Drosophila has eight Hox proteins, and they require factors acting in parallel to regulate different segmental morphologies. Here we find that the Drosophila gene split ends (spen), has a homeotic mutant phenotype, and appears to encode such a parallel factor. Our results indicate that spen plays two important segment identity roles. One is to promote sclerite development in the head region, in parallel with Hox genes; the other is to cooperate with Antennapedia and teashirt to suppress head-like sclerite development in the thorax. Our results also indicate that without spen and teashirt functions, Antennapedia loses its ability to specify thoracic identity in the epidermis. spen transcripts encode extraordinarily large protein isoforms (approx. 5,500 amino acids), which are concentrated in embryonic nuclei. Both Spen protein isoforms and Spen-like proteins in other animals possess a clustered repeat of three RNP (or RRM) domains, as well as a conserved motif of 165 amino acids (SPOC domain) at their C-termini. Spen is the only known homeotic protein with RNP binding motifs, which indicates that splicing, transport, or other RNA regulatory steps are involved in the diversification of segmental morphology. Previous studies by Dickson and others (Dickson, B. J., Van Der Straten, A., Dominguez, M. and Hafen, E. (1996). Genetics 142, 163-171) identified spen as a gene that acts downstream of Raf to suppress Raf signaling in a manner similar to the ETS transcription factor Aop/Yan. This raises the intriguing possibility that the Spen RNP protein might integrate signals from both the Raf and Hox pathways.

The homeotic genes in Drosophila can be divided into three classes: (1) genes of the homeotic selector, or Hox class, which are located in the Antennapedia and Bithorax complexes, (2) genes of the Polycomb (Pc) and trithorax (trx) classes, which regulate the spatial and temporal location of Hox transcription, (3)genes of the Hox modulator class, such as extradenticle, which act in parallel with Hox genes to regulate downstream targets. Hox genes are activated in distinct metameres on the anterior-posterior body axis and morphologically diversify those serially arranged fields of cells (Lewis, 1978; McGinnis and Krumlauf, 1992). The Hox proteins do this by their differential regulation of batteries of downstream ‘realizator’ genes (Garcia-Bellido, 1977; Pradel and White, 1998). Despite much research, the basis of differential realizator gene expression by Hox proteins is illuminated by only a few facts of unknown generality (Chan et al., 1997; Biggin and McGinnis, 1997; Pradel and White, 1998; Li and McGinnis, 1999; Li et al., 1999), and the basis of morphological differentiation under Hox control is even more obscure.

The ability of Hox proteins to differentially regulate batteries of downstream genes and thereby activate different developmental pathways requires a variety of modulatory factors that contribute to the binding occupancy and/or activity of Hox proteins on different target sequences. Such factors include the Hox proteins themselves, which can phenotypically suppress the activity of other family members, usually the posterior Hox function dominating the more anterior (Gonzalez-Reyes et al., 1990). Recent work by Capovilla and Botas (1998) suggests that this phenotypic suppression involves a short range enhancer repression mechanism (Gray et al., 1999). A few other Drosophila modulatory factors are apparently also involved in the specificity of Hox output, as evidenced by their homeotic mutant phenotypes in a wild-type genetic background. These include the DNA binding proteins extradenticle (exd) (Peifer and Weischaus, 1990; Rauskolb et al., 1993), teashirt (tsh) (Fasano et al., 1991; Alexandre et al., 1996), cap’n’collar B (cncB) (Mohler et al., 1995; McGinnis et al., 1998) and homothorax (hth) (Rieckhof et al., 1997; Pai et al., 1998).

Exd and its mammalian homologs in the Pbx class are homeodomain proteins (Flegel et al., 1993; Rauskolb et al., 1993) which participate in the selection and regulation of many Hox targets by cooperative binding with Hox proteins and concerted activity with them once bound (van Dijk and Murre, 1994; Chan and Mann, 1996; Mann and Chan, 1996; Chan et al., 1996; Pinsonneault et al., 1997; Li et al., 1999). The ability of Exd protein to enter nuclei and interact with Hox proteins is dependent on Hth, another Drosophila homeodomain protein (Rieckhof et al., 1997; Kurant et al., 1998; Pai et al., 1998). This activity of Hth, in combination with its pattern of expression, generates cellular patterns of nuclei that either contain or lack Exd protein. In cells that accumulate Exd in nuclei, Hox proteins are capable of serving as transcriptional activators, but in cells that lack nuclear Exd, Hox proteins have only been observed to function as repressors (Pinsonneault et al., 1997; Li et al., 1999). Tsh is a zinc-finger DNA binding protein, and is required in combination with the trunk HOX proteins to repress head-specific developmental pathways and promote trunk identity (Fasano et al., 1991; Röder et al., 1992; Alexandre et al., 1996). For example, Tsh acts with Antp to repress the ectopic activation of the anterior-most Hox gene, labial (lab; Röder et al., 1992). Finally, the CncB protein, a bZIP class of transcription factor, functions as a Hox modulator by preventing the Deformed Hox protein from activating maxillary-specific targets in the mandibular segment (Mohler et al., 1995; McGinnis et al., 1998).

Using genetic modifier screens, we have identified additional factors which influence the function of the Deformed (Dfd) gene in Drosophila, focusing on those that act in parallel to Dfd and other Hox genes (Harding et al., 1995; Gellon et al., 1997; Florence and McGinnis, 1998). The normal function of Dfd is required for the morphogenesis of structures arising from the mandibular and maxillary head segments, dorsal ridge, optic lobe and central nervous system (McGinnis et al., 1990; Restifo and Merrill, 1994). Our modifier screens identified mutant alleles of both exd and cncB, but also identified a gene with a particularly interesting mutant phenotype in which the ventral regions of thoracic (and occasionally abdominal) segments are homeotically transformed to head-like skeletal structures. We previously mapped this gene to the tip of the left arm of chromosome 2, at 21B4-6, and called it polycephalon (Gellon et al., 1997). However, we recently found that our polycephalon alleles disrupt the previously named split ends (spen) gene (Kolodziej et al., 1995), which was originally mapped to the right arm of the 2nd chromosome. We will hereafter use the name spen to denote the gene. Mutant alleles in spen [called E(Raf)2A] also were isolated as enhancers of a dominant Raf-induced eye phenotype by Dickson et al. (1996), and yet other alleles of spen [called E(E2F)2A or poc) were isolated as enhancers of an eye phenotype induced by overexpression of dE2F and dDp (Staehling-Hampton et al., 1999).

The homeotic phenotypes in spen mutant embryos are not associated with changes in Hox gene expression, nor is the spen expression pattern changed in Hox mutants, therefore spen appears to act in parallel with Hox segmental identity functions. In addition, over-expression of the thoracic determinants Antp or tsh can suppress the ectopic sclerites in spen mutants, and thus we propose that spen acts in combination with Antp and Tsh pathways to regulate the distinction between head and trunk identity. The spen gene encodes very large protein isoforms (approx. 5,500 amino acids) with an arrangement of RNP motifs and conserved C-terminal regions that can be used to identify Spen orthologs and paralogs in D. melanogaster, C. elegans and H. sapiens.

As members of the RNP motif family exert their diverse functions by sequence-specific binding to RNA (Burd and Dreyfuss, 1994), the Spen proteins may modulate Hox function through a regulatory influence on messenger RNAs produced from Hox downstream genes.

Strains and genetics

Most mutations, aberrations and abbreviations are described in Lindsley and Zimm, (1992), or in FlyBase (http://flybase.bio.indiana.edu). Other alleles and lines include: spenpoc361 and spenpoc231 (Gellon et al., 1997), spenE(Raf)2A16H1 and spenE(Raf)2A1401 (Dickson et al., 1996; these are listed as polycephalon alleles in the current FlyBase), spenE(CycE)D57 and spenE(CycE)e9 alleles (Christian Lehner, personal communication), Df(2L)TE21A, Dll5, cncVL110, lab1, oc2, ems3, btd1, Antp25, Antphs.PHTA, ScrC1AntpNsRC3, Dfd16, tsh8, tshScer/UAS.cGa, 69B-GAL4 (Brand and Perrimon, 1993; Chou et al., 1993), P{ry+t7.2=neoFRT}40-A (‘FRT-40A’), P{w+mC=ovoD1-18}2L1P{w+mC=ovoD1-18}2L2P{ry+t7.2=neoFRT} 40-A (‘ovoD1-FRT-2L’), P{ry+7.2=hsFLP}12; CyO/nocSco. The PZ1295 line [also known as l(2)03350)] was obtained from the Bloomington Stock Center. The l(2)K06805, l(2)K07612, l(2)K07721, l(2)K08102, l(2)K10325 and l(2)K13625 lines were obtained from the Berkeley Drosophila Genome Project (http://www.fruitfly.org). Embryonic stages are as described by Campos-Ortega and Hartenstein (1985).

