells in wild-type discs (Fig. 2G). However, 69% of ommatidia in ed1×5/edslH8 transheterozygotes exhibit five or six cone cells (Fig. 2H) and 10% contain three primary pigment cells (data not shown). Together, the over-recruitment of photoreceptor, cone and pigment cells in ed mutants is consistent with ED acting as a negative regulator of EGFR because previous analyses have shown that EGFR is required for differentiation of these three cell types (Freeman, 1996).

ed encodes an adhesion molecule-like protein with six immunoglobulin domains

We identified a P-element insertion, l(2)k01102 (Torok et al., 1993) located at 24D3-4, that failed to complement either ed1×5 or edlF20. To characterize ed molecularly, we recovered the DNA region flanking l(2)k01102 by plasmid rescue (see Materials and Methods). A 8 kb genomic DNA was isolated and used to screen an eye disc cDNA library. A 5.1 kb cDNA was isolated and sequence analysis revealed that it encodes an open reading frame of 3996 bp, which predicts a protein of 1332 amino acids (Fig. 3B). The translated protein contains six immunoglobulin (Ig) C2 type domains (Williams and Barclay, 1988) (Fig. 3C), a fibronectin type III domain (Hynes, 1986) (Fig. 3D) and a transmembrane domain, followed by a 315 amino acid C-terminal tail with no identifiable functional motif. A comparison of the genomic and cDNA sequence indicates that l(2)k1102 is inserted in the first intron, which is upstream of the coding region (Fig. 3A). To establish that the 5.1 kb cDNA identifies ed, we expressed the cDNA under the control of a heat shock promoter. Following heat shock treatments, we found that the hs-ed transgene rescues the lethality associated with the weak edslH8 allele (data not shown).

To detect the expression pattern of ED, we stained embryos with an antibody generated against the N-terminal ED peptide. The ED protein is widely expressed in the epidermis and is localized to the plasma membrane (Fig. 4A). Further, we find that ED is uniformly detected in all cells throughout the third instar larval eye and wing disc (data not shown).

The expression of aos and kek1, two other negative regulators of the EGFR pathway, is regulated by the EGFR pathway. To determine whether ed is regulated by the EGFR pathway, we examined the expression of ed in GMR-aos (Fig. 4B-D) and sev-RasV12 (data not shown) eye discs. In each case, the level of ed mRNA is not affected, as revealed by either the X-gal staining of the P insertion l(2)k1102. (Fig. 4B,C) or the ed-specific relative RT-PCR (Fig. 4D). These results indicate that ed, unlike aos and kek1, is not transcriptionally regulated by the activation of EGFR pathway.

Overexpression of ed antagonizes EGFR activity in the eye and wing

As shown above, loss of ed function is required for the formation of photoreceptor, cone and primary pigment cells. To determine the effect of overexpression of ed in the eye, we expressed UAS-ed using the GMR-Gal4 driver. GMR-Gal4; UAS-ed flies exhibit a small rough eye (Fig. 5A) and a reduced number of photoreceptors (Fig. 5B); this effect correlates with the reduced number of Elav-positive cells (Fig. 5C) in the eye disc. There are only four or five Elav-positive cells per cluster. In contrast, no obvious defects in the formation of cone cells were observed in response to ed overexpression, as most ommatidia still contain four Cut-positive cells (data not shown). Flies carrying two copies of GMR-GAL4-driven UAS-ed exhibit complete absence of the eye (data not shown).

To further document the interaction between ED and the EGFR pathway, we examined the effect of ectopic expression of ed in flies where other regulators were overexpressed. Overexpression of UAS-sty alone by GMR-GAL4 produces small rough eye (Fig. 5D). This phenotype can be partially suppressed by halving the dose of ed (Fig. 5E), and enhanced by GMR-GAL4-driven UAS-ed (Fig. 5F). Similar genetic interactions can also be observed between ed and kek1. The rough eye phenotype caused by GMR-GAL4-driven UAS-kek1 (Fig. 5G) is enhanced by GMR-GAL4-driven UAS-ed (Fig. 5H). Therefore ed, like sty and kek1, is a repressor of EGFR signaling during eye development.

Similarly, during wing vein development, ed genetically interacts with several components in the EGFR pathway. Flies of ed1×5/edslH8 have increased size of wing and extra wing vein (Fig. 5J). However, ectopic expression of ed using MS1096 GAL4 results in severe reduction in the size of wing, ranging from one quarter to one fifth of normal wing size. In addition, there is no vein material present (Fig. 5K).

