We describe the molecular characterization of the Drosophila gene Serrate (Ser), which encodes an integral membrane protein. The extracellular domain contains two cysteine-rich regions, one of which is organized in a tandem array of 14 EGF-like repeats. Antibodies directed against part of the extracellular region confirm the localization of the protein in the membrane. In the wing imaginai discs, the protein is detected in those regions that are affected in the wings of two dominant mutations, SerD and SerBd. Both mutations as well as three out of eight newly induced revertants of SerD could be mapped molecularly to the transcribed region, confirming the identity between the gene Ser and the transcription unit characterized. During embryonic development, RNA and protein exhibit a complex expression pattern, which is, however, not correlated with an appropriate embryonic phenotype. Phenotypic interactions of Ser alleles with the neurogenic genes Notch and Delta coupled with the structural similarity of the proteins encoded by these three genes suggest close interactions at the protein level.

. Cellular interactions play an important role during the development of multicellular organisms. Strikingly, individual elements necessary for cell–cell interactions, as signalling molecules, receptors or adhesion proteins, have been found in some cases to be conserved to a fairly high degree, suggesting that similar functions are mediated by similar proteins or domains of proteins. One group of proteins involved in cell–cell interactions is characterized by the presence of one or several repeated units, each of which shows similarity to the epidermal growth factor, EGF. Members of this family have been found in a variety of organisms, mainly in vertebrates, but also in invertebrates: in nematodes, sea urchins (Hursh et al. 1987; Suyemitsu et al. 1989) and Drosophila. They are either secreted, e.g. EGF itself, TGF-α(Lee et al. 1985; Marquard et al. 1984) or several factors of the blood coagulation system (see Furie and Furie, 1988, for a review); membrane-bound, e.g. the receptor for the low density lipoprotein, LDL (Russell et al. 1984; Südhoff et al. 1985); or membrane-associated, e.g. some proteins of the extracellular matrix (see Engel, 1989, for review). Two common characteristics of proteins containing EGF-like repeats are their participation in protein-protein interactions and their localization in the extracellular space. Several genes encoding EGF-like proteins have been described in Drosophila: Notch (N;Wharton et al. 1985a; Kidd et al. 1986), Delta (DI; Vâssin et al. 1987; Kopczynski et al. 1988), slit (sir, Rothberg et al. 1988) and crumbs (crb;TepaB et al. 1990) and in Caenorhabditis elegans (lin-12, glp-1; Greenwald, 1985; Yochem et al. 1988; Yochem and Greenwald, 1989). These genes participate in important developmental processes and, in some cases, mutations in the genes exhibit pleiotropic phenotypes, suggesting a multiplicity of interactions with other proteins.

Using probes of the neurogenic genes N and DI, several cross-hybridizing clones were isolated (Knust et al. 1987; Rothberg et al. 1988). Two of the clones could be associated with genes necessary for the proper organization of the central nervous system (CNS) (sli;Rothberg et al. 1988) or the epidermis (crb;TepaB et al. 1990; TepaB and Knust, 1990). Here, we present the molecular characterization of the gene Serrate (Ser), the DNA of which was isolated by cross-hybridization with an N probe. Two dominant mutations of this gene, which lead to defects in the morphogenesis of the wing, could be mapped molecularly to the transcription unit. Eight revertants of the dominant mutation, SerD, were induced by X-rays and molecular lesions were found in three of them within the Ser transcription unit. The expression pattern of the protein in wing imaginal discs is in good agreement with the observed phenotype. During embryogenesis, RNA and protein show a complex expression pattern in wild type, which is not, however, correlated with any obvious phenotype in embryos deficient for the Ser gene. Ser alleles display strong phenotypic interactions with mutations in the genes N and Dl, indicating interactions at the protein level.

Drosophila stocks and mutagenesis

Flies were grown on standard medium and crosses were performed at room temperature or at 25 °C unless otherwise stated.

The following alleles were used:

References a:Lindsley and Zimm, 1985, 1990; b: gift from C. Niisslein-Volhard; c: Vassin et al. 1985; d: Vhssin and Campos-Ortega, 1987; e:Lehmann et al. 1983.

Oregon-R was used as wild-type stock. All mutations were kept over appropriate balancer chromosomes.

For the induction of revenants of SerD, males of the genotypemwh red e SerD/TM1 were irradiated with 5000 rad and crossed immediately to females of the genotype In(3R)C Tb Sb e/TM2. Flies of the F1 generation, which no longer exhibited the SerD-phenotypc (mwh red e SerRx/ Balancer) were crossed to TM1/TM6 and balanced stocks were established.

Isolation of cDNA and genomic clones and Southern blots

cDNA clones were isolated from libraries made from embryonic RNA of different developmental stages (Poole et al. 1985; Brown and Kafatos, 1988). Genomic clones were isolated from two genomic libraries, derived from the Oregon-R wild-type strain and cloned either into the EMBL-4 phage vector (Pirrotta et al. 1983; kindly provided by Dr V. Pirrotta) or into λ-DASH (kindly provided by Dr H. Vassin). Screening of libraries, radioactive labelling of DNA probes, hybridizations, preparations of phage, plasmid and genomic DNA and Southern blot analysis were essentially as described by Maniatis et al. (1989).

Subcloning, sequencing and computer analysis

Genomic fragments and cDNAs were subcloned into the Bluescript vector (Stratagene). The protocol described by Hong (1982) was used to obtain progressive deletions of subcloned fragments, which were then sequenced using a modified chain termination method (Sanger et al. 1977) with bacteriophage T7 DNA polymerase. Computer analysis was carried out on an IBM PC/AT with the DNA/Protein Sequence analysis software of J. M. Pustell/Intemational Biotechnologies, Inc., New Haven. Computer homology search was carried out with the program of Lipman and Pearson (1985) and the NBRF protein bank.

Isolation of RNA, northern blot analysis and in situ hybridization

Total RNA from staged embryos, third instar larvae, pupae, male and female flies were isolated according to the method described by Auffray and Rougeon (1980). Poly(A)+ RNA was enriched by oligo(dT)-cellulose chromatography. The RNA was separated on agarose gels, blotted to nylon membranes and hybridized as described by Vassin et al. (1987). In situ hybridizations to whole-mount embryos were performed with a digoxigenin-labeled probe essentially as described by Tautz and Pfeifle (1989).

Production of polyclonal antibodies and immunohistochemistry

A BamHI fragment encoding amino acids 642–1024 (see Fig. 1D) was subcloned into the pATH 11 expression vector (Spindler et al. 1984). After induction with 5 μg ml−1 indolylacrylic acid, the trpE-fusion protein was isolated according to the method of Rio et al. (1986). About 1 μg of protein in Freund’s complete adjuvant was used to immunize Balb/c female mice intraperitoneally. After one boost with the same amount of protein in Freund’s incomplete adjuvant, the serum was tested on western blots and by immunohistochemistry. Antibody staining with the polyclonal anti-Ser serum was done essentially as described by Tepaß et al. (1990). Pictures were taken with a Zeiss microscope equipped with Nomarski optics.

Cuticle preparations

The cuticle of embryos up to 48 h old was prepared according to the method of Van der Meer (1977) or that of Wieschaus and Niisslein-Volhard (1986). Wings were dehydrated in 80 % isopropanol and mounted directly in Hoyer’s medium.

The transcription unit at 97F encodes a putative transmembrane protein with 14 EGF-like repeats

One of the clones isolated by cross-hybridization with an N probe (N10–00) was mapped to the chromosomal position 97F on the third chromosome (data not shown). Using this clone, we isolated 11 additional, overlapping cDNAs, which yielded a composite length of 5.0kb. Restriction site mapping, blot hybridizations and sequencing permitted the unambiguous alignment of the different cDNAs and thus confirmed that all cDNAs derived from the same transcription unit (Fig. 1). Using these cDNA clones, we isolated several overlapping phage clones from genomic libraries, representing a total length of about 35 kb of genomic DNA. By hybridization of the different cDNAs to the genomic region and restriction site mapping, we mapped the transcribed part represented by the cDNAs to an interval of about 20 kb that contains at least nine major introns. Several of the intron–exon boundaries were verified by sequencing (Fig. 2).

Fig. 1.