Preparation of embryonic cuticles

Mutant chromosomes were outcrossed to eliminate balancer chromosomes from the stocks before mutant cuticles were collected. Embryos were collected for approximately 12 hours and aged for more than 24 hours before preparing cuticles as described by Wieschaus and Nüsslein-Volhard (1986). Overexpression of tsh in a spen mutant background was accomplished by crossing spenpoc361/CyO;UAStsh/+ females to spenpoc361/+;69B-GAL4/+ males. Overexpression of Antp in a spen mutant background was accomplished by crossing spenpoc361/+;hsAntp/+ females to Df(2L)TE21A/+;hsAntp/+ males, collecting embryos for 1 hour intervals, aging at 25°C for 3 hours, and then heat shocking at 37°C for 20 minutes every 2 hours for a total of three heat treatments. Embryos were then aged at 25°C for at least an additional 22 hours.

Generation of maternal germline spen/spen clones

Maternal germline spen mutant clones were generated using the FLP-FRT system as detailed by Chou et al. (1993). spenpoc231 and spenpoc361 were recombined onto the FRT 40A chromosome. spen FRT 40A/CyO virgin females were then mated to hs-FLP12; ovoD1-FRT-2L/CyO males; progeny of this cross were heat shocked for 1 hour at 37°C at 3, 4, 5 and 6 days after egg deposition. The heat-shocked female progeny carrying both the spen and ovoD1 chromosomes were mated to males heterozygous for the other spen allele (e.g. spenpoc361/ovoD1 females were mated to spenpoc231/CyO,wg males).

Identification of P-element insertions and plasmid rescue

Genomic DNA from the PZ1295 line corresponding to l(2)03350 was prepared by standard protocols as described in Gellon et al. (1997). Rescue of the DNA surrounding the P-element insertion site was accomplished by digestion, ligation and transformation of the genomic DNA as described in Gellon et al. (1997).

Excision lines were generated from PZ1295 by using the Δ2-3 strain to mobilize the element (Robertson et al., 1988). Genomic DNA was prepared from homozygous viable and homozygous lethal excision lines and mapped by Southern analysis using probes surrounding the region obtained by plasmid rescue. A similar analysis was subsequently done to map the P insertion sites in the l(2)K06805, l(2)K07612, l(2)K07721, l(2)K08102, l(2)K10325 and l(2)K13625 alleles.

Isolation of genomic clones and cDNAs

We initiated a walk in the region surrounding the insertion point of PZ1295 using a 3.2 kb HindIII/XbaI genomic fragment obtained by plasmid rescue. Approximately 50 kb of overlapping λ clones were obtained from an EMBL-3 iso-1 genomic library provided by J. Tamkun using standard protocols and nick translated probes. An additional 30 kb proximal to our walk was kindly provided by M. Noll (λ clones y17-3, y16-1, y15-1).

Poly(A)+ RNA was prepared from 0-to 2-hour, 2-to 8-hour, 8 to 12-hour and 12-to 24-hour embryos, purified once over an oligo-dT column, electrophoresed in a formaldehyde gel and blotted according to standard procedures. Northern blots were hybridized with 32P-labeled DNA probes.

cDNAs were isolated from an oligo-T primed 8-to 12-hour embryonic cDNA library provided by Nick Brown (Brown and Kafatos, 1988), and a randomly primed 0-to 18-hour embryonic λGT10 cDNA library (Clontech). Probes were labeled by nick translation. Hybridization and washes were done at high stringency following standard protocols. More than 100 overlapping partial cDNAs corresponding to the putative spen transcript were isolated using genomic and cDNA fragments as 32P-labeled probes from the region shown in Fig. 2. Thirty cDNA clones were subcloned and mapped.

Fig. 2.

(A) A molecular map of the spen locus. The top, thin line represents genomic DNA; the restriction sites are marked, BamHI (B), EcoRI (R), SalI (S). A size marker of 8 kb is shown in upper right. The location of P-inserts which fail to complement spen alleles are indicated with open triangles. The locations of transcripts within this region are shown, with arrows indicating extent and orientation of each. The intron/exon structures of the spenL and spenS transcription units are shown below the transcript arrow. Alternative 5′ spen exons are shown in light gray, and these include translation start codons and short ORFs that join the long common ORF. Common exons are shown in black. The stop codon for the predicted protein is indicated (TAA), and the positions of domains conserved with human and C. elegans homologs are shown below the sequence in dark gray (see text and Fig. 4 for descriptions of labels). (B) Northern blots of poly(A)+ embryonic RNA from selected ages (0-2 hours, 2-8 hours, 8-12 hours, 12-24 hours) probed with two DNA fragments (shown above the cDNA and genomic structures by 1 and 2). Probe 1 was generated from spenL-specific cDNA sequences, and hybridizes to a single band of approx. 20 kb (M, size marker lane). The spenL transcript is not detected in 0-2 hour embryos when only maternal transcripts are present. The spenL transcript levels are abundant at 2-8 hours and 8-12 hours of development, and less abundant at 12-24 hours. Probe 2 was generated from a genomic fragment in the common region, and detects a broader band of approx. 20 kb in 2-24 hour embryos. The common probe detects spen transcripts in 0-to 2-hour embryos, presumably due to maternal transcription from the spenS promoter.

Fig. 2.

(A) A molecular map of the spen locus. The top, thin line represents genomic DNA; the restriction sites are marked, BamHI (B), EcoRI (R), SalI (S). A size marker of 8 kb is shown in upper right. The location of P-inserts which fail to complement spen alleles are indicated with open triangles. The locations of transcripts within this region are shown, with arrows indicating extent and orientation of each. The intron/exon structures of the spenL and spenS transcription units are shown below the transcript arrow. Alternative 5′ spen exons are shown in light gray, and these include translation start codons and short ORFs that join the long common ORF. Common exons are shown in black. The stop codon for the predicted protein is indicated (TAA), and the positions of domains conserved with human and C. elegans homologs are shown below the sequence in dark gray (see text and Fig. 4 for descriptions of labels). (B) Northern blots of poly(A)+ embryonic RNA from selected ages (0-2 hours, 2-8 hours, 8-12 hours, 12-24 hours) probed with two DNA fragments (shown above the cDNA and genomic structures by 1 and 2). Probe 1 was generated from spenL-specific cDNA sequences, and hybridizes to a single band of approx. 20 kb (M, size marker lane). The spenL transcript is not detected in 0-2 hour embryos when only maternal transcripts are present. The spenL transcript levels are abundant at 2-8 hours and 8-12 hours of development, and less abundant at 12-24 hours. Probe 2 was generated from a genomic fragment in the common region, and detects a broader band of approx. 20 kb in 2-24 hour embryos. The common probe detects spen transcripts in 0-to 2-hour embryos, presumably due to maternal transcription from the spenS promoter.

A short region (150 bp) within the 4.8 kb BamHI-EcoRI fragment that spans most of the 3′ half of the large exon of the putative spen transcript sequence was not included in any of the cDNA clones. RT-PCR was used to determine that this sequence is included in spen transcripts. Poly(A)+ RNA, isolated as described above, was used as a template for reverse transcriptase following the protocols included with the enzyme, Superscript II Reverse Transcriptase (GibcoBRL). The reaction was primed with a sequence-specific 3′ primer. PCR was performed using the 1st strand reaction as template and following the protocol included with the RT enzyme (GibcoBRL).

Sequencing and database searches

Sequencing was performed mainly at the Center for AIDS Research CORE Facility at UCSD. The BLAST algorithm (Altschul et al., 1997) was used to search protein and EST databases for sequences homologous to short fragments of protein sequence. The 3 RNP-type RNA binding domains were identified in this manner. Comparison of the Spen protein sequences (accession no. AF188205) to the ISREC PROSITE database identified five bipartite nuclear localization motifs. Searches of genomic and EST databases revealed cDNA and genomic sequences which are predicted to encode the protein domains shown in Fig. 4; from D. melanogaster (accession no. for short form –AAD38639), Homo sapiens Spen-like (accession no. for long form –AL034555, for short form – EST sequence AI621042, and genomic L13434), H. sapiens RNP2 from U1a protein (accession no. 2554638), Caenorhabditis elegans (accession no. for long form – CAA91320; for short form – AAC19192) and Nicotiana tabacum (accession no. U90212), Xenopus laevis (accession no. P20397).

Fig. 4.

Spen and the family of predicted Spen-like proteins. (A) The Spen orthologs from human and worm are all larger than 2738 amino acids, and each contain three RNP domains, bipartite nuclear localization sequences (NLS), and a novel, conserved domain, the Spen Paralog and Ortholog C-terminal domain (SPOC domain). Additional similarity among Spen orthologs consists of an acidic/basic domain (EK) and glutamine-rich regions (Poly Q). An additional group of proteins, the Short Spen Like Proteins (SSLPs) are encoded in fly, human and nematode genomes (Materials and Methods). These SSLPs are all smaller than 800 amino acids and possess Spen-like RNP domains, a SPOC domain, and an RGG motif. (B) Amino acid alignment of the three RNP domains from the six Spen family proteins, along with similar RNPs from other families. The first RNP (RNPa) is more similar to the first RNP of Nucleolin than to other RNP domains. The RNPb domain has more identity to an RNP domain of AC Binding Factor (ACBF) than to other RNP domains. ACBF is a tobacco protein which binds to double stranded, AC-rich DNA sequences (Séguin et al., 1997). The schematic structure of an RNP domain from human snRNP U1A is shown at the RNP1 (red) sequences are boxed in all sequences. Hs, Homo sapiens; Dm, D. melanogaster; Ce, C. elegans; Xl, Xenopus laevis, Nt, Nicotiana tabacum. Accession numbers are provided in the Materials and Methods. (C). Amino acid alignment of the SPOC domains from fly, human and nematode. A nucleotide substitution (TGC to TAC) in spenE(CycE)e9changes the conserved cysteine (C) to tyrosine (Y) at amino acid 5444, while a substitution (GGC to GAC) in spenE(CycE)D57 changes the invariant glycine (G) to aspartate (D) at amino acid 5480.