The intracellular domain of ED is required for the repression

ED contains six Ig domains and a 315 amino acid intracellular domain. To determine whether the intracellular domain of ED is required for the repression of the EGFR signaling, we generated UAS-edΔintra flies. Overexpression ofUAS-edΔintra usingGMR-GAL4 had no phenotypes in the eye, indicating that the cytoplasmic domain of ED is required for the repression of the EGFR signaling pathway.

ed acts parallel or downstream of sina but upstream of ttk to specify R7 cells

To determine where in the RAS/RAF/MAPK signaling pathway ed acts, we conducted a number of genetic epistasis experiments. sevd2 is a loss-of-function sevenless (sev) allele (Simon et al., 1991) and sevd2 mutant flies lack R7 cells (Fig. 6A). Although ommatidia within a ed1×5/edslH8 mutants contain an average of 1.34 R7 cell (Fig. 2F), we found that ommatidia within a sevd2; ed1×5/edslH8 double mutant contain an average of 1.37 R7 cells (n=61) (Fig. 6B). This demonstrates that in ed mutants, the formation of supernumary R7 cells is independent of sev function. In addition, ed1×5 enhances the rough eye phenotype caused by overexpressing constitutive active forms of either the EGFR (Fig. 1C,F), RAS1 (Fig. 6C,D), or RAF (Fig. 6G,H). Conversely, ed1×5 suppresses the rough eye phenotype caused by overexpressing dominant negative RAS1 (Fig. 6E,F). While 61% of ommatidia in a sev-RasN17/+ mutant lack R7 cells (n=164), only 10% of ommatidia in ed1×5/edslH8; sev-RasN17/+ double mutants lack R7 photoreceptors. In addition, at 25°, ed1×5 also rescues the lethality of RafHM7, a temperature-sensitive Raf allele. Therefore, ed acts either downstream of the Ras/Raf pathway or in parallel.

Fig. 6.

Genetic epistatic analysis of ed. sevd2 ommatidia have no R7 cells (A) but sevd2; ed1×5/edslH8 ommatidia contain ectopic R7 cells (B, arrowheads). The rough eye phenotype caused by sev-Ras1V12/+ (C) is enhanced when heterozygous for ed1×5 (D). Overexpression of RasN17 under the control of the sevenless enhancer (sev-RasN17/+) causes a rough eye (E), and this phenotype is suppressed when heterozygous for ed1×5 (F). The rough eye phenotype associated with sev-tor4021Raf /+ (G) is enhanced when heterozygous for ed1×5 (H). Ommatidia within a pntΔ88/pnt1277 (I) and sev-yanACT/+ (K) mutants contain 0.69 and 0.05 R7 cell, respectively. However, ommatidia within ed1×5/edslH8; pntΔ88/pnt1277 (J) and ed1×5/edslH8; sev-yanACT/+ (L) double mutants contain 1.44 and 1.01 R7 cells (arrowheads), respectively. sina2/sina3 (M) ommatidia contain 0.01 R7 cells, however, ed1×5/edslH8; sina2/sina3 (N) double mutants contain 1.29 R7 cells (arrowhead). Overexpression of ttk88 under the control of the GMR enhancer (GMR-ttk88/+) blocks photoreceptor determination (O), and this phenotype can not be suppressed in ed1×5/edslH8; GMR-ttk88/+ (P).

Fig. 6.

Genetic epistatic analysis of ed. sevd2 ommatidia have no R7 cells (A) but sevd2; ed1×5/edslH8 ommatidia contain ectopic R7 cells (B, arrowheads). The rough eye phenotype caused by sev-Ras1V12/+ (C) is enhanced when heterozygous for ed1×5 (D). Overexpression of RasN17 under the control of the sevenless enhancer (sev-RasN17/+) causes a rough eye (E), and this phenotype is suppressed when heterozygous for ed1×5 (F). The rough eye phenotype associated with sev-tor4021Raf /+ (G) is enhanced when heterozygous for ed1×5 (H). Ommatidia within a pntΔ88/pnt1277 (I) and sev-yanACT/+ (K) mutants contain 0.69 and 0.05 R7 cell, respectively. However, ommatidia within ed1×5/edslH8; pntΔ88/pnt1277 (J) and ed1×5/edslH8; sev-yanACT/+ (L) double mutants contain 1.44 and 1.01 R7 cells (arrowheads), respectively. sina2/sina3 (M) ommatidia contain 0.01 R7 cells, however, ed1×5/edslH8; sina2/sina3 (N) double mutants contain 1.29 R7 cells (arrowhead). Overexpression of ttk88 under the control of the GMR enhancer (GMR-ttk88/+) blocks photoreceptor determination (O), and this phenotype can not be suppressed in ed1×5/edslH8; GMR-ttk88/+ (P).