Schematic representation of the Ser transcript and protein. (A) Alignment of 7 out of 12 cDNAs and the genomic fragment (g9229; see also Fig. 2) that were used for sequence analysis. cDNA N10–00, which was the original clone isolated under low-stringency conditions, carries a deletion of 176bp. Sequenced regions are indicated by filled bars. An, poly(A) tail. (B) Restriction map of the combined cDNAs. .The stippled region indicates the long open reading frame of 4224 nucleotides. Abbreviations used are: B, BamHI; H, HindIII; N, NruI; RV, EcoRV; X, Xba I. (C) Diagrammatic representation of the 1408 amino acid SER protein. SP, signal peptide, TM, transmembrane domain, Cys, second cysteine-rich region. Hatched region: EGF-homology region, with the insertions in the 4th, 6th and 10th repeat shown as light hatched regions. (D) The 1.1 kb BamHI fragment derived from a deletion clone of the cDNA N10–12 (see Materials and methods for further explanation), representing amino acids 642–1024, was expressed in bacteria as a trpE-fusion protein and was used to raise antibodies.

Fig. 1.

Schematic representation of the Ser transcript and protein. (A) Alignment of 7 out of 12 cDNAs and the genomic fragment (g9229; see also Fig. 2) that were used for sequence analysis. cDNA N10–00, which was the original clone isolated under low-stringency conditions, carries a deletion of 176bp. Sequenced regions are indicated by filled bars. An, poly(A) tail. (B) Restriction map of the combined cDNAs. .The stippled region indicates the long open reading frame of 4224 nucleotides. Abbreviations used are: B, BamHI; H, HindIII; N, NruI; RV, EcoRV; X, Xba I. (C) Diagrammatic representation of the 1408 amino acid SER protein. SP, signal peptide, TM, transmembrane domain, Cys, second cysteine-rich region. Hatched region: EGF-homology region, with the insertions in the 4th, 6th and 10th repeat shown as light hatched regions. (D) The 1.1 kb BamHI fragment derived from a deletion clone of the cDNA N10–12 (see Materials and methods for further explanation), representing amino acids 642–1024, was expressed in bacteria as a trpE-fusion protein and was used to raise antibodies.

Fig. 2.

Schematic representation of the cloned genomic region from 97F. (A) A collection of overlapping genomic clones isolated from different libraries. They span the translated region as well as the 3′ untranslated and at least part of the 5′ untranslated region of Ser. (B) Restriction map of the cloned region and map units. g9229 represents the 3 kb EcoRI-Nrul fragment that was used for sequence analysis. Zero indicates the end of the leftmost phage. B, BamHI; E, EcoRI; H, HindIII; X, XbaI. (C) Exon-intron structure of the Ser gene. Nine introns were identified by restriction mapping and hybridization of different cDNA fragments to genomic DNA. Exon-intron boundaries that were defined by sequencing are indicated by small dots. The stippled region corresponds to the long open reading frame. The black horizontal bars indicate the fragments in which restriction fragment length polymorphisms were detected in the DNA of Bd (map units 3.5–4.1), SerD (map unit 20.1–21.0), and three revenants of SerD, SerRX106, SerRX82 and SerRX107.

Fig. 2.

Schematic representation of the cloned genomic region from 97F. (A) A collection of overlapping genomic clones isolated from different libraries. They span the translated region as well as the 3′ untranslated and at least part of the 5′ untranslated region of Ser. (B) Restriction map of the cloned region and map units. g9229 represents the 3 kb EcoRI-Nrul fragment that was used for sequence analysis. Zero indicates the end of the leftmost phage. B, BamHI; E, EcoRI; H, HindIII; X, XbaI. (C) Exon-intron structure of the Ser gene. Nine introns were identified by restriction mapping and hybridization of different cDNA fragments to genomic DNA. Exon-intron boundaries that were defined by sequencing are indicated by small dots. The stippled region corresponds to the long open reading frame. The black horizontal bars indicate the fragments in which restriction fragment length polymorphisms were detected in the DNA of Bd (map units 3.5–4.1), SerD (map unit 20.1–21.0), and three revenants of SerD, SerRX106, SerRX82 and SerRX107.

The nucleotide sequence representing the entire length of the composite cDNAs was determined. As none of the cDNAs contained a translational start site, we sequenced part of a genomic fragment immediately 5′ to the cDNAs (g 9229 in Figs 1 and 2), and found the putative translation start site. The combined genomic and cDNA sequences reveal one long open reading frame of 4224bp (120bp genomic and 4104bp cDNA), which is followed by a 3′ non-translated region of 903 bp (Fig. 3). The primary gene product consists of 1408 amino acids and has a calculated relative molecular mass of 150×103. A hydropathy profile (Kyte and Doolittle, 1982) revealed three hydrophobic regions, a C-terminal region of 25 amino acids with characteristics of a membrane-spanning domain and an N-terminal stretch of 10 hydrophobic amino acids as part of a signal peptide (Perlman and Halverson, 1983). The third hydrophobic domain (amino acid 535 to 579) corresponds to the insertion within the 6th EGF-like repeat (see below). Thus, the protein encoded by this transcription unit represents a putative transmembrane protein, which is schematically depicted in Fig. 1. No attempts were made to determine the transcriptional start site. However, a heptanucleotide located 1.1 kb upstream of the translation start matches well (6 out of 7 nucleotides) to the consensus of Drosophila transcription start sites (ATCAG/TTC/T; Hultmark et al. 1986).

Fig. 3.

Nucleotide and deduced amino acid sequence of Ser. None of the cDNAs contained a translation start site, but two potential translation start sites were detected in the genomic fragment g9229 (see Fig. 1) 120 bp and 129 bp upstream from the 5′ end of the cDNA N10–29. The second ATG, which is preceded by the tetranucleotide CAAA and fits exactly the consensus sequence for translational start sites in Drosophila (Cavener, 1987), will be regarded as the more likely one. The hydrophobic core of the signal sequence and the transmembrane domain are shaded, the putative cleavage site of the signal peptide (arrowhead) obeys the −1, −3’ rule of von Heijne (1985). EGF-like repeats are underlined, the insertions in the 4th, the 6th and the 10th repeat are marked with broken lines. The insertion in the 10th repeat results in two additional cysteine residues, one in the insertion itself and the other in an 8 bp in frame direct repeat flanking the insertion. Several RNA-degradation signals in the 3′ untranslated region (Shaw and Kamen, 1986) and the second cysteine-rich domain (amino acid 995–1060) are indicated by dotted lines. The RGD tripeptide is boxed, as are three potential Ser/Thr-phosphorylation sites in the intracellular part. 14 possible N-linked glycosylation sites are denoted by asterisks. Two canonical polyadenylation signals (Bimstiel et al. 1985) close to the end of the sequence are underlined.

Fig. 3.

Nucleotide and deduced amino acid sequence of Ser. None of the cDNAs contained a translation start site, but two potential translation start sites were detected in the genomic fragment g9229 (see Fig. 1) 120 bp and 129 bp upstream from the 5′ end of the cDNA N10–29. The second ATG, which is preceded by the tetranucleotide CAAA and fits exactly the consensus sequence for translational start sites in Drosophila (Cavener, 1987), will be regarded as the more likely one. The hydrophobic core of the signal sequence and the transmembrane domain are shaded, the putative cleavage site of the signal peptide (arrowhead) obeys the −1, −3’ rule of von Heijne (1985). EGF-like repeats are underlined, the insertions in the 4th, the 6th and the 10th repeat are marked with broken lines. The insertion in the 10th repeat results in two additional cysteine residues, one in the insertion itself and the other in an 8 bp in frame direct repeat flanking the insertion. Several RNA-degradation signals in the 3′ untranslated region (Shaw and Kamen, 1986) and the second cysteine-rich domain (amino acid 995–1060) are indicated by dotted lines. The RGD tripeptide is boxed, as are three potential Ser/Thr-phosphorylation sites in the intracellular part. 14 possible N-linked glycosylation sites are denoted by asterisks. Two canonical polyadenylation signals (Bimstiel et al. 1985) close to the end of the sequence are underlined.

In addition, we found a putative TATA box (GTATA-TAAAG) in a distance of 30 bp further upstream. Whether this and/or a further upstream promoter is used, or whether there are even additional introns in this region, is still under investigation.