Fig. 4.

Spen and the family of predicted Spen-like proteins. (A) The Spen orthologs from human and worm are all larger than 2738 amino acids, and each contain three RNP domains, bipartite nuclear localization sequences (NLS), and a novel, conserved domain, the Spen Paralog and Ortholog C-terminal domain (SPOC domain). Additional similarity among Spen orthologs consists of an acidic/basic domain (EK) and glutamine-rich regions (Poly Q). An additional group of proteins, the Short Spen Like Proteins (SSLPs) are encoded in fly, human and nematode genomes (Materials and Methods). These SSLPs are all smaller than 800 amino acids and possess Spen-like RNP domains, a SPOC domain, and an RGG motif. (B) Amino acid alignment of the three RNP domains from the six Spen family proteins, along with similar RNPs from other families. The first RNP (RNPa) is more similar to the first RNP of Nucleolin than to other RNP domains. The RNPb domain has more identity to an RNP domain of AC Binding Factor (ACBF) than to other RNP domains. ACBF is a tobacco protein which binds to double stranded, AC-rich DNA sequences (Séguin et al., 1997). The schematic structure of an RNP domain from human snRNP U1A is shown at the RNP1 (red) sequences are boxed in all sequences. Hs, Homo sapiens; Dm, D. melanogaster; Ce, C. elegans; Xl, Xenopus laevis, Nt, Nicotiana tabacum. Accession numbers are provided in the Materials and Methods. (C). Amino acid alignment of the SPOC domains from fly, human and nematode. A nucleotide substitution (TGC to TAC) in spenE(CycE)e9changes the conserved cysteine (C) to tyrosine (Y) at amino acid 5444, while a substitution (GGC to GAC) in spenE(CycE)D57 changes the invariant glycine (G) to aspartate (D) at amino acid 5480.

RNA interference

RNAi was performed essentially as described by Kennerdell and Carthew (1998). A 972 bp DNA fragment from the region that encodes the 3′ UTR of the spen transcript (from 11 nucleotides to 983 nucleotides downstream of the Spen TAA stop codon) was transcribed with T7 RNA polymerase to generate RNA copies of both strands. The resulting double-stranded RNA was injected into the middle region of embryos less than one hour after egg deposition. Controls consisted of injected buffer at the same site and injection of cap’n’collar B (cncB) RNA. cncB loss-of-function mutants have an obvious homeotic phenotype in embryos (McGinnis et al., 1998) that is mimicked by injection of ds cncB-specific RNA (A.Veraksa and W. M., unpublished results). Injected embryos were aged and their cuticles prepared after removing the vitelline membranes by hand.

Mutation detection

Point mutations in spen open reading frames were detected by denaturing gel electrophoresis of 400-500 base pair fragments of genomic DNA isolated from spen heterozygotes and parental strains. Genomic DNA was purified using the QIAamp Tissue Kit (QIAGEN), and PCR was performed using primers designed to produce fragments with the melting behavior described by Myers et al. (1987). Melting point variations resulting from point mutations in the fragments were detected using a Denaturing Gel Electrophoresis System (BioRad). Fragments which showed melting point variation were sequenced on both strands to determine the specific nucleotide position that was changed in mutant chromosomes.

Antibodies and in situ hybridization

A GST-Spen fusion protein containing amino acids 3203-3714 from the common region of the SpenL protein was injected into mice to generate anti-Spen antiserum. The fusion protein was bacterially expressed and purified using glutathione-Sepharose (Pharmacia) and standard techniques. For staining with this antibody, embryos were collected, dechorionated and fixed for 15 minutes in 4% formaldehyde. The antiserum was used at a 1:100 dilution, and detected with FITC-labeled anti-mouse secondary antibody (Jackson Labs).

In situ hybridizations to detect transcript expression were performed using a variation of the protocol described by Tautz and Pfeifle (1989). DIG-labeled antisense probes were made from the following cDNA sequences: Dfd probe was generated from a 2.3 kb EcoRV-SmaI fragment isolated from pcDfd41 (Regulski et al., 1985), cloned into the EcoRV site of pBluescript, linearized with BamHI and transcribed from the T3 promoter. Scr probe was generated from a 2.1 kb EcoRI-XbaI genomic fragment containing the homeobox (Kuroiwa et al., 1985), subcloned into pGEM2, linearized with EcoRI and transcribed from the T7 promoter. Antp probe was generated from a 2.2 kb BamHI-EcoRI fragments from p903 (Hafen et al., 1983) subcloned into pBluescript, linearized with BamHI and transcribed from the T7 promoter. spen antisense transcript probes were generated from a 5.2 kb cDNA fragment in clone pc1.65BH10-5.2R that included the RNP coding region subcloned into pBluescript, linearized with NotI and transcribed from the T3 promoter.

spen mutants develop head-like skeletal features in the trunk

We previously reported the isolation of six EMS-induced mutant alleles of spen based on its interaction with the Hox gene Dfd (spen was called polycephalon by Gellon et al., 1997). Some of these alleles behave as genetic nulls for the cuticular phenotype, i.e. spen361/spen361 cuticles are indistinguishable from spen361/Df(2L)TE21A(deletes the entire spen locus), and spen231 is slightly weaker than spen361. Loss of zygotic spen function results in embryonic lethality that is associated with a loss of the anterior portion of the H-piece, and with a kinked median tooth (Fig. 1B,F, compare to Fig. 1A). The anterior parts of the H-piece are derived from the ventral maxillary segment, and are dependent on Dfd function (Merrill et al., 1987; Regulski et al., 1987). Head development is profoundly disrupted in maternal/zygotic spen mutants. These mutant embryos have non-involuted heads that are missing many head sclerites, including the base of the mouth hooks, median tooth, anterior regions of the lateralgräten, and the dorsal bridge. The ventral arms and vertical plates are also strongly reduced (Fig. 1D,F,). Many of these spen-dependent sclerites are also dependent on head Hox genes such as labial (lab), and Sex combs reduced (Scr), gap/homeotic genes such as empty spiracles (ems) and orthodenticle (otd), and other head patterning genes (Merrill et al., 1989; Sato et al., 1985; Pattatucci et al., 1991; Dalton et al., 1989; Finkelstein and Perrimon, 1991). In the maternal/ zygotic mutants, the trunk region of the embryos appears to be largely unaffected. The denticle belts and posterior structures such as filzkörper appear normally shaped, although these structures, as well as the remnants of head skeleton (and ecopic head-like sclerites, see below), are somewhat less pigmented/ sclerotized than normal. The overall low level of sclerotization may be due to embryonic death in maternal/zygotic mutants before the cuticle is fully developed (Fig. 1D).

Fig. 1.

spen is required for thorax and head development. (A) Wild-type cuticle showing normal head skeleton and thoracic segments. T2, second thoracic segment. Anterior is left and dorsal is up in this and all figures, except as noted. (B) spenpoc361/Df(2L)TE21A mutant phenotype with weak expressivity. The anterior H-piece is lost (arrowhead), and ectopic sclerites develop in the middle of T2 and T3 (arrows). (C) spenpoc361/Df(2L)TE21A mutant phenotype with strong expressivity. The ectopic sclerites cover more of the naked epidermis and are found in more segments, including T1, A1 and A2. Abdominal segments develop ectopic sclerites in lateral regions. (A)A maternal/zygotic spen mutant in which the germ line was homozygous for spenpoc361, half of these maternal mutant embryos then received the spenpoc231 mutant allele. The maternal/zygotic mutant heads are severely disrupted, with posterior mouth hooks (MH), median tooth (MT), hypostomal (HS), ectostomal (EcS), and epistomal sclerites (EpS), dorsal bridge (DB) and lateralgräten (LG) either missing or disrupted (see F). The denticle belts (arrows) are slightly less sclerotized than normal. (E) In this close up of the mid thoracic cuticule of a spenpoc361/Df(2L)TE21A mutant, anterior is at the top. When present in large patches as seen here, the ectopic sclerites (ES) in spen mutants consist of brown, striated, thick sclerotic cuticle that is similar in color and texture to head skeleton. This can be seen by comparing the ectopic sclerites to the normal vertical plates (VP), ventral arms (VA) and dorsal arms (DA), which can be seen underlying the ectopic sclerites. (F) Schematic diagram of spen functional requirements in head and thorax. Blue structures are defective in zygotic spen mutants, while red components are only defective in embryos lacking both the maternal and zygotic spen function.