To determine whether ED acts in the nucleus, we generated flies double mutant for ed;pnt, ed;yan or ed;sina. We found that pntΔ88/pnt1277 (Fig. 6I) and sev-yanACT/+ (Fig. 6K) ommatidia contain an average of 0.69 (n=264) and 0.05 (n=250) R7 cells, respectively. However, ed1×5/edslH8; pntΔ88/pnt1277 (Fig. 6J) and ed1×5/edslH8; sev-yanACT/+ ommatidia (Fig. 6L) contain an average of 1.44 (n=102) and 1.01 (n=125) R7 cells, respectively. Strikingly, ed1×5/edslH8; sina2/sina3 ommatidia (Fig. 6N) contain an average of 1.29 (n=180) R7 cells, as compared with 0.01 (n=173) R7 cells in sina2/sina3 mutant (Fig. 6M). Therefore, in ed mutants, the formation of supernumary R7 cells is independent of sina function. Finally, loss of ttk activity has been shown to produce ectopic R7 cells in a sina-independent manner (Lai et al., 1996). To determine whether ed acts downstream of ttk, we overexpressed ttk in ed mutants. Overexpression of TTK88 under the control of either the GMR enhancer that completely inhibits photoreceptor cell development (Fig. 6L) or the sev enhancer that only deletes R3, R4 and R7 photoreceptors (data not shown). However, this TTK88-mediated neuronal repression cannot be suppressed by removing ed activity (Fig. 6P), indicating that ed acts upstream of ttk to specify R7 development. Together, our genetic epistatic analysis suggests that ed acts either parallel or downstream of Ras, Raf, pnt, yan and sina, but upstream of ttk to specify R7 cell fates.

ED does not regulate ttk88 expression or protein stability

Our genetic epistatic analyses suggest that ed acts upstream of ttk88 to specify R7. ed might regulate ttk88 mRNA expression or TTK88 protein levels. Alternatively, ed might regulate the activity of TTK88 through protein modification, like phosphorylation. To determine whether ed regulates ttk expression, we examined the expression of ttk in ed mutant disc using the X-gal staining of the P-element insertion ttk0219 (Li et al., 1997) and detected no obvious changes (Fig. 7B). Furthermore, TTK88 is expressed at high levels in the cone cells but is not expressed in developing photoreceptor cells (Li et al., 1997; Dong et al., 1999). To determine whether ed regulates TTK88 protein levels, we examined TTK88 levels in ed (Fig. 7D) and GMR-Gal4; UAS-ed eye discs (data not shown). In each case, the level of TTK88 was unaffected. Together, our results suggest that ED does not regulate ttk88 mRNA expression or TTK88 protein stability.

Fig. 7.

ed does not regulate ttk88 expression or protein stability. ttk0219 enhancer trap reporter gene is expressed in the cone cells (A) using anti-β-galactosidase. The expression levels of ttk0219 are unaffected in ed1×5/edslH8 mutant disc (B), which contain extra cone cells. In wild type, TTK88 is shown in the cone cell nuclei (C) using anti-TTK88. The levels of TTK88 do not change in ed1×5/edslH8 mutant disc (D). A six-cell cluster is marked.

Fig. 7.

ed does not regulate ttk88 expression or protein stability. ttk0219 enhancer trap reporter gene is expressed in the cone cells (A) using anti-β-galactosidase. The expression levels of ttk0219 are unaffected in ed1×5/edslH8 mutant disc (B), which contain extra cone cells. In wild type, TTK88 is shown in the cone cell nuclei (C) using anti-TTK88. The levels of TTK88 do not change in ed1×5/edslH8 mutant disc (D). A six-cell cluster is marked.

The ed mutation acts nonautonomously to generate extra R7 cells

To determine in which cells ed is required, we used ey-FLP (Newsome et al., 2000) to generate clones of homozygous edslA12 mutant cells in a sevd2 background. As shown in Fig. 6A, no R7 cells develop in the sevd2 background. We scored 54 mosaic ommatidia that contain R7-like cells. Among them, 57% of the R7-like cells were ed (Fig. 8A), while 43% were ed+ (Fig. 8B). Similar results were obtained when we generated ed mutant clones in sina and sev-yanACT mutant backgrounds (data not shown). The observation that R7 cells can be derived from either wild-type or ed mutant cells, leads us to propose that the ed mutation acts cell non-autonomously in the generation of supernumerary R7 cells.