The most characteristic features of the extracellular domain are two cysteine-rich regions, the first of which is composed of 14 EGF-like units. EGF-like repeats, especially those found in other Drosophila proteins, are usually composed of about 38 amino acids and are characterized by six cysteine residues with a regular spacing and by additional conserved amino acids. The 4th, the 6th and the 10th repeat differ from this type of repeat in that they carry insertions of 68, 46 and 36 amino acids, respectively (Fig. 3). The clustered appearance of a limited variety of amino acids may reflect their origin from repetitive elements, which is actually supported, at least for the first insertion, by a 53% similarity at the DNA level to the opa repeat (Wharton et al. 1985b). The first insertion contains 11 clustered, positively charged amino acids, which might be involved in electrostatic interactions, while the insertion in the 6th repeat would allow hydrophobic interactions with other proteins. The 10th EGF-like repeat contains 13 threonine residues, which is similar to a threonine stretch found in the Drosophila glutactin (Olson et al. 1990).

The second cysteine-rich domain between the EGF-like repeats and the transmembrane domain consists of 11 cysteine residues concentrated within a fragment of 66 amino acids:

formula

where C is cysteine and Xn designates any other amino acid. Part of the spacing pattern of the cysteine residues is reminiscent of the pattern found in the N-terminal propeptide of the αl-chain of the human collagen I precursor (Chu et al. 1984) (Fig. 4A). However, the functional significance of this similarity is not known.

Fig. 4.

Sequence similarity between part of the SERRATE protein and part of the collagen pro-al(I) chain (A) and the DELTA protein (B). In A, amino acids 995–1019 of the SERRATE protein and amino acids 56–81 of the collagen pro-al(I) chain (taken from Chu et al. 1984) and, in B, amino acids 85-405 of the SERRATE protein and amino acids 24–347 of the DELTA protein (taken from Vassin et al. 1987) are aligned to give the best match. SP, signal peptide; I, identical amino acids; conservative exchanges. The region of similarity of the collagen pro-al(I) chain lies within the cleaved portion of the collagen molecule that is involved in several processes, e.g. the formation of the triplehelix (Fessler and Fessler, 1978). In B, the EGF-homology region is underlined. The overall homology within this region to DELTA is 43 % (143 out of 332 amino acids), 53% (71/124) in the EGF-like region and 36 % (72/198) in the N-terminal region.

Fig. 4.

Sequence similarity between part of the SERRATE protein and part of the collagen pro-al(I) chain (A) and the DELTA protein (B). In A, amino acids 995–1019 of the SERRATE protein and amino acids 56–81 of the collagen pro-al(I) chain (taken from Chu et al. 1984) and, in B, amino acids 85-405 of the SERRATE protein and amino acids 24–347 of the DELTA protein (taken from Vassin et al. 1987) are aligned to give the best match. SP, signal peptide; I, identical amino acids; conservative exchanges. The region of similarity of the collagen pro-al(I) chain lies within the cleaved portion of the collagen molecule that is involved in several processes, e.g. the formation of the triplehelix (Fessler and Fessler, 1978). In B, the EGF-homology region is underlined. The overall homology within this region to DELTA is 43 % (143 out of 332 amino acids), 53% (71/124) in the EGF-like region and 36 % (72/198) in the N-terminal region.

The predicted protein exhibits a region of striking similarity to the DELTA protein, encoded by one of the neurogenic genes (Vassin et al. 1987; Kopczynski et al. 1988), which extends from the putative signal peptide cleavage site to the insertion of the 4th EGF-like repeat (Fig. 4B). This region has a 43 % similarity to DELTA. It is also remarkable that the block of EGF-like repeats in both proteins starts with an incomplete repeat.

The putative intracellular part of the protein contains three possible phosphorylation sites, which fit the consensus sequence for phoshorylation by cyclic AMPdependent serine or threonine protein kinases (Krebs and Beavo, 1979; Cohen, 1985).

DNA heterogeneities associated with the mutations SerD, Bd and revenants of SerD map within the transcription unit at 97F

The transcription unit encoding the EGF-like protein described above was mapped by in situ hybridization to the chromosomal region 97F. Two mutations were known to map in this region: Serrate (Ser-, here named SerD) and Beaded (Bd), which have been reported to map meiotically at 3–92.5 and 3–93.8, respectively (Lindsley and Zimm, 1985,1990). Only one allele of Bd was available, BdG (Beaded of Goldschmidt), which has been reported to be lethal in homozygosity (Lindsley and Zimm, 1985). Using overlapping deficiencies, we could map the lethality of Bd to the chromosomal interval 97F1-98A1/2, which is uncovered by Df(3R)D605. SerD is viable over any deficiency of this region, but the close proximity to Bd suggests its localization in the same interval.

In a mutagenesis screen, we isolated eight revertants of Ser°. One of these is a cytological visible deletion (Df(3R)SerRX3; 97E7–11;97F3–11) and is embryonic lethal, whereas the seven others die as second instar larvae or pupae. By Southern blots using genomic fragments representing the entire transcribed region, we detected restriction fragment length polymorphisms in the DNAs of SerD, Bd and three revertants as compared to wild-type or the parental DNA, respectively (Fig. 2). The polymorphisms detected in the DNAs of SerD and Bd after digestion with different restriction enzymes can be explained by insertions of unknown sequences into the last exon (SerD) or into the intron following the translational start (Bd). The revertant SerRX106 is associated with an internal deletion, which eliminates about 9 kb of the transcribed region. The polymorphisms detected in the DNA of SerRX82 are likely to be caused by a small deletion of about 0.5 kb, whereas the defects mapped in SerRX107 cannot be interpreted unambiguously.

SerDand Bd are dominant gain-of-function mutations In the genetic crosses performed, both SerD and Bd behave as dominant gain-of-function mutations. These mutant phenotypes cannot be produced by simple hemizygosity. The phenotype of SerD is fully penetrant over the deficiency as well as over the duplication the locus. Bd is lethal in homozygosity and in heterozygosity over any deficiency of this region and its phenotype, although still recognizable, is reduced by the duplication.

The phenotype of SerD is characterized by notches at the tip of the wings, especially between the third and the fifth vein. In nearly all animals, hairs on the posterior margin of the wing and on the alula and some bristles at the anterior margin of the wing are missing (Fig. 5B). Penetrance and expressivity of the phenotype of the heterozygotes are slightly temperature dependent. At 18°C, the phenotypes described above are nearly fully penetrant, whereas at 28°C the phenotype is weaker, mainly with respect to the notches, and the penetrance is slightly reduced. In agreement with Belt (1971), we found that homozygous SerD flies are viable and fertile and that their phenotype is enhanced (Fig. 5C).

Fig. 5.

Wing phenotypes of (A) wild-type. (B) SerD/+. (C) SerD/SerD: the phenotype is enhanced in comparison to B, i.e. the notches are more pronounced, the third and fifth wing veins are thickened and nearly all hairs on the posterior margin and most of the bristles on the anterior margin of the wing are missing. The alula (arrowhead) is reduced in size and contains less hairs than in wild type. (D) Bd/ + (weak phenotype). (E) Bd/+ (intermediate phenotype). (F) Bd/ + (strong phenotype).

Fig. 5.

Wing phenotypes of (A) wild-type. (B) SerD/+. (C) SerD/SerD: the phenotype is enhanced in comparison to B, i.e. the notches are more pronounced, the third and fifth wing veins are thickened and nearly all hairs on the posterior margin and most of the bristles on the anterior margin of the wing are missing. The alula (arrowhead) is reduced in size and contains less hairs than in wild type. (D) Bd/ + (weak phenotype). (E) Bd/+ (intermediate phenotype). (F) Bd/ + (strong phenotype).

The phenotype of Bd heterozygotes shows a high degree of variability. The weak Bd phenotype is characterized by one or several small indentations in the wing, in most cases at the anterior margin (Fig. 5D) and resembles that of SerD. In the intermediate phenotype (Fig. 5E), the indentations are more pronounced and affect most of the first vein. In its strongest form, the first vein and most parts of the second vein are completely missing and the wing blade is reduced to about half its normal size (Fig. 5F). Bd also affects the development of the halteres, which are often reduced in size and, in the most extreme case, they are only rudimentary (not shown). In contrast to SerD, the hairs at the posterior margin of the wing and the alula are only slightly affected even in the most extreme Bd phenotype. The Bd phenotype is strongly temperature dependent, in that, at 18°C, nearly all animals show a strong phenotype. At 28°C, about 78% of the animals are wild type, most of the remaining ones exhibit a weak and only a few per cent express the intermediate phenotype. The enhancement of the Bd phenotype can be obtained by any condition that prolongs the time of development, e.g. reduced temperature or overcrowded cultures. This observation is in good agreement with the already reported sensitivity of Bd to the concomitant presence of Minute mutations, which are also known to delay development (Lindsley and Zimm, 1990).