Fig. 1.

spen is required for thorax and head development. (A) Wild-type cuticle showing normal head skeleton and thoracic segments. T2, second thoracic segment. Anterior is left and dorsal is up in this and all figures, except as noted. (B) spenpoc361/Df(2L)TE21A mutant phenotype with weak expressivity. The anterior H-piece is lost (arrowhead), and ectopic sclerites develop in the middle of T2 and T3 (arrows). (C) spenpoc361/Df(2L)TE21A mutant phenotype with strong expressivity. The ectopic sclerites cover more of the naked epidermis and are found in more segments, including T1, A1 and A2. Abdominal segments develop ectopic sclerites in lateral regions. (A)A maternal/zygotic spen mutant in which the germ line was homozygous for spenpoc361, half of these maternal mutant embryos then received the spenpoc231 mutant allele. The maternal/zygotic mutant heads are severely disrupted, with posterior mouth hooks (MH), median tooth (MT), hypostomal (HS), ectostomal (EcS), and epistomal sclerites (EpS), dorsal bridge (DB) and lateralgräten (LG) either missing or disrupted (see F). The denticle belts (arrows) are slightly less sclerotized than normal. (E) In this close up of the mid thoracic cuticule of a spenpoc361/Df(2L)TE21A mutant, anterior is at the top. When present in large patches as seen here, the ectopic sclerites (ES) in spen mutants consist of brown, striated, thick sclerotic cuticle that is similar in color and texture to head skeleton. This can be seen by comparing the ectopic sclerites to the normal vertical plates (VP), ventral arms (VA) and dorsal arms (DA), which can be seen underlying the ectopic sclerites. (F) Schematic diagram of spen functional requirements in head and thorax. Blue structures are defective in zygotic spen mutants, while red components are only defective in embryos lacking both the maternal and zygotic spen function.

Approximately 50% of the embryos that are zygotic or maternal/zygotic spen mutants also develop sclerites in the thoracic segments. These sclerites appear most frequently in the second and third thoracic segments (T2 and T3), and overlap the anterior/posterior compartment boundary. The expressivity of this phenotype ranges widely. The weakest phenotypes exhibit a row of small chunks of ectopic sclerotic material, usually limited to T2 and T3 (Fig. 1B). Strong phenotypes show broad bands of sclerotic material in all thoracic segments, often accompanied by sclerotic patches in lateral regions of the abdominal segments (Fig. 1C). No matter how extensive, the sclerites do not form in the fields of denticle belts, nor do they change the overall size of a segment. In small patches, the ectopic material is variably sclerotized, scar-like, and not recognizable as any other embryonic structure. When manifest in broad patches, the sclerotic material is brown and striated, reminiscent of the brown, striated appearance of the ventral arms, vertical plates and dorsal arms of the head skeleton (Fig. 1E). We conclude that one role of spen is to suppress the production of head-type sclerotization from ventral thoracic cells, and it is needed to a lesser extent for this function in lateral abdominal cells.

This spen homeotic phenotype is not specific to alleles isolated on the basis of Dfd interaction. In a screen for suppressors and enhancers of an activated Raf construct, Dickson et al. (1996) isolated three alleles of E(Raf)2A. These alleles do not complement the lethality of our spen alleles, and we find that the E(Raf)2A alleles also exhibit ectopic head-like sclerites, as do 11 additional alleles of spen that enhance the rough eye phenotype generated by ectopic expression of Cyclin E in the eye (Christian Lehner, personal communication).

Production of ectopic head-like sclerotic material in the trunk indicates that loss of spen function might result in de-repression of head patterning genes. We examined the cuticular phenotypes of embryos doubly mutant for spen and Distal-less (Dll), cnc, lab, Dfd, buttonhead, otd and ems, all of which have been shown to be required for determining head-specific pathways (Mohler et al., 1995; Cohen and Jurgens, 1990; Finkelstein and Perrimon, 1991). All of the double mutant combinations show the same degree of thoracic sclerotization as spen single mutants. We also examined the transcription patterns of Dll, cnc, lab, Dfd, ems, otd and apontic in spen mutant embryos and found them to be indistinguishable from wild type (data not shown). Therefore we conclude that spen does not repress ectopic head-like cuticle by repressing the expression or function of these known head-determining genes. Finally, ectopic expression of Dfd and lab in spen mutant embryos did not alter the degree or character of the ectopic sclerites in the thorax (data not shown).

Identification and sequence of spen transcripts

To understand how mutations in spen produce ectopic sclerites, we cloned the gene which encodes the spen function. All six of the EMS-induced spen mutant alleles recovered from the Dfd modifier screen (Gellon et al., 1997) fail to complement the lethality of the P-element insertion PZ1295 (l(2)03350), located within cytological interval 21B4-6. Precise excisions of the PZ1295 insertion are viable over all six spen alleles, whereas imprecise excision lines are still lethal over spen mutant chromosomes. These results suggest that PZ1295 interrupts a sequence crucial to the function of the spen locus. Genomic DNA adjacent to the site of PZ1295 insertion was rescued and overlapping λ clones spanning approximately 50 kb of surrounding genomic DNA were isolated. In addition, genomic clones covering the 3′ end of what turned out to be the spen transcription unit were kindly provided by Markus Noll. A structural map of the locus is shown in Fig. 2A. Northern analysis with genomic probes detected two transcripts from the region, one of 1.3 kb (X in Fig. 2A) and another of 20 kb (spen in Fig. 2A). The 3′ end of kismet (kis) is located 7.5 kb from the 3′ end of spen (Fig. 2A; Daubresse et al., 1999).

Five of the spen P-element insertions map either in the first exon or the first intron of the long isoform of the 20 kb transcript, while one maps immediately 5′ of this transcript (Fig. 2A). Evidence that the 20 kb transcript corresponds to the spen locus is based on: (a) reversion of the mutant phenotype upon excision of the P element inserts; (b) identification of amino acid changes in two EMS-induced spen alleles, both isolated as enhancers of CycE (Christian Lehner, personal communication; see Fig. 4); and (c) the generation of spen head defects and ectopic head-like sclerites upon injection of double stranded RNA specific to the 20 kb transcript sequence (data not shown).

Using genomic DNA fragments from these regions, approximately 200 cDNA clones were isolated from a poly-T primed, 8-12 hour embryonic cDNA library (Brown and Kafatos, 1988), and from a randomly primed, 0-18 hour embryonic cDNA library. Overlapping cDNA clones were mapped and sequenced, resulting in the exon/intron structure shown in Fig. 2A. The structure of the cDNA clones that we analyzed suggests that at least two alternative promoters initiate transcripts. A probe made from the most 5′ exons, which initiate the long (spenL) transcript isoform does not detect maternal transcripts in the 0-2 hour RNA lane of a northern blot (Fig. 2B, probe 1), whereas a probe from the common region does (Fig. 2B, probe 2). However, RT-PCR analysis (not shown) does detect spenL and spenS transcripts at all stages, including 0-2 hours, indicating that low levels of the spenL isoform exist in maternal RNA. Both probes detect transcripts of approximately 20 kb at later stages of embryonic development, although the spenL isoform is predicted to be approximately 1 kb larger based on the known cDNA structures. A third Spen protein isoform has been characterized by B. Kuang, S. Wu, Y. Shin, K. A. Mace, L. Luo and P. Kolodziej (personal communication), and is initiated from an alternative exon that maps upstream of the spenS 5′ exon. The third isoform has a few unique amino acid residues at its N terminus that differ from two Spen protein isoforms reported here.

Spen proteins contain RNP motifs and a novel conserved C-terminal domain

The first ATG in the long open reading frame of the spenLtranscript (Fig. 2) begins a long open reading frame of 5533 codons, which is predicted to make a 597 kDa protein (Fig. 3). The first ATG in the spenS transcript begins an open reading frame of 5476 codons. The SpenS protein isoform is essentially an N-terminal truncation of the SpenL protein with the addition of a Met Arg dipeptide, encoded in S exon 2, substituted for the first 59 aa residues of SpenL (Fig. 3). The N terminus of the SpenL isoform may contain unique functions due to its poly-Asn and poly-Gln sequences (Fig. 3). Three RNA-binding domains of the RNP type (amino acids 554 – 806 of L) are present in the 5474 aa common region of the predicted SpenL and SpenS proteins. Five bipartite nuclear localization sequences are scattered throughout the middle of the common sequence (positions 1872-1889, 1949-1966, 2135-2152, 2450-2467, 4556-4573 of SpenL). In addition, a region that is highly enriched in Glu (E) and Lys (K) residues is present from positions 1917-2053 of SpenL, and strings of glutamine (Q) residues are found throughout the C-terminal 1/3rd of the Spen proteins (Fig. 4A).

Fig. 3.