Fig. 8.

ed functions in a non cell-autonomous manner. Phase-contrast images of section through an edslA12 homozygous mutant clone induced in a sevd2/Y; edslA12/+ animals. The clones are marked by the lack of pigmentation. sevd2 ommatidia have no R7 cells. A total of 54 phenotypically normal mosaic ommatidia were scored for the presence of pigment in the R7 cells. The ectopic R7 cells (arrowheads) can be derived from either ed mutant (A) or wild-type cells (B), indicating that the ed mutation acts cell non-autonomously in the generation of supernumerary R7 cells.

Fig. 8.

ed functions in a non cell-autonomous manner. Phase-contrast images of section through an edslA12 homozygous mutant clone induced in a sevd2/Y; edslA12/+ animals. The clones are marked by the lack of pigmentation. sevd2 ommatidia have no R7 cells. A total of 54 phenotypically normal mosaic ommatidia were scored for the presence of pigment in the R7 cells. The ectopic R7 cells (arrowheads) can be derived from either ed mutant (A) or wild-type cells (B), indicating that the ed mutation acts cell non-autonomously in the generation of supernumerary R7 cells.

ED is not a universal repressor

ED is uniformly expressed in the follicle cells during stage 1-10 oogenesis (data not shown). To determine whether ed acts during oogenesis in the establishment of EGFR-dependent dorsal/ventral polarity, we examined the eggs derived from edslH8/Df(2L)ed-dp females. These females are fertile and do not exhibit any overt morphological defects (data not shown). As loss-of-function mutations in many cell adhesion molecule have subtle mutant phenotypes (Ghiglione et al., 1999), we overexpressed UAS-ed in the follicle cells using the GAL4 drivers T155 or CY2 (Queenan et al., 1997; Ghiglione et al., 1999). The eggs derived from such females have completely normal dorsal appendages (data not shown) suggesting that ED does not interfere with EGFR signaling in follicle cells.

The EGFR plays important roles at various stages of Drosophila development and is subject to modulation by multiple positive and negative regulators. We have identified ED as a novel adhesion molecule-like protein that negatively regulates the EGFR signaling pathway. ed genetically interacts with several components in the EGFR pathway. Flies of ed mutant produce extra photoreceptor and cone cells. Conversely, ectopic overexpression of ed in the eye leads to reduction of photoreceptor number. We demonstrate that ED acts by converging on TTK88, the most downstream component known in EGF receptor signaling. Our results not only demonstrate the active role of an adhesion molecule in the EGFR signal transduction pathway but also identify a previously unknown regulatory mechanism.

ed acts nonautonomously

ED contains six Ig domain with extensive homology to vertebrate neural adhesion molecule L1. The L1-family of cell adhesion molecules exert their functions through homophilic or heterophilic interactions with other Ig domaincontaining adhesion molecule (Hortsch, 1996). We found that ED is expressed in every cell of the eye disc. In addition, our genetic analysis demonstrate that ed acts in a cell nonautonomous manner to generate extra R7 cells. If ED transmits the negative signal from the sending cell via homophilic interaction to the receiving cell, loss of ed in either sending or receiving cells would result in the same phenotype, owing to the failure to receive the inhibitory signal. Therefore the extra R7 cells found in the receiving cells could be either wild type or mutant for ed. However, if ED transmits the negative signaling via heterophilic interaction, ed is only required in the sending cells but not the receiving cells. Therefore, the extra R7 cells found in the receiving cells could be either wild type or mutant for ed. Alternatively, ED might act as a ligand that activates an unidentified receptor on receiving cells. All three models are consistent with our results showing that ed functions cell nonautonomously. However, only the homophilic interaction model would require the cytoplasmic domain of ED to be required in both the sending and receiving cells. Since we found that the cytoplasmic domain of ED is required for the repression of the EGFR pathway, we favor the homophilic interaction model between ED molecules to specify photoreceptor cell formation (Fig. 9).

Fig. 9.

Model for ED function. The membrane-spanning ED protein transmits the negative signal, via homotypic interactions, into the receiving cell where it antagonizes the EGFR signaling by converging on TTK88.

Fig. 9.