The adult phenotype of transheterozygous Bd/SerD flies shows characteristics of both mutations (not shown). Due to the overlapping features of the Bd and SerD phenotypes, the phenotype of the transheterozygotes cannot be unambiguously ascribed to one or the other mutation. We could not observe an enhancement of any of the phenotypic traits described.

Whereas SerD is homozygous viable (see above), Bd is lethal in homozygosity as well as with a deficiency of this region. The embryonic lethality ranges between 4% (Bd/Bd) to 7% (Bd/ Df(3R)ro82b) and the remaining homo- or hemizygous animals die as first or second instar larvae. Deletion of the 97F interval (Df(3R)D605/Df(3R)ro82b) leads to embryonic lethality in homozygosity. The cuticle phenotype of the dead embryos did not show any major defects compared to wild-type embryos.

All revertants are wild type in the heterozygous condition and homozygous lethal. SerRX3, which is a deficiency of several chromosomal bands, is embryonic lethal and exhibits no obvious phenotype, whereas all others, which do not show any cytological abnormalities, are larval lethals. All revertants are lethal with a deficiency of this region and do not complement each other. With one exception (SerRX119), all revertants are lethal with Bd.

Ser alleles exhibit pronounced interactions with N and D1

Bd, SerD, revertants of SerD and the deficiency of the 97F region exhibit pronounced interactions with mutations of the neurogenic genes N and DI. The adult phenotypes of N and DI mutations have been described by Welshons (1965), Vassin et al. (1985), Vassin and Campos-Ortega (1987) and Alton et al. (1988), and are shown in Fig. 6A–C. SerD affects the phenotypes of N and of DI differently. In N/+;SerD/+ animals, both phenotypes seem to be enhanced (Fig. 6D). This effect is even more pronounced in combination with the recessive mutation notchoid (nd), in which nd becomes dominant in the presence of SerD and the phenotype is enhanced in hemizygous males (Fig. 6E). In contrast to this, SerD reduces the phenotype of the deficiency DlFX3 and. of other DI alleles (Fig. 6F).

Fig. 6.

Wing phenotypes of: (A) N8/ +, the phenotype resembles that of Bd and SerD. Note the small notches at the tip of the wing (hence the name) and the slight thickening of the third and fifth wing vein. (B) nd1 /Y, a recessive viable allele of N. shows a similiar wing phenotype in homo- or hemizygosity as N/+. (C) DTFX3/+, note the irregular broadening of the wing veins, mainly vein II and V, and Delta-like thickenings at the end of the veins. (D) N8/+\ SerD/+. The wings arc very small and strongly indented and the wing veins, especially the third and fifth vein, are thickened. Note the hairs missing in the alula and the posterior margin of the wing and most of the bristles at the anterior margin. (E) nd1/Y; SerD/ +. The phenotype is enhanced compared with B. Note the thickening of the third and fifth wing vein, which is similar to D. (F) Dl,FX3+/+ SerD. The DI phenotype is reduced compared to C. (G) N8/+; Bd/+. (H) nd1/Y; Bd/+. (I) Dl9P39+/+ Bd. This wing was obtained from an animal raised at 29°C. The wing blade is completely missing in this case.

Fig. 6.

Wing phenotypes of: (A) N8/ +, the phenotype resembles that of Bd and SerD. Note the small notches at the tip of the wing (hence the name) and the slight thickening of the third and fifth wing vein. (B) nd1 /Y, a recessive viable allele of N. shows a similiar wing phenotype in homo- or hemizygosity as N/+. (C) DTFX3/+, note the irregular broadening of the wing veins, mainly vein II and V, and Delta-like thickenings at the end of the veins. (D) N8/+\ SerD/+. The wings arc very small and strongly indented and the wing veins, especially the third and fifth vein, are thickened. Note the hairs missing in the alula and the posterior margin of the wing and most of the bristles at the anterior margin. (E) nd1/Y; SerD/ +. The phenotype is enhanced compared with B. Note the thickening of the third and fifth wing vein, which is similar to D. (F) Dl,FX3+/+ SerD. The DI phenotype is reduced compared to C. (G) N8/+; Bd/+. (H) nd1/Y; Bd/+. (I) Dl9P39+/+ Bd. This wing was obtained from an animal raised at 29°C. The wing blade is completely missing in this case.

The effects of Bd on N and DI are, like the Bd phenotype itself, strongly dependent on temperature. Flies heterozygous for N or nd in combination with Bd show a strong enhancement of probably both phenotypes (Figs 6G, H). On the other hand, Bd is embryonic lethal with amorphic DI alleles at room temperature, without showing any obvious cuticle phenotype. How-ever, at 29°C, a small per cent of transheterozygous flies were recovered that exhibit a very strong wing phenotype (Fig. 61). In addition, the tarsal segments are fused, the ocelli are enlarged and sometimes fused, and the compound eyes are smaller and rough (not shown). These phenotypic traits have also been described for antimorphic Dl alleles (Vassin and Campos-Ortega, 1987).

The deficiency of the 97F region does not change the adult N phenotype nor does nd become dominant. In contrast, the deficiency of the 97F region is lethal with the deficiency of Dl or other amorphic Dl alleles. About 4% of the progenies of these crosses die as embryos. The revenants of SerD behave differently with respect to Dl, in that six of them are viable as transheterozygotes and two are embryonic lethal. Interestingly, one of the latter two, SerRX3, is a cytological visible deficiency, which eliminates the whole gene and the other, SerRX106, carries a large deletion within the gene (see Fig. 2C).

Besides minor defects in the tracheal system observed in some of the dead embryos, which were obtained by different crosses described above, we detected no other major embryonic phenotype even after staining with tissue-specific antibodies, e.g. neural- or epidermal-specific antibodies.

Ser exhibits a complex expression pattern

The temporal expression pattern of Ser was analyzed by northern blot hybridization of cDNA clones to poly(A)+ RNA isolated from different developmental stages. In the RNA isolated from embryonic, larval and pupal stages, two transcripts of 5.6 and 5.9kb can be detected. As 8 out of a total number of 12 cDNAs isolated from embryonic and larval cDNA libraries have the same 3′ end, we suggest that the difference in size is not caused by differential termination. In RNA of adult flies, a transcript of 2.5 kb is present in both sexes, whereas in males two additional transcripts of 4.5kb and 5.2kb RNA can be detected (data not shown).

The spatial expression pattern of Ser was analyzed both by in situ hybridization of digoxigenin-labelled probes to RNA in whole-mount embryos, and by staining whole-mount embryos and imaginal wing discs with a polyclonal mouse serum that was directed against part of the extracellular domain (Fig. 1). The specificity of the serum was tested by staining an egg collection from a stock heterozygous for a deficiency of the 97F region. A total of 25 % of these embryos did not stain with the serum (data not shown).

The temporal and spatial expression patterns of the Ser RNA and protein are the same during embryonic development and will therefore be presented together. The pattern comprises domains of expression in the fore- and hindgut, the epidermis, the tracheal system, the salivary glands and the CNS. The first expression is visible at stage 11 (all stages are according to Campos-Ortega and Hartenstein, 1985) in the anlage of the clypeolabrum (Fig. 7A) and shortly after this also in two longitudinal rows of cells in the hypopharyngeal lobe (Fig. 7B). These regions will later form the roof and floor of the pharynx, respectively. Expression in the roof continues until the end of embryogenesis, while expression in the epithelium of the floor ceases after germ-band retraction (Fig. 7D). At late stage 11, expression starts in a ring of cells surrounding the stomodeum (Fig. 7B), which finally come to lie in the anterior part of the proventriculus (Fig. 7B, D and G).

Fig. 7.