The predicted amino acid sequences of Spen proteins. The transcripts that include the spenL-specific exons are predicted to encode a longer protein (SpenL), the sequence of which is shown. Two putative translation start sites, α and β, are shown; α is the first initiation codon in the transcript, while β has a context closer to the consensus. The SpenS protein is predicted to contain only MetArg (MR) residues as unique N-terminal amino acids before joining the common sequence at residue 60. Protein domain predictions include three RNP domains within residues 553-806 (highlighted in black), five bipartite nuclear localization sequences (18-amino-acid blocks highlighted in black between residues 1800-4600), and a 165 residue Spen paralog and ortholog C-terminal domain, also highlighted in black, at the C terminus.

Fig. 3.

The predicted amino acid sequences of Spen proteins. The transcripts that include the spenL-specific exons are predicted to encode a longer protein (SpenL), the sequence of which is shown. Two putative translation start sites, α and β, are shown; α is the first initiation codon in the transcript, while β has a context closer to the consensus. The SpenS protein is predicted to contain only MetArg (MR) residues as unique N-terminal amino acids before joining the common sequence at residue 60. Protein domain predictions include three RNP domains within residues 553-806 (highlighted in black), five bipartite nuclear localization sequences (18-amino-acid blocks highlighted in black between residues 1800-4600), and a 165 residue Spen paralog and ortholog C-terminal domain, also highlighted in black, at the C terminus.

Database searches combined with GENESCAN and FGENE predictions of exon structure reveal apparent structural orthologs of Spen in genome sequences from human and C. elegans. These Spen family members are defined by their large size (predicted to be at least 3300 aa for human and 2738 aa for C. elegans), by their possession of three RNP domains that are closely related to those in Spen isoforms, and by a lengthy sequence match to a 165 amino acid motif at the C terminus of the D. melanogaster Spen proteins (Fig. 4A). The sequence of the first RNP domain of the Spen family (RNPa in Fig. 4B) also is closely related to the first RNP domain of Nucleolin, a protein involved in ribosome biosynthesis (Ginisty et al., 1999). The second Spen family RNP domain (RNPb in Fig. 4B) is a good match to the canonical RNP domain (Fig. 4B, bottom line), and shares even more identities with an RNP domain in AC binding factor (ACBF), an RNP-motif protein isolated from tobacco. ACBF binds to DNA regulatory sequences of genes in the phenylpropanoid biosynthetic pathways of many plants (Séguin et al., 1997).

The C-terminal motif, which we call the SPOC domain (Spen Paralog and Ortholog C-terminal domain) is very similar among family members (57% identity between D. melanogaster Spen and H. sapiens Spen1; Fig. 4C), but matches no other sequence motifs with known biochemical function. The entire open reading frames of two EMS-induced spen alleles (spenE(CycE++)D57 and spenE(CycE++)e9) were sequenced. Single base substitutions in each of these alter the predicted amino acid sequence at highly conserved positions within the SPOC domain: codon position 5444 of spenE(CycE++)e9, TGC – Cys to TAC – Tyr; and codon position 5480 of spenE(CycE++)D57, GGC – Gly to GAC – Asp (Fig. 4C). With regard to the embryonic head and thoracic defects observed in spen361 zygotic mutants, both of these alleles are hypomorphs. The order of allele strengths with respect to zygotic defects is spen361 > spen231 > spenE(CycE++)e9 > spenE(CycE++)D57.

Our sequence similarity searches also reveal a separate sub-family of short Spen-like proteins encoded in D. melanogaster and other animal genomes. The short Spen-like proteins have a similar arrangement of Spen-like RNP motifs and SPOC domains to those in the Spen orthologs, but the short Spen-like proteins are approximately one tenth the size of Spen (Fig. 4). All of the known short Spen-like proteins also encode an RGG motif, which is capable of destablizing RNA helices and often found in combination with RNP motifs (Siomi and Dreyfuss, 1997). The current sequence evidence suggests that a duplication event giving rise to the genes that encode the large and small Spen-like proteins occurred before the divergence of the deuterostome and protostome lineages in metazoan evolution (Aguinaldo et al., 1997).

Spen proteins are ubiquitously expressed and localized in nuclei

Antibodies were raised against a common fragment of the Spen proteins (amino acids 3203-3714) to assay the expression pattern and subcellular localization. Immunostaining of whole-mount embryos shows that Spen antigen is expressed in most or all cell types, and is concentrated in nuclei (Fig. 5A-C). Spen protein staining is first detectable prior to cellularization in stage 3 embryonic nuclei, and is present in all blastoderm cells, including the pole cells (Fig. 5A). The protein is expressed throughout the rest of embryogenesis, and is concentrated in nuclei. At later embryonic stages (9 through 14), nuclear Spen staining appears to be most abundant in CNS and epidermal cells (Fig. 5C), but most tissues exhibit nuclear staining at detectable levels.

Fig. 5.

Spen protein antigen is concentrated in nuclei of most embryonic cells. (A) Stage 5 embryo showing anti-Spen staining in all nuclei. The antibodies used were raised against the protein fragment 3203-3714, which is common to both SpenL and SpenS proteins. (B) Close-up of embryo in A. (C) Dorsolateral cells of the second and third thoracic segments in a stage 13 embryo; Spen protein staining is concentrated in nuclei.

Fig. 5.

Spen protein antigen is concentrated in nuclei of most embryonic cells. (A) Stage 5 embryo showing anti-Spen staining in all nuclei. The antibodies used were raised against the protein fragment 3203-3714, which is common to both SpenL and SpenS proteins. (B) Close-up of embryo in A. (C) Dorsolateral cells of the second and third thoracic segments in a stage 13 embryo; Spen protein staining is concentrated in nuclei.

spen cooperates with Antp and tsh to repress head-like sclerites in the thorax

Null mutations in Antp result in a transformation of T2 and T3 towards T1 in the embryonic body plan (Wakimoto and Kaufman, 1981; Martinez-Arias, 1986). In addition, Antp mutant embryos develop ectopic head-like sclerites in the dorsal thorax (between T1 and T2; Fig. 6A), similar in kind but not in position to the ectopic sclerite phenotype seen in spen mutants. To test whether spen and Antp function in an additive or synergistic manner in the repression of head-like sclerites in the thorax, we first studied spen; Antp cuticle phenotypes. Embryos mutant for both spen and Antp have more sclerotic material in dorsal T2 than do Antp mutants alone (Fig. 6B). In addition, the ectopic head-like sclerites in the ventral thorax of spen; Antp mutants are more sclerotized and extensive than in spen mutants alone (Fig. 6D, compare to 6E). The sclerotic material in spen; Antp mutants frequently appears in two distinct bands, one in the center of the segment similar to the position in spen mutants, and at another position in the posterior of T1 and T2 (Fig. 6D, arrow). These posterior ectopic sclerites do not develop in T3 (Fig. 6D). The enhanced formation of head-like sclerites in spen; Antp mutants suggests that spen and Antp function in a common or interacting pathway(s) in subregions of T1 and T2.

Fig. 6.

Antp and spen cooperatively repress head-like sclerites in thoracic segments. (A) Dorsal head of an Antp embryo showing ectopic sclerotization (arrowhead). B) Dorsal head of a spen; Antp embryo, a sibling of the embryo in A. The amount of dorsal ectopic sclerotization is increased (black arrowhead). (C-H) Ventral T2 and T3 embryonic cuticle. (C) Antp control embryo has sclerites in dorsal thoracic cuticle. There is no sclerotization in ventral regions of the thorax. (D) spen; Antp embryos have more extensive sclerotization in the thorax than spen siblings, particularly in the posterior compartment (arrow). (E) spen control sibling of the embryo in D. (F) hsAntp control embryo treated with heat shock. (G)One of the rare spen; hsAntp siblings of the embryo in F that shows bits of thoracic sclerotization. (H) spen ; hsAntp control embryo that was not treated with heat shock.

Fig. 6.

Antp and spen cooperatively repress head-like sclerites in thoracic segments. (A) Dorsal head of an Antp embryo showing ectopic sclerotization (arrowhead). B) Dorsal head of a spen; Antp embryo, a sibling of the embryo in A. The amount of dorsal ectopic sclerotization is increased (black arrowhead). (C-H) Ventral T2 and T3 embryonic cuticle. (C) Antp control embryo has sclerites in dorsal thoracic cuticle. There is no sclerotization in ventral regions of the thorax. (D) spen; Antp embryos have more extensive sclerotization in the thorax than spen siblings, particularly in the posterior compartment (arrow). (E) spen control sibling of the embryo in D. (F) hsAntp control embryo treated with heat shock. (G)One of the rare spen; hsAntp siblings of the embryo in F that shows bits of thoracic sclerotization. (H) spen ; hsAntp control embryo that was not treated with heat shock.

The synergistic effect of Antp and spen might be due to a regulatory effect of Antp on spen transcription pattern, or to Spen effects on Antp transcript pattern or translation. However, Antp transcript and protein expression patterns are unchanged in spen mutant embryos, and spen transcript expression is unchanged in Antp mutant embryos (data not shown). Therefore, spen and Antp appear to be acting in parallel, presumably due to direct or indirect regulation of common downstream genes.