Model for ED function. The membrane-spanning ED protein transmits the negative signal, via homotypic interactions, into the receiving cell where it antagonizes the EGFR signaling by converging on TTK88.

ed define an independent pathway to repress EGFR signaling

Studies on RTK signaling in both vertebrates and invertebrates have converged on an evolutionarily conserved DRK/RAS/ RAF/MAPK signaling cassette that is required to transmit the signal from the receptor to the nucleus. Previous studies on TORSO (TOR) signaling, however, indicated that TOR RTK transduces its signals through both a RAS-dependent and an unidentified RAS-independent pathways that converge on RAF (Hou et al., 1995). In addition, genetic analysis of daughter of sevenless (Dos) has revealed that it functions upstream of RAS1 and defines a signaling pathway that is independent of direct binding of DRK/GRB2 to the SEV RTK (Raabe et al., 1996; Herbst et al., 1996). Nevertheless, either a RAS-or KSR-independent pathway still acts underneath a RTK.

In contrast to the previous examples, our genetic data indicates that ED functions either downstream of RAS1/MAPK/PNT/YAN or in a parallel pathway. However, based on the following data we argue that ED is unlikely to act downstream of PNT/YAN/SINA in the nucleus, but instead defines an independent pathway that antagonizes EGFR signaling (Fig. 9). First, antiserum against ED N-terminal peptide localizes ED to the plasma membrane, but not the nucleus, of every cell in the eye disc. Second, ED functions non-autonomously in the signal-sending cells. Third, ed is not transcriptionally regulated by the activation of the EGFR pathway, a situation that is different from other negative regulators of EGFR such as aos, sty and kek1. Fourth, SINA has been shown to form a complex with PHYL to target TTK for degradation. The production of ectopic R7 cells in mutations of most negative regulators, like Gap1 and yan, all require sina (Gaul et al., 1992; Lai and Rubin, 1992). However, the formation of extra R7 cells in ttk mutant is only partially sina dependent. This observation led Lai et al., to suggest that the production of extra R7 cells in ttk mutant is partly influenced by both the normal R7 developmental signals, which are sina dependent, and another sin-independent signaling (Lai et al., 1996). Our genetic analysis demonstrates that the production of ectopic R7 cells in ed mutant is completely sina independent. The observation that ed functions upstream of ttk88, implies that the independent inhibitory pathway, although sina independent, converges on TTK.

Mechanism of inhibition of the EGFR by ED

ED, a putative cell adhesion protein, is constitutively expressed on every cells on the eye disc. There are several ways in which ed expression can influence signaling. For example, it could induce polarization and adherens junction formation of undifferentiated cells. The EGFR is localized to the apical microvillar border where it binds its inductive ligand (Zak and Shilo, 1992). The apical restriction of EGFR may concentrate these receptors at a high density and allows efficient capture of the SPI ligand, thus restricting SPI diffusion over a long distance. In the absence of ED, the EGFR might diffuse to the basolateral membrane and the density of EGFR may be too low to capture SPI efficiently. According to this model, SPI may diffuse to distant cells. These cells which normally do not encounter the ligand would then differentiate extra photoreceptor or cone cells. In this case, ED would function as a mechanical force to affect the binding efficiency of EGFR and the diffusion distance of SPI. However, we do not favor this mechanism because we observe that halving the dose of ed can enhance the rough eye phenotype caused by constitutively active EGFR (sev-tor4021Egfr), which does not require SPI.

The other possibility is that ED, via homophilic interactions, may directly transmit a negative signal. This signal would counteract the basal ligand-independent activity of RTK, caused by nonspecific RTK oligomerization, and establish an inherent inhibitory network to prevent cells from differentiating as photoreceptor or cone cells. According to this model, only when a cell receives its ligand can it activate its RTK and antagonize this negative effect and differentiates. Thus, loss of ed activity behind furrow produces ectopic photoreceptor and cone cells. In this case, a photoreceptor differentiation response would be elicited in region where only the RTK signaling pathway was activated. Thus, RTK activation functions in a permissive manner.

Loss of ed gene activity results in ectopic photoreceptor and cone cells formation. Thus ED, like GAP1 and YAN, functions as a general repressor of differentiation in the developing eye. The negative signal that ED transmits might be mediated through TTK to repress photoreceptor cell formation and another negative regulator to repress cone cells formation. How does ed transmit an inhibitory signal into the receiving cells? The neurite outgrowth and axonal fasciculation mediated by L1-family neural adhesion molecule require both the homotypic or heterotypic interactions of extracellular Ig domains and the conserved ankyrin binding site (FIGQY) in the cytoplasmic domain (Dubreuil et al., 1996; Hortsch et al., 1998). Unlike members of L1-family, ED contains a 315 amino acid cytoplasmic domain with no apparent sequence homology to the conserved FIGQY ankyrin-binding site. However, deletion analysis indicates that the intracellular domain is required for the repression of the EGFR pathway. Finally, the repressing effect of ED signaling in photoreceptors is mediated through TTK. There are several ways that ED signaling might affect TTK. It might elevate the repressing activity of TTK through posttranslational modification, like phosphorylation, to increase its DNA binding affinity to its target genes. Alternatively, it might directly upregulate ttk mRNA or its protein levels. The activation of EGFR signaling has been shown to downregulate TTK88 protein stability (Tang et al., 1997; Li et al., 1997). We favor the former possibility, as we found that the levels of both ttk mRNA and ED protein are unaffected in ed mutant background.