Expression pattern of Ser RNA and protein. In situ hybridizations (A–F) were performed with a digoxigenin-labelled cDNA fragment (cDNA N10–29; see Fig. 1A). The antiserum used for staining (G, H) is directed against part of the extracellular domain (see Fig. 1D). (A) Early stage 11 embryo. Note the transcript in a few cells of the anlage of the clypeolabrum (cl). (B) Late stage 11 embryo. The expression domain in the epithelium of the clypeolabrum (cl) is enlarged and RNA can be clearly seen in the hypopharyngeal lobe (hy). Transcripts appear in a ring around the stomodeum (st) and in a metameric pattern in the germ band. (C) Same stage as B, showing the posterior end of the germ band. The stripes in each parasegment are separated by a gap laterally and are out of register with each other. RNA is also seen in the anlagcn of the anal pads (ap) and the posterior’ spiracles (ps). (D, E and F) Stage 14 embryos. Transcripts are visible in the main tracheal trunks (tr), the roof of the pharynx (ph), the proventriculus (pr), two rings of cells in the hindgut (hg), the anterior and posterior spiracles (as, ps), at the base of the anal pads (ap), the roof of the pharynx (ph), the frontal sac (fs), and in a striped pattern in the epidermis. (G) Stage 15 embryo, stained with the polyclonal anti-Ser antiserum. Beside the tissues stained in the stage 14 embryo (D, E, and F, same abbreviations), the protein can be detected in the duct of the salivary gland (sd) and in the commissures of the CNS (co). In A to G, anterior is left and the dorsal side up. The yolk has a brown color in all embryos. (H) Wing imaginal disc of a third instar larva, stained with the polyclonal anti-Ser antiserum. Staining can be seen in the anlagcn of the wing margin (m), in three stripes perpendicular to this and in several patches in the posterior region of the disc, one of which corresponds to the anlage of the alula (al) (according to the fate map of Bryant, 1978).

Fig. 7.

Expression pattern of Ser RNA and protein. In situ hybridizations (A–F) were performed with a digoxigenin-labelled cDNA fragment (cDNA N10–29; see Fig. 1A). The antiserum used for staining (G, H) is directed against part of the extracellular domain (see Fig. 1D). (A) Early stage 11 embryo. Note the transcript in a few cells of the anlage of the clypeolabrum (cl). (B) Late stage 11 embryo. The expression domain in the epithelium of the clypeolabrum (cl) is enlarged and RNA can be clearly seen in the hypopharyngeal lobe (hy). Transcripts appear in a ring around the stomodeum (st) and in a metameric pattern in the germ band. (C) Same stage as B, showing the posterior end of the germ band. The stripes in each parasegment are separated by a gap laterally and are out of register with each other. RNA is also seen in the anlagcn of the anal pads (ap) and the posterior’ spiracles (ps). (D, E and F) Stage 14 embryos. Transcripts are visible in the main tracheal trunks (tr), the roof of the pharynx (ph), the proventriculus (pr), two rings of cells in the hindgut (hg), the anterior and posterior spiracles (as, ps), at the base of the anal pads (ap), the roof of the pharynx (ph), the frontal sac (fs), and in a striped pattern in the epidermis. (G) Stage 15 embryo, stained with the polyclonal anti-Ser antiserum. Beside the tissues stained in the stage 14 embryo (D, E, and F, same abbreviations), the protein can be detected in the duct of the salivary gland (sd) and in the commissures of the CNS (co). In A to G, anterior is left and the dorsal side up. The yolk has a brown color in all embryos. (H) Wing imaginal disc of a third instar larva, stained with the polyclonal anti-Ser antiserum. Staining can be seen in the anlagcn of the wing margin (m), in three stripes perpendicular to this and in several patches in the posterior region of the disc, one of which corresponds to the anlage of the alula (al) (according to the fate map of Bryant, 1978).

In the hindgut, two defined regions of expression can be distinguished from stage 12 onward: one region lies just caudalwards from the insertion sites of the Malphighian tubules, and the other region lies in the posterior-most 20% of the hindgut, caudalwards to a small restriction of the hindgut (Figs 7D, G and 8A).

Expression in the epidermis starts at stage 11 in a metameric pattern, which is maintained until after germ-band retraction (Fig. 7B, Cand F). In each segment, a dorsal and a ventral stripe can be distinguished (Fig. 7C and F), which are not in register with each other. In the thoracic segments, only dorsal stripes can be detected. After germ-band retraction, RNA is very abundant in the first thoracic segment. Furthermore, RNA can be detected in the gnathal segments and in the anlagen of the anal pads (Fig. 7C, D and G).

In the tracheal system, RNA and protein are expressed from stage 13 onward in the two main lateral trunks (Figs 7E, F and 8A), except a short region just between the ends of the tracheal stems and the anlagen of the posterior and anterior spiracles. The latter start expression in stages 11 and 13, respectively, and continue until the end of embryogenesis (Fig. 7C, F and G). Expression can also be detected in the secretory ducts of the salivary glands (Fig. 7D and G) and at the ventral side of the frontal sac (from stage 14 onward; Fig. 7D and G).

The protein is also clearly visible from stage 15 onward in the anterior and posterior commissures of each segment as well as in the roots of the segmental nerves (Fig. 7G). In addition, several axons within the brain hemispheres can be stained with this antibody (not shown).

In all tissues where the SERRATE protein is expressed it is associated with the membranes. In some tissues, most obvious in the hindgut, the trachea and the proventriculus at late stages of development, the SERRATE protein is localized on the apical surface of the respective cells (Fig. 8A). In some regions, e.g. in the roof of the pharynx or in the epidermis, the protein is often found in vesicle-like structures (Fig. 8B, C).

Fig. 8.

Higher magnification of some embryonic tissues stained with the anti-SER antibody. (A) Parts of the hindgut and the trachea of a stage 15 embryo to demonstrate the localization of the SER-protein on the apical membranes of the hindgut (hg) and the trachea (tr). (B) Stage 15 embryo (same embryo as in Fig. 7G). Note the presence of vesicle-like structures in the roof of the pharynx (ph); pv, proventriculus. (C) Part of the epidermis of a stage 15 embryo. The SER-protein is expressed in narrow stripes (arrowheads) and appears in vesicle-like structures.

Fig. 8.

Higher magnification of some embryonic tissues stained with the anti-SER antibody. (A) Parts of the hindgut and the trachea of a stage 15 embryo to demonstrate the localization of the SER-protein on the apical membranes of the hindgut (hg) and the trachea (tr). (B) Stage 15 embryo (same embryo as in Fig. 7G). Note the presence of vesicle-like structures in the roof of the pharynx (ph); pv, proventriculus. (C) Part of the epidermis of a stage 15 embryo. The SER-protein is expressed in narrow stripes (arrowheads) and appears in vesicle-like structures.

In wing imaginal discs of third instar larvae, the SERRATE protein is expressed in a row of cells located across the wing pouch and in three stripes crossing the former line perpendicularly and in some regions at the border of the disc (Fig. 7H). According to the fate map established by Bryant (1978), the stained regions correspond to the future wing margin and the anlage of the alula. It is not clear whether the expression in the three stripes reflects parts of the wing vein pattern.

The transcription unit at 97F defines the gene Ser

At the beginning of our analysis, only two dominant mutations of this region were available, SerD and Bd. Both show restriction fragment length polymorphisms within the transcription unit at 97F presented in this paper (Fig. 2) and it is most likely that the mapped differences are brought about by insertions into an intron (Bd) or an exon (SerD). Since the parental chromosomes of these mutations are not available, we cannot verify that these defects cause the mutant phenotypes. In order to isolate more alleles of Ser, we induced new mutations by irradiating SerD flies, and screened the offspring for the absence of the dominant wing phenotype. Three out of eight revertants show restriction fragment length polymorphisms in the transcription unit under discussion (Fig. 2). The revertants genetically behave in the same way as the deficiency with respect to SerD, Bd and, in two cases, Dl, and can be regarded as hypomorphic or amorphic Ser alleles. Taken together, the genetic results, the expression pattern of the protein in the wing imaginal disc of third instar larvae, which is in agreement with the observed wing phenotype of both dominant mutations, and the molecular data confirm the identity between the described transcription unit at 97F and the genetically defined locus. However, final proof for the identity can only be obtained by transformation experiments.

According to the convention of Drosophila geneticists (Lindsley and Grell, 1968), the gene should be named Beaded, because this was the first name given to a mutation of this gene (Lindsley and Zimm, 1985, 1990). However, since many balancer chromosomes are marked with SerD, the name Serrate is more widely used. Thus, we would propose to make an exception to the rule and give the name Serrate to this transcription unit and the designation SerD and SerBd for the dominant alleles.