If spen and Antp regulate common targets, then induction of high levels of exogenous Antp expression might result in suppression of the spen mutant phentoype. We therefore tested the ability of excess Antp protein to suppress the spen mutant phenotype. Overexpression of Antp under heat shock promoter control (hsAntp) causes a transformation of head regions to thoracic identity, but leaves T2 and T3 nearly unchanged (Fig. 6F; Gibson and Gehring, 1988). When Antp is overexpressed in a spen mutant background, the ectopic head-like sclerites are strongly suppressed (Fig. 6G). The number of hsAntp, spen embryos which exhibit any detectable ectopic sclerites was less than half the expected number compared to spen mutant siblings from the same cross, or compared to spen, hsAntp embryos that were not subjected to heat shock. In addition, the sclerites which do occasionally appear in heat shocked hsAntp; spen embryos are smaller than those in their spen siblings (Fig. 6G,H) The ability of excess Antp to suppress the spen homeotic transformation indicates that the two genes interact to repress ectopic head-like sclerites.

In the head region, where spen is required for the development of sclerites that also require Dfd, Scr and other head genes, it is also possible that spen might work in parallel to Hox pathways. This is the case for Dfd and Scr, since mutations in these genes had no effect on spen transcript expression pattern, and conversely, spen mutants had no effect on Dfd or Scr transcript or protein expression patterns (data not shown). We attempted to test whether the overexpression of Dfd could rescue the H-piece defect in spen mutant embryos, but the morphology of the head skeleton was so disrupted by heat shock induced ectopic Dfd protein (Kuziora and McGinnis, 1988) that it was not possible to conclude whether the anterior H-piece was restored.

Another Drosophila gene involved in distinguishing head from body is teashirt (tsh). Tsh protein is expressed only in the labial segment and trunk region of embryos, where it is required to repress head identity and to promote thoracic and abdominal segment identities (Fasano et al., 1991). tsh transcription levels in the thorax are maintained by Antp (Röder et al., 1992; McCormick et al., 1995), but a variety of genetic interaction tests have shown that Antp and Tsh have independent functions in repressing head development (Röder et al., 1992). Is spen integrated into the Antp; tsh pathways by regulation of the tsh or Antp expression patterns. Our experiments show that: (1) expression pattern of Tsh protein (Röder et al., 1992) is unchanged in spen mutant embryos, (2) protein expression patterns of Scr or Antp in spen, tsh double mutant embryos are unchanged from the pattern seen in tsh mutants alone (data not shown), and (3) the spen transcript pattern is normal in tsh mutants (data not shown). Therefore, Spen suppression of head-like sclerites is not exerted by a regulatory effect on the Tsh protein expression pattern, nor by Tsh effects on the spen transcript pattern, nor through combinatorial effects of spen and tsh on Scr or Antp protein abundance.

The phenotype of tsh, spen mutant embryos suggests that the two genes act to promote thoracic development. In tsh mutant embryos, the T1 denticle belt is absent and although the remaining denticle belts appear to have the appropriate segmental identities, the denticles themselves are disorganized and smaller than in wild type (Fig. 7A; Fasano et al., 1991; Röder et al., 1992). In contrast, tsh, spen double mutants completely lack denticle belts in the thorax (Fig. 7B). This may be due to the the death of cells in the denticle field in the thorax of the double mutants, or to the inability of Antp protein, still expressed in the remaining cells, to promote the development of thorax-specific structures.

Fig. 7.

teashirt (tsh) and spen cooperate in normal thoracic development. (A) Ventral cuticle of a tsh embryo. T1 denticles are absent. T2, T3 and first abdominal (A1) denticle belts are present (arrowheads), but have small denticles in disorganized patterns. The segments are also much narrower than wild type. (B) A comparable region from a spen; tsh embryo. Some ectopic sclerotic material can be seen in the thoracic region, but the denticle belts are completely absent, which is not seen in tsh alone. The A1 denticle belt is still barely identifiable (arrowhead). (C) T2 and T3 of a control spen sibling of the embryo shown in (D). D) T2 and T3 segments of a spen; UAStsh/69B embryo, in which tsh was ubiquitously over-expressed under control of a 69B-GAL4 driver. This completely represses the ectopic sclerites observed in spen mutant embryos.

Fig. 7.

teashirt (tsh) and spen cooperate in normal thoracic development. (A) Ventral cuticle of a tsh embryo. T1 denticles are absent. T2, T3 and first abdominal (A1) denticle belts are present (arrowheads), but have small denticles in disorganized patterns. The segments are also much narrower than wild type. (B) A comparable region from a spen; tsh embryo. Some ectopic sclerotic material can be seen in the thoracic region, but the denticle belts are completely absent, which is not seen in tsh alone. The A1 denticle belt is still barely identifiable (arrowhead). (C) T2 and T3 of a control spen sibling of the embryo shown in (D). D) T2 and T3 segments of a spen; UAStsh/69B embryo, in which tsh was ubiquitously over-expressed under control of a 69B-GAL4 driver. This completely represses the ectopic sclerites observed in spen mutant embryos.

As to whether tsh and spen collaborate in repressing head-like sclerites, we find that the tsh, spen double mutants still have bits of sclerite in the ‘thorax’ of the double mutants, so this phenotype is not enhanced. However,we also examined the effects of Tsh overexpression on the ectopic head-like sclerites in spen mutants. In wild-type embryos, overexpression of Tsh protein throughout the embryo results in transformation of head regions toward thoracic identity (de Zulueta et al., 1994), as well as poorly differentiated denticle belts, especially in the thorax. In the thorax of spen mutant embryos that also overexpress Tsh, the ectopic ventral head-like sclerites are strongly suppressed (Fig. 7D). Taken together, these results suggest that spen, tsh and Antp function in a combinatorial manner to repress the development of head-like sclerites and promote the development of thoracic identity.

Spen function in Hox, Tsh, and Raf pathways

Encoding RNA binding domains, spen is a unique type of homeotic function. It is neither in the Hox class, nor is it required for the proper spatial regulation of Hox transcription, such as trx or Pc group homeotics. It falls into the growing class of functions that act in parallel to modulate the output of Hox pathways. The direction of spen’s modulatory effect, either helping to promote or to repress head-type sclerotization, is apparently dictated by the type of Hox gene expressed in a certain region of the body.

Although expressed throughout the embryo, spen appears to be especially important to the morphogenesis of head cuticular structures, presumably due to modulation of its function on the anterior/posterior axis by specific Hox proteins and specific signaling pathways. The epistasis experiments reported here, along with previous results, indicate that in its role of promoting sclerite formation in the head, spen acts in parallel to head patterning genes such as Dfd and ems (Fig. 8). There are also other embryonic head structures that are missing or abnormal in both maternal/zygotic spen mutants and other head homeotic mutants such as lab and Scr (Merrill et al., 1989; Sato et al., 1985; Pattatucci et al., 1991). The homeotic function of spen is exerted in the trunk, where it represses head-like sclerite formation in combination with both the segment identity homeotic Antp and the regional homeotic gene tsh (Fig. 8).

Fig. 8.

A summary of Spen functions in the head and thorax. In the head, spen is required for production of specific head sclerites, which are also dependent on Hox gene function. Both Dfd and spen functions are required for the production of the base of the mouth hooks, anterior portion of the H-piece, lateralgräten and ectostomal sclerite (all shown in red). Thus, together spen and Dfd regulate the formation or sclerotization of head structures, either in a common pathway (1), or independently (2). In regions of ventral thoracic cuticle (region indicated in blue hatching), spen cooperates with Antp and tsh to repress head-like sclerites and promote normal thoracic development.

Fig. 8.

A summary of Spen functions in the head and thorax. In the head, spen is required for production of specific head sclerites, which are also dependent on Hox gene function. Both Dfd and spen functions are required for the production of the base of the mouth hooks, anterior portion of the H-piece, lateralgräten and ectostomal sclerite (all shown in red). Thus, together spen and Dfd regulate the formation or sclerotization of head structures, either in a common pathway (1), or independently (2). In regions of ventral thoracic cuticle (region indicated in blue hatching), spen cooperates with Antp and tsh to repress head-like sclerites and promote normal thoracic development.

The thorax-promoting spen function also genetically interacts with both tsh and Antp. In double mutants for spen and tsh, the thorax has few segmental specializations. This is an interesting result, for Antp expression is unaffected in the epidermal cells that remain in the thoracic region. Another result indicating a close genetic interaction between tsh and spen is that overexpression of Tsh protein can suppress the ectopic sclerite phenotype of spen mutants. The Spen and Tsh functions, both of which act in parallel to the Hox proteins, seem likely to exert this effect through by their collaboration (direct or indirect) on the regulation of Hox downstream target genes. Presumably this occurs by the Tsh DNA binding protein operating as a transcription factor (Alexandre et al., 1996), and by Spen acting at the level of RNA processing/transport.