ED is a tissue specific repressor

ED is widely expressed at various stages of Drosophila development (J.-C. H., unpublished observations). Our results demonstrate that ED is a negative regulator of EGFR and Sev signaling pathways during eye and wing development. However, ED does not appear to be involved in EGFR signaling during oogenesis. Therefore, ED differentially functions as an inhibitor of RTK in a tissue-specific manner.

We thank M. B. Sokolowski, I. Kiss, G. M. Rubin, Z.-C. Lai, M. Freeman, M. Simon, C. Klambt, B. Dickson, E. Hafen, H. Y. Sun and the Bloomington Stock Center who sent us fly stocks, and Z.-C. Lai, S. L. Zipursky and R. W. Carthew, who sent us anti-TTK88 antibody. We also thank D. F. Ready for help with cobalt sulphide staining, C.-T. Chien for help with confocal microscopy, C.-K. Chen for help with antibody staining and C.-Y. Tang for help with microinjection. We are grateful to M. P. Zeidler, C.-T. Chien and Henry Y. Sun for helpful discussion and critical reading of the manuscript. This study was supported by grants to J. C. H. from the National Science Council (NSC), Taiwan, Republic of China, and a grant from the US Army Breast Cancer Research (N. P).

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
.
Bang
,
A. G.
and
Kintner
,
C.
(
2000
).
Rhomboid and Star facilitate presentation and processing of the Drosophila TGF-homolog Spitz
.
Genes Dev
.
14
,
177
186
.
Brand
,
A. H.
and
Perrimon
,
N.
(
1993
).
Targeted gene expression as a means of altering cell fates and generating dominant phenotypes
.
Development
118
,
401
415
.
Brunner
,
D.
,
Oellers
,
N.
,
Szabad
,
J.
,
Biggs
,
W. H.
,
Zipursky
,
S. L.
and
Hafen
,
E.
(
1994
).
A gain of function mutation in Drosophila MAP kinase activates multiple receptor kinase signaling pathways
.
Cell
76
,
875
888
.
Brunner
,
D.
,
tDucker
,
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
, 1994.
Capdevila
,
J.
and
Guerrero
,
I.
(
1994
).
Targeted expression of the signaling molecule decapentaplegic induces pattern duplications and growth alterations in Drosophila wings
.
EMBO J
.
13
,
4459
4468
.
Carthew
,
R. W.
and
Rubin
,
G. M.
(
1990
).
seven in absentia, a gene required for specification of R7 cell fate in the Drosophila eye. Cell
63
,
561
577
.
Casci
,
T.
,
Vinos
,
J.
and
Freeman
,
M.
(
1999
).
tSprouty, an intracellular inhibitor of Ras signaling
.
Cell
96
,
655
665
.
Cavener
,
D. R.
(
1987
).
Comparison of the consensus sequence flanking translational start sites in Drosophila and vertebrates
.
Nucleic Acids Res
.
15
,
1353
1361
.
Chou
,
T. B.
,
Noll
,
E.
and
Perrimon
,
N.
(
1993
).
Autosomal P[ovoD1] dominant female-sterile insertions in Drosophila and their use in generating germ-line chimeras
.
Development
119
,
1359
1369
.
de Belle
,
J. S.
,
Sokolowski
,
M. B.
and
Hilliker
,
A. J.
(
1993
).
Genetic analysis of foraging microregion of Drosophila melanogaster
.
Genome
36
,
94
101
.
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
.
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
.
Dubreuil
,
R. R.
,
MacVicar
,
G.
,
Dissanayake
,
S.
,
Liu
,
C.
,
Homer
,
D.
and
Hortsch
,
M.
(
1996
).
Neuroglian-mediated cell adhesion induces assembly of the membrane skeleton at cell contact sites
.
J. Cell Biol
.
133
,
647
655
.
Fortini
,
M. E.
,
Simon
,
M. A.
and
Rubin
,
G. M.
(
1992
).
Signaling by the sevenless protein tyrosine kinase is mimicked by Ras1 activation
.
Nature
355
,
559
561
.
Freeman
,
M.
,
Klambt