The structure of the SERRATE protein suggests its participation in protein-protein interactions

According to the sequence analysis, the putative SERRATE protein is a transmembrane protein. Staining with specific antibodies shows localization of the protein in the membrane and thus confirms this assumption. The hydrophobic core of the putative signal peptide is preceded by an unusual long stretch of 68 amino acids, among them 18 basic residues. This situation is similar to the structure of the CRUMBS protein, where 67 amino acids, among them 19 basic residues, precede the hydrophobic core (TepaB et al. 1990). Despite its length, it could act as a functional signal peptide (Rottier et al. 1987), which is further confirmed by the presence of a potential cleavage site (Fig. 3). A second hydrophobic region of 25 amino acids, the putative membrane-spanning segment, is flanked at its carboxyl side by positively charged amino acids, which are presumed to interact with the polar groups of membrane lipids at the junction between the membrane and the cytoplasmic portion of the protein.

The 14 EGF-like repeats of the SERRATE protein exhibit a striking similarity to those found in other EGF-like proteins of Drosophila, i.e. the NOTCH, DELTA, SLIT and CRUMBS proteins (Wharton et al. 1985a, Kidd et al. 1986; Vassin et al. 1987; Kopczyinzki et al. 1988; Rothberg et al. 1988; TepaB et al. 1990), and to the blood coagulation factors among the vertebrate proteins (Rees et al. 1988, and references therein). Seven of the EGF-like repeats (repeat 3, 5–8, 11 and 14) carry three conserved AsX residues, which can be regarded as putative Ca2+-binding domains. The corresponding residues in the first EGF-like repeat of the blood coagulation factor IX have been shown to be essential for Ca2+-dependent interactions with other factors of the clotting cascade (Handford et al. 1990). Thus, the corresponding repeats of the SERRATE protein are likely to be involved in protein-protein interactions, as has been shown for other members of this family, such as EGF itself (Komoriya et al. 1984) or laminin (Graf et al. 1987).

A second cysteine-rich region, located between the EGF-like repeats and the transmembrane domain, has also been described for the NOTCH protein and the C. elegans proteins LIN-12 and GLP-1 (Greenwald, 1985; Yochem et al. 1988; Yochem and Greenwald, 1989). However, except for the high amount of cysteine residues, the SERRATE protein does not show any sequence similarity to the corresponding regions of these proteins. At its C terminus, the cysteine-rich region contains the tripeptide arginine-glycine-aspartic acid (RGD). This sequence occurs also in several proteins of the extracellular matrix, and seems to be crucial for their attachment to membrane receptors, the integrins (see Ruoslahti and Pierschbacher, 1987, for review). However, the possible function of this motif in the SERRATE protein remains to be tested biochemically.

The SERRATE protein is expressed in wing discs of third instar larvae as well as during embryogenesis. Strikingly, some of the regions stained in the wing disc correspond to parts that are affected in the adult wing of SerD and SerBd, suggesting a direct correlation between the expression of the SERRATE protein and the development of parts of the wing. However, the cellular basis of the mutant phenotype is not yet known. One explanation would be cell death, similar to the process described to occur in nd wing disc development (Lindsley and Zimm, 1990).

The finding that expression begins relatively late in embryogenesis, that no major defects are detectable in embryos without any SERRATE protein (in a deletion, for example) and that embryos that only lack Ser (in the revertants) survive and only die as larvae indicate that the SERRATE protein may be mainly required during later stages of development. The lack of a major embryonic phenotype could also be explained by redundancy of functions, such that a deficit in the SERRATE protein can be compensated by another protein. An example to this is the fasciclin I protein, which is expressed on a subset of axons (Zinn et al. 1988). Whereas a deletion of this gene by itself does not produce any major defects in the embryo, the concomitant absence of the abl gene leads to severe defects in CNS development (Elkins et al. 1990).

Genetic interactions between Ser, N and Dl suggest direct protein–protein interactions

The complex phenotypic interactions of Ser alleles with N and Dl indicate a close functional relationship between these genes. The similar adult phenotypes produced by SerD and N mutations lead us to speculate that they may represent aspects of the same phenotype. In particular, the enhancement of the N phenotype and the dominant character of nd in the presence of SerD could indicate that SerD reduces the amount of functional N product. Thus, in the presence of two copies of N+, the phenotype of Ser° is similar to N/+ and, in the presence of only one N+ copy, the mutant phenotype is enhanced. This assumption is further supported by the observations that the phenotype of the duplication of the N gene, Confluens, is reduced by SerD and vice versa (data not shown).

It has been previously proposed that N and Dl may interact in a dose-dependent manner and that their phenotypes are produced by an imbalance of their gene products (Alton et al. 1989; Godt, 1990). An N to Dl ratio of 1 (+/+;+/+ or +/N-, DI/+) will thus yield a wild-type or nearly wild-type phenotype, a ratio of less than 1 (e.g. in +/N; +/+) gives the N phenotype and a ratio of more than 1 (e.g. in +/+; DI/+) determines the Dl phenotype. According to the proposed model, the combination of Ser0 with the latter genotype would reduce the functional amount of N product and thus lower the ratio, which results in the reduction of the Dl phenotype, and this was indeed observed (Fig. 6F).

So far, only interactions during postembryonic development can be described by this model. The involvement of Ser during early neurogenesis is unlikely, as the expression of Ser starts well after the decision between neuroblasts and epidermoblasts, which is controlled by the neurogenic genes (e.g. see Knust and Campos-Ortega, 1989, for review) has been made. Afterwards, however, the expression domains of Ser partly overlap those of N (Johansen et al. 1989; Kidd et al. 1989) and Dl (Vassin et al. 1987; Kopczynski and Muskavitch, 1989), which would allow interactions between these proteins to occur. Particularly, the lethality of double heterozygotes of Dl and SerBd or Dl and amorphic Ser alleles indicates considerable interactions between the two genes.

The observed phenotypical interactions may reflect interactions between the NOTCH, DELTA and SERRATE proteins at the molecular level. Recently, biochemical analysis has demonstrated that direct interactions between the NOTCH and the DELTA protein, expressed on cell membranes of Schneider cells, lead to the formation of specific aggregates of the otherwise non-adhesive cells (Fehon et al. 1990; T. Lieber, J. F. Krane, B. Hassel, J. A. Campos-Ortega and M. Young, personal communication). Sequence analysis of some point mutations in N (Hartley et al. A98T, Kelley et al. 1987) and Dl (B. Hassel and J. A. Campos-Ortega, personal communication) implies that at least some of the different EGF-like repeats have individual functions and seem to be involved in specific protein-protein interactions, which do not allow single amino acid exchange. The mutant phenotypes can be explained by disrupting specific interactions mediated by individual EGF-like repeats with different proteins at different times of development. The results obtained from our analysis indicate that the SERRATE protein is a good candidate for a molecule involved in interactions with the two other EGF-like proteins.

We thank J. A. Campos-Ortega for many helpful discussions, F. Grawe for excellent technical assistance and C. Niisslein-Volhard and the Bowling Green Stock Center for fly stocks. We thank J. A. Campos-Ortega and J. Deatrick for critical comments on the manuscript. U.T. was supported by a graduate studentship of the Graduiertenfôrderung of Nordrhein-Westfalen. This research was supported by grants and a Heisenberg fellowship to E.K. from the Deutsche Forsch-ungsgemeinschaft (Kn 250/1–3 and Kn 250/2–1).