Staehling-Hampton et al. (1999) recently discovered that in Drosophila eye cells, wild-type spen genetically antagonizes ectopically expressed E2F and Dp proteins, and genetically assists p21 in blocking S-phase entry. Thus at least in the eye, spen apparently has a role in suppressing progression of the cell cycle. Other evidence consistent with this derives from the study of interactions between spen and Raf mutations. spenE(Raf)2A alleles were found to enhance a dominant, constitutively active Raf phenotype in the R7 photoreceptors of the Drosophila eye (Dickson et al., 1996). Clones of spen mutant cells in the eye paradoxically exhibited either loss of R7 cells, or supernumerary R7 and other photoreceptors (Dickson et al., 1996). Given the interactions and phenotypes they observed, Dickson et al. (1996) concluded that at least in R7 cells, spen acts downstream of Raf, as a constitutive repressor of Raf signaling, in a manner similar to the ETS transcription factor Aop/Yan (Nüsslein-Volhard et al., 1984; Lai and Rubin, 1992). It is interesting to note that the maternal/zygotic loss of function mutations reported here for spen, and those for null zygotic mutations of aop/yan (Rogge et al., 1995) are similar in that thoracic and abdominal features are relatively normal, while head skeletal structures are missing or malformed. So it is possible that Spen may act along with Aop/Yan in the nuclei of embryonic head cells as a repressor of Ras/Raf signaling.

Genetic interactions between Ras signaling and Hox function have been described previously; Boube et al. (1997) found that proboscipedia (pb) loss-of-function homeotic phenotypes in the Drosophila labium were enhanced in Ras1/Ras1+ adults, and Ultrabithorax (Ubx) loss-of-function homeotic phenotypes were enhanced when haltere cells contained only one wild-type allele of Gap1 (a Ras inhibitor). In C. elegans, the ability of the lin-39 Hox gene to specify vulval cell fates requires activation of the Ras pathway (Clandinin et al., 1997; Maloof and Kenyon, 1998; Eisenmann et al., 1998). Ras acts both in parallel to lin-39 as a co-activator of vulval identity, and also amplifies the amount of lin-39 protein expression in vulval precursor cells. Another C. elegans Hox gene, egl-5, also requires Ras signaling to provide a neuroectoblast cell fate to the P12 cell, which is accomplished at least in part by Ras pathway activation of the transcription of egl-5 (Jiang and Sternberg, 1998). Thus the impact of Ras/Raf signaling on Hox gene expression and Hox protein function may be commonly used in developmental genetic circuitry. In light of all the genetic interactions of Spen, it is possible that it has a general function in inhibiting cell division and acts in combination with Hox, E2F, and Ras/Raf signaling pathways to accomplish this role. How this putative general growth control function might be connected (if at all) to the loss of sclerites in the head and the ectopic production of head-like sclerites in the trunk is unknown.

The extent to which Spen connects Ras/Raf pathways and Hox pathways is a fascinating but still largely unexplored area, as are the mechanisms by which Spen fits into these pathways. RNA-binding proteins have been postulated to function in signaling pathways by binding signal-responsive gene transcripts (Siomi and Dreyfuss, 1997). In this postulated mechanism, a transcript might be released for translation in response to activation of a signaling pathway, providing very rapid protein induction; or a different protein isoform might be generated by differential splicing in response to signaling pathways. However, these and other possible mechanisms are speculative until common target genes of Hox and Spen, or Spen and Raf, are identified.

Downstream targets for Hox and Spen

The ectopic head-like sclerites which develop in the thorax of spen mutant embryos are darker, more striated, and thicker than normal thoracic body wall cuticle. In addition, spen is required for the formation and sclerotization of head structures such as the H-piece and ventral arms. Thus a common theme of spen function in the epidermis is regulation of sclerotization. What are possible ‘sclerite’ target genes that might be activated by Dfd and Spen, and repressed by Antp and Spen?

At the molecular level, the formation of sclerites involves the incorporation of a complex array of substances in addition to chitin. Among these substances are cross-linking agents that generate darkly pigmented, thick, and hard sclerites (Poodry, 1980). Therefore, in thoracic cells, Spen and Antp are directly or indirectly repressing the expression or activities of the enzyme and transport pathways that synthesize and secrete these sclerotization-promoting substances. Genes in the pathways that synthesize some of these substances have been long studied in Drosophila, and include Dopa decarboxylase (Ddc) and pale (tyrosine hydroxylase; Wright, 1996; Jurgens et al., 1984; Neckameyer and White, 1993). These enzymes are required for the production of dopamine, which functions both as a neurotransmitter in the fly nervous system, and a precursor of cross-linking molecules required for the sclerotization of cuticle in the body wall. It will be interesting to investigate whether Ddc, pale, or other genes in the dopamine pathway are targets of the repressive functions of spen and Antp in thoracic cells.

Spen orthologs and Spen-like proteins

Spen and its homologs contain three clustered RNP-type RNA binding domains (Siomi and Dreyfuss, 1997; Handa et al., 1999). Other proteins which contain RNP domains include Sex lethal (Sxl), U1A snRNP, Nucleolin, and the circadian rhythm regulator Lark (Burd and Dreyfuss, 1994; Serin et al., 1997; Newby and Jackson, 1996). It therefore seems highly likely that Spen family protein functions will be exerted by RNA binding.

We thank T. Kaufman, J. Mahaffey, G.Morata, G. Struhl, S. Kerridge, Christian Lehner, Barry Dickson and Ernst Hafen for providing fly stocks, J. Tamkun and N. Brown for sharing their Drosophila libraries. Christian Lehner and Peter Kolodziej generously shared unpublished results. We are also grateful to M. Noll for providing genomic clones in the spen region, Eileen Furlong and Matt Scott for anti-Antp and anti-tsh antisera, and J. Tamkun for sharing the northern analysis of the kis region prior to publication. Both FlyBase and the Berkeley Drosophila Genome Project were of great value at many stages in the completion of these experiments. This work was supported by NIH grant HD 28315.

The isolation of additional spen mutant alleles, along with the sequence of one of the Spen protein isoforms is also reported by Rebay, I., Chen, F., Hsaio, F., Kolodziej, P. A., Kuang, B. H., Laverty, T., Suh, C., Voas, M., Williams, A. and Rubin, G. M. (1999). A genetic screen for novel components of the Ras/MAPK signaling pathway that interact with the yan gene of Drosophila identifies split ends, a new RRM motif containing protein. Genetics, in press.