,
C.
,
Goodman
,
C. S.
and
Rubin
,
G. M.
(
1992
).
The argos gene encodes a diffusible factor that regulate cell fate decisions in the Drosophila eye
.
Cell
69
,
963
975
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
.
Freeman
,
M.
(
1997
).
Cell determination strategies in the Drosophila eye
.
Development
124
,
261
270
.
Gaul
,
U.
,
Mardon
,
G.
and
Rubin
,
G. M.
(
1992
).
A putative Ras GTPase activating protein acts as a negative regulator of signaling by the sevenless receptor tyrosine kinase
.
Cell
68
,
1007
1019
.
Ghiglione
,
C.
,
Carraway III
,
K. L.
,
Amundadottir
,
L. T.
,
Boswell
,
R. E.
,
Perrimon
,
N.
and
Duffy
,
J. B.
(
1999
).
The transmembrane molecule kekkon1 acts in a feedback loop to negatively regulate the activity of the Drosophila EGF receptor during oogenesis
.
Cell
96
,
847
856
.
Herbst
,
R.
,
Carroll
,
P. M.
,
Allard
,
J. D.
,
Schilling
,
J.
,
Raabe
,
T.
and
Simon
,
M. A.
(
1996
).
Daughter of sevenless is a substrate of the phosphotyrosine phosphatase corkscrew and functions during sevenless signaling
.
Cell
85
,
899
909
.
Hortsch
,
M.
(
1996
).
The L1 family of neuronal cell adhesion molecules: old proteins performing new tricks
.
Neuron
17
,
587
593
.
Hortsch
,
M.
,
Homer
,
D.
,
Malhotra
,
J. D.
,
Chang
,
S.
,
Frankel
,
J.
,
Jefford
,
G.
and
Dubreuil
,
R. R.
(
1998
).
Structural requirements for outside-in and inside-out signaling by Drosophila Neuroglian, a member of the L1 family of cell adhesion molecules
.
J. Cell Biol
.
142
,
251
261
.
Hou
,
X. S.
,
Chou
,
T.-B.
,
Melnick
,
M. B.
and
Perrimon
,
N.
(
1995
).
The torso receptor tyrosine kinase can activate raf in a Ras-independent pathway
.
Cell
81
,
63
71
.
Hynes
,
R. O.
(
1986
).
Molecular biology of fibronectin
.
Annu. Rev. Cell Biol
.
1
,
67
90
.
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
.
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 development
.
Genetics
143
,
315
329
.
Karim
,
F. D.
and
Rubin
,
G. M.
(
1999
).
PTP-ER, a novel tyrosine phosphatase, functions downstream of Ras1 to downregulate MAPK kinase during Drosophila eye development
.
Cell
3
,
741
750
.
Kimmel
,
B. E.
,
Heberlein
,
U.
and
Rubin
,
G. M.
(
1990
).
The homeodomain protein Rough is expressed in a subset of cells in the developing Drosophila eye where it can specify photoreceptor cell subtype
.
Genes Dev
.
4
,
712
727
.
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
.
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. USA
93
,
5025
5030
.
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
.
Miyamoto
,
S.
,
Teramoto
,
H.
,
Gutkind
,
J. S.
and
Yamada
,
K. M.
(
1996
).
Integrins can collaborate with growth factors for phosphorylation of receptor tyrosine kinases and MAP kinase activation: roles of integrin aggregation and occupancy of receptors
.
J. Cell Biol
.
135
,
1633
1642
.
Moghal
,
N.
and
Neel
,
B. G.
(
1998
).
Integration of growth factor, extracellular matrix, and retinoid signals during bronchial epithelial cell differentiation
.
Mol. Cell. Biol
.
18
,
6666
6678
.
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
.
Moro
,
L.
,
Venturino
,
M.
,
Bozzo
,
C.
,
Silengo
,
L.
,
Altruda
,
F.
,
Beguinot
,
L.
,
Tarone
,
G.
and
Defilippi
,
P.
(
1998
).
Integrins induce activation of EGF receptor: role in MAP kinase induction and adhesion-dependent cell survival
.
EMBO J
.
17
,
6622
6632
.
Newsome
,
T. P.
,
Asling
,
B.
and
Dickson
,
B. J.
(
2000
).
Analysis of Drosophila photoreceptor axon guidance in eye-specific mosaics