1
Alton
,
A. K.
,
Fechtel
,
K.
,
Kopczynski
,
C. C.
,
Shephard
,
S. B.
,
Kooh
,
P. J.
and
Muskavitch
,
M. A. T.
(
1989
).
Molecular genetics of Delta, a locus required for ectodermal differentiation in Drosophila
.
Dev. Genetics
10
,
261
272
.
2
Alton
,
A. K.
,
Fechtel
,
K.
,
Terry
,
A. L.
,
Meikle
,
S. B.
and
Muskavitch
,
M. A. T.
(
1988
).
Cytogenetic definition and morphogenetic analysis of Delta, a gene affecting neurogenesis in Drosophila melanogaster
.
Genetics
118
,
235
245
.
3
Auffray
,
C.
and
Rougeon
,
F.
(
1980
).
Purification of mouse immunoglobulin heavy chain messenger RNAs from total myeloma tumor RNA
.
Eur. J. Biochem
.
107
,
303
314
.
4
Belt
,
A. L.
(
1971
).
A non-lethal allele of Serrate?
DIS
46
,
116
.
5
Birnstiel
,
M. L.
,
Busslinger
,
M.
and
Strub
,
K.
(
1985
).
Transcription termination and 3’ processing: The end is in sitel Cell
41
,
349
359
.
6
Brown
,
N. M.
and
Kafatos
,
F. C.
(
1988
).
Functional cDNA libraries from Drosophila embryos
.
J. molec. Biol
.
203
,
425
437
.
7
Bryant
,
P. J.
(
1978
).
Pattern formation in imaginai discs
.
In The Genetics and Biology of Drosophila
, vol.
2c
(ed.
M.
Ashbumer
and
T. R. F.
Wright
), pp.
230
335
.
Academic Press
,
New York
.
8
Campos-Ortega
,
J. A.
and
Hartenstein
,
V.
(
1985
).
The Embryonic Development of Drosophila melanogaster
.
Springer, Berlin Heidelberg
,
New York, Tokyo
.
9
Cavener
,
D.
(
1987
).
Comparison of the consensus sequence flanking translational start sites in Drosophila and vertebrates
.
Nucl. Acids Res
.
15
,
1353
1361
.
10
Chu
,
M. L.
,
De Wet
,
W.
,
Bernard
,
M.
,
Ding
,
J. F.
,
Morabito
,
M.
,
Myers
,
J.
and
Ramirez
,
F.
(
1984
).
Human pro-a-l(I) collagen structure reveals evolutionary conservation of a pattern of introns and exons
.
Nature
310
,
337
340
.
11
Cohen
,
P.
(
1985
).
The role of phophorylation in the hormonal control of enzyme activity
.
Eur. J. Biochem
.
151
,
439
448
.
12
Elkins
,
T.
,
Zinn
,
K.
,
Mcallister
,
L.
,
Hoffmann
,
F. M.
and
Goodman
,
C. S.
(
1990
).
Genetic analysis of a Drosophila neural cell adhesion molecule: Interaction of fasciclin I and Abelson tyrosine kinase mutations
.
Cell
60
,
565
575
.
13
Engel
,
J.
(
1989
).
EGF-like domains in extracellular matrix proteins: localized signals for growth and differentiation?
FEBS Lett
.
251
,
1
7
.
14
Fehon
,
R. G.
,
Kooh
,
P. J.
,
Rebay
,
I.
,
Regan
,
C. L.
,
Xu
,
T.
,
Muskavitch
,
M. A. T.
and
Artavanis-Tsakonas
,
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
.
15
Fessler
,
J. H.
and
Fessler
,
L. I.
(
1978
).
Biosynthesis of collagen
.
A Rev. Biochem
.
47
,
129
162
.
16
Furie
,
B.
and
Furie
,
B. C.
(
1988
).
The molecular basis of blood coagulation
.
Cell
53
,
505
518
.
17
Godt
,
D.
(
1990
).
Wechselwirkungen zwischen den neurogenen Genen bei Drosophila melanogaster
.
Ph.D. thesis
,
Cologne
.
18
Graf
,
J.
,
Iwamoto
,
Y.
,
Sasaki
,
M.
,
Martin
,
G. R.
,
Kleinman
,
H. K.
,
Robey
,
F. A.
and
Yamada
,
Y.
(
1987
).
Identification of an amino acid sequence in laminin mediating cell attachment, chemotaxis and receptor binding
.
Cell
48
,
989
996
.
19
Greenwald
,
I.
(
1985
).
lin-12, a nematode homeotic gene, is homologous to a set of mammalian proteins that includes epidermal growth factor
.
Cell
43
,
583
590
.
20
Handford
,
P. A.
,
Baron
,
M.
,
Mayhew
,
M.
,
Willis
,
A.
,
Beesley
,
T.
,
Brownlee
,
G. G.
and
Campbell
,
I. D.
(
1990
).
The first EGF-like domain from human factor IX contains a high-affinity calcium binding site
.
EMBO J
.
9
,
475
480
.
21
Hartley
,
D. A.
,
Xu
,
T.
and
Artavanis-Tsakonas
,
S.
(
1987
).
The embryonic expression of the Notch locus of Drosophila melanogaster and the implications of point mutations in the extracellular EGF-like domain of the predicted protein
.
EMBO J
.
6
,
3407
3417
.
22
Hong
,
G. F.
(
1982
).
A systematic DNA sequencing strategy
.
J. molec. Biol
.
158
,
539
549
.
23
Hultmark
,
D.
,
Klemenz
,
R.
and
Gehring
,
W.
(
1986
).
Translational and transcriptional control elements in the untranslated leader of the heat shock gene hsp22
.
Cell
52
,
429
438
.
24
Hursh
,
D. A.
,
Andrews
,
M. E.
and
Raff
,
R. A.
(
1987
).
A sea urchin gene encodes a polypeptide homologous to epidermal growth factor
.
Science
237
,
1487
1490
.
25
Johansen
,
K. M.
,
Fehon
,
R. G.
and
Artavanis-Tsakonas
,
S.
(
1989
).
The Notch gene product is a glycoprotein expressed on the cell surface of both epidermal and neuronal precursor cells during Drosophila development
.
J. Cell Sci
.
109
,
2427
2440
.
26
Kelley
,
M. R.
,
Kidd
,
S.
,
Deutsch
,
W. A.
and
Young
,
M. W.
(
1987
).
Mutations altering the structure of epidermal growth factor-like coding sequences at the Drosophila Notch locus
.
Cell
51
,
539
548
.
27
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
.
28
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
.
Molec. cell. Biol
.
6
,
3094
3108
.
29
Knust
,
E.
and
Campos-Ortega
,
J. A.
(
1989
).
The molecular genetics of early neurogenesis in Drosophila melanogaster
.
BioEssays
11
,
95
100
.
30
Knust
,
E.
,
Dietrich
,
U.
,
Tepab
,
U.
,
Bremer
,
K. A.
,
Weigel
,
D.
,
Vâssin
,
H.
and
Campos-Ortega
,
J. A.
(
1987
).
EGF homologous sequences encoded in the genome of Drosophila melanogaster, and their relation to neurogenic genes
.
EMBO J
.
6
,
761
766
.
31
Komortya
,
A.
,
Hortsch
,
M.
,
Meyers
,
C.
,
Smith
,
M.
,
Kanety
,
H.
and
Schlesinger
,
J.
(
1984
).
Biologically active synthetic fragments of epidermal growth factor: localization of a major receptor binding region
.
Proc. natn. Acad. Sci. U.S.A
.
81
,
1351
1355
.
32
Kopczynski
,
C. C.
,
Alton
,
A. K.
,
Fechtel
,
K.
,
Kooh
,
P. J.
and
Muskavitch
,
M. A. T.
(
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
,
1723
1735
.
33
Kopczynski
,
C. C.
and
Muskavitch
,
M. A. T.
(
1989
).
Complex spatio-temporal accumulation of alternative transcripts from the neurogenic gene DELTA during embryogenesis
.
Development
107
,
623
636
.
34
Krebs
,
E. G.
and
Beavo
,
J. A.
(
1979
).
Phosphorylation-Déphosphorylation of enzymes
.
A. Rev. Biochem
.
48
,
923
959
.
35
Kyte
,
J.
and
Doolittle
,
R. F.
(
1982
).
A simple method for displaying the hydropathic character of a protein
.
J. molec. Biol
.
157
,
105
132
.
36
Lee
,
D. C.
,
Rose
,
T. M.
,
Webb
,
N. R.
and
Todaro
,
G. J.
(
1985
).
Cloning and sequence analysis of a cDNA for rat transforming growth factor-α
.
Nature
313
,
489
491
.
37
Lehmann
,
R.
,
Jimenez
,
F.
,
Dietrich
,
U.
and
Campos-Ortega
,
J. A.
(
1983
).
On the phenotype and development of mutants of early neurogenesis in Drosophila melanogaster