.
Development
127
,
851
860
.
O’Neill
,
E. M.
,
Rebay
,
I.
,
Tijian
,
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
.
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
.
Queenan
,
A. M.
,
Ghabrial
,
A.
and
Schupbach
,
T.
(
1997
).
Ectopic activation of torpedo/Egfr, a Drosophila receptor tyrosine kinase, dorsalizes both the eggshell and the embryo
.
Development
124
,
3871
3880
.
Raabe
,
T.
,
Riesgo-Escovar
,
J.
,
Liu
,
X.
,
Bausenwein
,
B. S.
,
Deak
,
P.
,
Maroy
,
P.
and
Hafen
,
E.
(
1996
).
tDos, a novel pleckstrin homology domain-containing protein required for signal transduction between sevenless and Ras1 in Drosophila
.
Cell
85
,
911
920
.
Read
,
D.
and
Manley
,
J. L.
(
1992
).
Alternatively spliced transcripts of the Drosophila tramtrack gene encode zinc finger proteins with distinct DNA binding specifications
.
EMBO J
.
11
,
1035
1044
.
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
.
Reichman-Fried
,
M.
,
Dickson
,
B.
,
Hafen
,
E.
and
Shilo
,
B.-Z.
(
1994
).
Elucidation of the role of breathless, a Drosophila FGF receptor homolog, in tracheal cell migration
.
Genes Dev
.
8
,
428
439
.
Schweitzer
,
R.
,
Howes
,
R.
,
Smith
,
R.
,
Shilo
,
B.-Z.
and
Freeman
,
M.
(
1995
).
Inhibition of Drosophila EGF receptor activated 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
.
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.
and
Rubin
,
G. M.
(
1982
).
Transposition of cloned P elements into Drosophila germ line chromosome
.
Science
218
,
341
347
.
Tan
,
P. B. O.
and
Kim
,
S. K.
(
1999
).
Signaling specificity
.
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
.
Torok
,
T.
,
Tick
,
G.
,
Alvarado
,
M.
and
Kiss
,
I.
(
1993
).
P-lacW insertional mutagenesis on the second chromosome of Drosophila melanogaster: isolation of lethals with different overgrowth phenotypes
.
Genetics
135
,
71
80
.
Wasserman
,
J. D.
and
Freeman
,
M.
(
1997
).
Control of EGF receptor activation in Drosophila
.
Trends Cell Biol
.
7
,
431
435
.
Williams
,
A. F.
and
Barclay
,
A. N.
(
1988
).
The immunoglobulin superfamily-domains for cell surface recognition
.
Annu. Rev. Immunol
.
6
,
381
405
.
Wolff
,
T.
and
Ready
,
D. F.
(
1991
).
Cell death in normal and rougheye mutants of Drosophila
.
Development
113
,
825
839
.
Wolff
,
T.
and
Ready
D. F.
(
1993
). Pattern formation in Drosophila retina. In
The Development of Drosophila melanogaster
(ed.
M.
Bate
and
A.
Martinez-Arias
), pp.
1277
1325
.
New York: Cold Spring Harbor Press
.
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
,
T.
and
Rubin
,
G. M.
(
1993
).
Analysis of genetic mosaics in developing and adult Drosophila tissues
.
Development
117
,
1223
1237
.
Yamamoto
,
D.
,
Nihonmatsu
,
I.
,
Matsuo
,
T.
,
Miyamoto
,
H.
,
Kondo
,
S.
,
Hirata
,
K.
and
Ikegami
,
Y.
(
1996
).
Genetic interactions of pokkuri with seven in absentia, tramtrack and downstream components of the sevenless pathway in R7 photoreceptor induction in Drosophila melanogaster
.
Roux’s Arch. Dev. Biol
.
205
,
215
224
.
Yamauchi
,
T.
,
Ueki
,
K.
,
Tobe
,
K.
,
Tamemoto
,
H.
,
Sekine
,
N.
,
Wada
,
M.
,
Honjo
,
M.
,
Takahashi
,
M.
,
Takahashi
,
T.
,
Hirai
,
H.
et al. 
. (
1997
).
Tyrosine phosphorylation of the EGF receptor by the kinase Jak2 is induced by growth hormone
.
Nature
390
,
91
96
.
Zak
,
N. B.
and
Shilo
,
B.-Z.
(
1992
).
Localization of DER and the pattern of cell division in wild-type and Ellipse eye imaginal discs
.
Dev. Biol
.
149
,
448
456
.