.
Wilhelm Roux’s Arch, devl Biol
.
192
,
62
74
.
38
Lindsley
,
D. L.
and
Grell
,
E. H.
(
1968
).
Genetic variations of Drosophila melanogaster. Carnegie Inst publ no
.
627
,
Washington DC
39
Lindsley
,
D. L.
and
Zimm
,
G.
(
1985
)
The genome of Drosophila melanogaster
.
Drosophila Inform. Serv
.
62
.
40
Lindsley
,
D. L.
and
Zimm
,
G.
(
1990
).
The genome of Drosophila melanogaster
.
Drosophila Inform. Serv
.
68
.
41
Lipman
,
D. J.
and
Pearson
,
W. R.
(
1985
).
Rapid and sensitive protein similarity searches
.
Science
227
,
1435
1441
.
42
Maniatis
,
T.
,
Fritsch
,
E. F.
and
Sambrock
,
J.
, eds
(
1989
).
Molecular Cloning. A Laboratory Manual
.
Cold Spring Harbor Laboratory
,
Cold Spring Harbor, NY
.
43
Marquardt
,
H.
,
Hunkapillar
,
M.
,
Hood
,
L. E.
and
Todaro
,
G. J.
(
1984
).
Rat transforming growth factor I: Structure and relation to epidermal growth factor
.
Science
223
,
1079
1082
.
44
Olson
,
P. F.
,
Fessler
,
L. L
,
Nelson
,
R. E.
,
Sterne
,
R. E.
,
Campbell
,
A. G.
and
Fessler
,
J. H.
(
1990
).
Glutactin, a novel Drosophila basement membrane-related glycoprotein with sequence similarity to serine esterases
.
EMBO J
.
9
,
1219
1227
.
45
Perlman
,
D.
and
Halverson
,
H. O.
(
1983
).
A putative signal peptidase recognition site and sequence in eukaryotic and prokaryotic signal peptides
.
J. molec. Biol
.
167
,
391
409
.
46
Pirrotta
,
V.
,
Hadfield
,
C.
and
Pretorius
,
G. H. J.
(
1983
).
Microdissection and cloning of the white locus and the 3B1-3C2 region of the Drosophila X chromosome
.
EMBO J
.
2
,
927
934
.
47
Poole
,
S. J.
,
Kauvar
,
L. M.
,
Drees
,
B.
and
Kornberg
,
T.
(
1985
).
The engrailed locus of Drosophila: Structural analysis of an embryonic transcript
.
Cell
40
,
37
43
.
48
Rees
,
D. J. G.
,
Jones
,
I. M.
,
Hanford
,
P. A.
,
Walters
,
S. J.
,
Smith
,
K. J.
and
Brownlee
,
G. G.
(
1988
).
The role of β-hydroxyaspartate and adjacent carboxylate residues in the first EGF domain of human factor IX
.
EMBO J
.
7
,
2053
2061
.
49
Rio
,
D. C.
,
Laski
,
F. A.
and
Rubin
,
G.
(
1986
).
Identification and immunochemical analysis of biologically active Drosophila P element transposase
.
Cell
44
,
21
32
.
50
Rothberg
,
J. M.
,
Hartley
,
D. A.
,
Walter
,
Z.
and
Artavanis-Tsakonas
,
S.
(
1988
).
slit: an EGF-homologous locus of Drosophila melanogaster involved in the development of the embryonic central nervous system
.
Cell
55
,
1047
1059
.
51
Rottier
,
P. M. J.
,
Florkiewicz
,
R. Z.
and
Rose
,
J. K.
(
1987
).
An internalized amino-terminal signal sequence retains full activity in vivo but not in vitro
.
J. biol. Chem
.
262
,
8889
8895
.
52
Ruoslahti
,
E.
and
Pierschbacher
,
M. D.
(
1987
).
New perspectives in cell adhesion: RGD and integrins
.
Science
238
,
491
497
.
53
Russell
,
D. W.
,
Schneider
,
W. J.
,
Yamamoto
,
T.
,
Luskey
,
K. L.
,
Brown
,
M. S.
and
Goldstein
,
J. L.
(
1984
).
Domain map of the LDL receptor: sequence homology with the epidermal growth factor
.
Cell
37
,
577
585
.
54
Sanger
,
F.
,
Nicklen
,
S.
and
Coulson
,
A. R.
(
1977
).
DNA sequencing with chain termination inhibitors
.
Proc. natn. Acad. Sci. U.S.A
74
,
5463
5467
.
55
Shaw
,
G.
and
Kamen
,
R.
(
1986
).
A conserved Ao sequence from the 3’untranslated region of GM-CSF mRNA mediates selective mRNA degradation
.
Cell
46
,
659
667
.
56
Spindler
,
K. R.
,
Rosser
,
D. S. E.
and
Berk
,
A. J.
(
1984
).
Analysis of adenovirus transforming proteins from early regions 1A and IB with antisera to inducible fusion antigens produced in Escherichia coli
.
J. Virol
.
49
,
132
141
.
57
Südhoff
,
T. C.
,
Goldstein
,
J. L.
,
Brown
,
M. S.
and
Russel
,
D. W.
(
1985
).
The LDL receptor gene: A mosaic of exons shared with different proteins
.
Science
228
,
815
822
.
58
Suyemitsu
,
T.
,
Asami-Yoshizumi
,
T.
,
Noguchi
,
S.
,
Tonegawa
,
Y.
and
Ishihara
,
K.
(
1989
).
The exogastrula-inducing peptides in embryos of the sea urchin, Anthocidaris crassispina. isolation and determination of the primary structure
.
Cell Diff. Dev
.
26
,
53
66
.
59
Tautz
,
D.
and
Pfeifle
,
C.
(
1989
).
A non-radioactive in situ hybridization method for the localization of specific RNAs in Drosophila embryos reveals a translational control of the segmentation gene hunchback
.
Chromosoma
98
,
81
85
.
60
Tepab
,
U.
and
Knust
,
E.
(
1990
).
Phenotypic and developmental analysis of mutations at the crumbs locus, a gene required for the development of epithelia in Drosophila melanogaster
.
Roux’s Arch. devl. Biol
.
199
,
189
206
.
61
Tepab
,
U.
,
Theres
,
C.
and
Knust
,
E.
(
1990
).
The Drosophila gene crumbs encodes an EGF-like protein, expressed on apical membranes of epithelial cells and required for organization of epithelia
.
Cell
61
,
787
799
.
62
Vassin
,
H.
,
Bremer
,
K. A.
,
Knust
,
E.
and
Campos-Ortega
,
J. A.
(
1987
).
The neurogenic locus Delta of Drosophila melanogaster is expressed in neurogenic territories and encodes a putative transmembrane protein with EGF-like repeats
.
EMBO J
.
6
,
3431
3440
.
63
Vassin
,
H.
and
Campos-Ortega
,
J. A.
(
1987
).
Genetic analysis of Delta, a neurogenic gene of Drosophila melanogaster
.
Genetics
116
,
433
445
.
64
Vâssin
,
H.
,
Vielmetter
,
J.
and
Campos-Ortega
,
J. A.
(
1985
).
Genetic interactions in early neurogenesis of Drosophila melanogaster
.
J. Neurogenet
.
2
,
291
308
.
65
Van Der Meer
,
J.
(
1977
).
Optical clean and permanent mount preparations for phase contrast microscopy of cuticular structures of insect larvae
.
Drosophila Inform. Serv
.
52
,
160
.
66
Von Heune
,
G.
(
1985
).
Signal Sequences. The limits of variation
.
J. molec. Biol
.
184
,
99
105
.
67
Welshons
,
W. J.
(
1965
).
Analysis of a gene in Drosophila
.
Science
150
,
1122
1129
.
68
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
.
69
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
.
70
Wieschaus
,
E.
and
Nüsslein-Volhard
,
C.
(
1986
).
Looking at embryos
.
In Drosophila A Practical Approach
(ed.
D. B.
Roberts
), pp.
199
227
.
IRL Press
,
Oxford, Washington DC
.
71
Yochem
,
J.
and
Greenwald
,
I.
(
1989
).
glp-1 and lin-12, genes implicated in distinct cell-cell interactions in C. elegans, encode similar transmembrane proteins
.
Cell
58
,
553
563
.
72
Yochem
,
J.
,
Weston
,
K.
and
Greenwald
,
I.
(
1988
).
The Caenorhabditis elegans lin-12 gene encodes a transmembrane protein with overall similarity to Drosophila Notch
.
Nature
335
,
547
550
.
73
Zinn
,
K.
,
Mcallister
,
L.
and
Goodman
,
C. S.
(
1988
).
Sequence analysis and neuronal expression of fasciclin I in grasshopper and Drosophila
.
Cell
53
,
577
587
.