Aguinaldo
,
A.
,
Turbeville
,
J.
,
Linford
,
L.
,
Rivera
,
M.
,
Garey
,
J.
,
Raff
,
R.
and
Lake
,
J.
(
1997
).
Evidence for a clade of nematodes, arthropods and other moulting animals
.
Nature
387
,
489
493
.
Alexandre
,
E.
,
Graba
,
Y.
,
Fasano
,
L.
,
Gallet
,
A.
,
Perrin
,
L.
,
De Zulueta
,
P.
,
Pradel
,
J.
,
Kerridge
,
S.
and
Jacq
,
B.
(
1996
).
The Drosophila teashirt homeotic protein is a DNA-binding protein and modulo, a HOM-C regulated modifier of variegation, is a likely candidate for being a direct target gene
.
Mech. Dev
.
59
,
191
204
.
Altschul
,
S. F.
,
Madden
,
T. L.
,
Schäffer
,
A. A.
,
Zhang
,
J.
,
Zhang
,
Z.
,
Miller
,
W.
and
Lipman
,
D. J.
(
1997
).
Gapped BLAST and PSI-BLAST: a new generation of protein database search programs
.
Nucl. Acids Res
.
25
,
3389
3402
.
Biggin
,
M. D.
and
McGinnis
,
W.
(
1997
).
Regulation of segmentation and segmental identity by Drosophila homeodomain proteins: the role of DNA binding in functional activity and specificity
.
Development
124
,
4425
4433
.
Boube
,
M.
,
Benassayag
,
C.
,
Seroude
,
L.
and
Cribbs
,
D. L.
(
1997
).
Ras1-mediated modulation of Drosophila homeotic function in cell and segment identity
.
Genetics
146
,
619
628
.
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
.
Brown
,
N. H.
and
Kafatos
,
F. C.
(
1988
).
Functional cDNA libraries from Drosophila embryos
.
Mol. Biol
.
203
,
425
437
.
Burd
,
C. G.
and
Dreyfuss
,
G.
(
1994
).
Conserved structures and diversity of RNA-binding proteins
.
Science
265
,
615
621
.
Campos-Ortega
,
J. A.
and
Hartenstein
,
V.
(
1985
).
The Embryonic Development of Drosophila melanogaster
.
Springer-Verlag
,
Berlin
.
Capovilla
,
M.
and
Botas
,
J.
(
1998
).
Functional dominance among Hox genes: repression dominates activation in the regulation of Dpp
.
Development
125
,
4949
4957
.
Chan
,
S. K.
and
Mann
,
R. S.
(
1996
).
A structural model for a HOX-extradenticle-DNA complex accounts for the choice of the HOX protein in the heterodimer
.
Proc. Natl. Acad. Sci. USA
93
,
5223
5228
.
Chan
,
S. K.
,
Popperl
,
H.
,
Krumlauf
,
R.
and
Mann
,
R. S.
(
1996
).
An extradenticle-induced conformational change in a HOX protein overcomes an inhibitory function of the conserved hexapeptide motif
.
EMBO J
.
15
,
2476
2487
.
Chan
,
S. K.
,
Ryoo
,
H. D.
,
Gould
,
A.
,
Krumlauf
,
R.
and
Mann
,
R. S.
(
1997
).
Switching the in vivo specificity of a minimal Hox-responsive element
.
Development
124
,
2007
2014
.
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
.
Clandinin
,
T. R.
,
Katz
,
W. S.
and
Sternberg
,
P. W.
(
1997
).
Caenorhabditis elegans HOM-C genes regulate the response of vulval precursor cells to inductive signal. Dev Biol
.
182
,
150
161
.
Cohen
,
S. M.
and
Jurgens
,
G.
(
1990
).
Mediation of Drosophila head development by gap-like segmentation genes
.
Nature
346
,
482
485
.
Dalton
,
D.
,
Chadwick
,
R.
and
McGinnis
,
W.
(
1989
).
Expression and embryonic function of empty spiracles: a Drosophila homeo box gene with two patterning functions on the anterior-posterior axis of the embryo
.
Genes Dev
.
3
,
1940
1956
.
Daubresse
,
G.
,
Deuring
,
R.
,
Moore
,
L.
,
Papoulas
,
O.
,
Zakrajsek
,
I.
,
Waldrip
,
W. R.
,
Scott
,
M. P.
,
Kennison
,
J. A.
and
Tamkun
,
J. W.
(
1999
).
The Drosophila kismet gene is related to chromatin-remodeling factors and is required for both segmentation and segment identity
.
Development
126
,
1175
1187
.
de Zulueta
,
P.
,
Alexandre
,
E.
,
Jacq
,
B.
and
Kerridge
,
S.
(
1994
).
Homeotic complex and teashirt genes co-operate to establish trunk segmental identities in Drosophila
.
Development
120
,
2287
2296
.
Dickson
,
B. J.
,
Van Der Straten
,
A.
,
Dominguez
,
M.
and
Hafen
,
E.
(
1996
).
Mutations modulating Raf signaling in Drosophila eye development
.
Genetics
142
,
163
171
.
Eisenmann
,
D. M.
,
Maloof
,
J. N.
,
Simske
,
J. S.
,
Kenyon
,
C.
and
Kim
,
S. K.
(
1998
).
The beta-catenin homolog BAR-1 and LET-60 Ras coordinately regulate the Hox gene lin-39 during Caenorhabditis elegans vulval development
.
Development
125
,
3667
3680
.
Fasano
,
L.
,
Röder
,
L.
,
Core
,
N.
,
Alexandre
,
E.
,
Vola
,
C.
,
Jacq
,
B.
and
Kerridge
,
S.
(
1991
).
The gene teashirt is required for the development of Drosophila embryonic trunk segments and encodes a protein with widely spaced zinc finger motifs
.
Cell
64
,
63
79
.
Finkelstein
,
R.
and
Perrimon
,
N.
(
1991
).
The molecular genetics of head development in Drosophila
.
Development
112
,
899
912
.
Flegel
,
W. A.
,
Singson
,
A. W.
,
Margolis
,
J. S.
,
Bang
,
A. G.
,
Posakony
,
J. W.
and
Murre
,
C.
(
1993
).
Dpbx, a new homeobox gene closely related to the human proto-oncogene pbx1: molecular structure and developmental expression. Mech. Dev
.
41
,
155
161
.
Florence
,
B.
and
McGinnis
,
W.
(
1998
).
A genetic screen of the Drosophila X chromosome for mutations that modify Deformed function
.
Genetics
150
,
1497
1511
.
FlyBase
. (
1999
).
The FlyBase database of the Drosophila Genome Projects and community literature. The FlyBase Consortium
.
Nucleic Acids Res
.
27
,
85
88
.
Garcia-Bellido
,
A.
(
1977
).
Homeotic and atavic mutations in insects
.
Amer. Zool
.
17
,
613
629
.
Gellon
,
G.
,
Harding
,
K.
,
McGinnis
,
N.
,
Martin
,
M. M.
and
McGinnis
,
W.
(
1997
).
A genetic screen for modifiers of Deformed homeotic function identifies novel genes required for head development
.
Development
124
,
3321
3331
.
Gibson
,
G.
and
Gehring
,
W. J.
(
1988
).
Head and thoracic transformation caused by ectopic expression of Antennapedia during Drosophila development
.
Development
102
,
657
675
.
Ginisty
,
H.
,
Sicard
,
H.
,
Roger
,
B.
and
Bouvet
,
P.
(
1999
).
Structure and functions of nucleolin
.
J. Cell Sci
.
112
,
761
772
.
Gonzalez-Reyes
,
A.
,
Urquia
,
N.
,
Gehring
,
W. J.
,
Struhl
,
G.
and
Morata
,
G.
(
1990
).
Are cross-regulatory interactions between homeotic genes functionally significant
.
Nature
344
,
78
80
.
Gray
,
S.
,
Szymanski
,
P.
and
Levine
,
M.
(
1994
)
Short-range repression permits multiple enhancers to function autonomously within a complex promoter
.
Genes Dev
.
8
,
1829
1838
.
Hafen
,
E.
,
Levine
,
M.
,
Garber
,
R. L.
and
Gehring
,
W. J.
(
1983
).
An improved in situ hybridization method for the detection of cellular RNAs in Drosophila tissue sections and its application for localizing transcripts of the homeotic Antennapedia gene complex
.
EMBO J
.
2
,
617
623
.
Handa
,
N.
,
Nureki
,
O.
,
Kurimoto
,
K.
,
Kim
,
I.
,
Sakamoto
,
H.
,
Shimura
,
Y.
,
Muto
,
Y.
and
Yokoyama
,
S.
(
1999
).
Structural basis for recognition of the tra mRNA precursor by the Sex-lethal protein
.
Nature
398
,
579
585
.
Harding
,
K. W.
,
Gellon
,
G.
,
McGinnis
,
N.
and
McGinnis
,
W.
(
1995
).
A screen for Dfd modifier mutations in Drosophila
.
Genetics
140
,
1339
1352
.
Jiang
,
L. I.
and
Sternberg
,
P. W.
(
1998
).
Interactions of EGF, Wnt and HOM-C genes specify the P12 neuroectoblast fate in C. elegans
.
Development
125
,
2337
2347
.
Jurgens
,
G.
,
Wieschaus
,
E.
,
Nüsslein-Volhard
,
C.
and
Kluding
,
H.
(
1984
).
Mutations affecting the pattern of the larval cuticle in Drosophila melanogaster. II. Zygotic loci on the third chromosome
.
Roux’s Arch. Dev. Biol
.
193
,
283
295
.
Kennerdell
,
J. R.
and
Carthew
,
R. W.
(
1998
).
Use of dsRNA-mediated genetic interference to demonstrate that frizzled and frizzled 2 act in the wingless pathway
.
Cell
95
,
1017
1026
.
Kolodziej
,
P. A.
,
Jan
,
L. Y.
and
Jan
,
Y. N.
(
1995
).
Mutations that affect the length, fasciculation, or ventral orientation of specific sensory axons in the Drosophila embryo
.
Neuron
15
,
273
286
.
Kurant
,
E.
,
Pai
,
C.-y.
,
Sharf
,
R.
,
Halachmi
,
N.
,
Sun
,
Y. H.
and
Salzberg
,
A.
(
1998
).
dorsotonals/homothorax, the Drosophila homologue of meis1, interacts with extradenticle in patterning of the embryonic PNS
.
Development
125
,
1037
1048
.
Kuroiwa
,
A.
,
Kloter
,
U.
,
Baumgartner
,
P.
and
Gehring
,
W. J.
(
1985
).
Cloning of the homeotic Sex combs reduced in Drosophila and in situ localization of its transcripts
.
EMBO J
.
4
,
3757
3764
.
Kuziora
,
M. A.
and
McGinnis
,
W.
(
1988
).
Autoregulation of a Drosophila homeotic selector gene
.
Cell
55
,
477
485
.
Lai
,
Z. C.
and
Rubin
,
G. M.
(
1992
).
Negative control of the photoreceptor development in Drosophila by the product of the yan gene, an ETS domain protein
.
Cell
70
,
609
620
.
Lewis
,
E. B.
(
1978
).
A gene complex controlling segmentation in Drosophila
.
Nature
276
,
565
570
.
Li
,
X.
and
McGinnis
,
W.
(
1999
).
Activity regulation of Hox proteins, a mechanism for altering functional specificity in development and evolution
.
Proc. Natl. Acad. Sci. USA
96
,
6802
6907
.
Li
,
X.
,
Murre
,
C.
and
McGinnis
,
W.
(
1999
).
Activity regulation of a Hox protein and a role for the homeodomain in inhibiting transcriptional activation
.
EMBO J
.
18
,
198
211
.
Lindsley
,
D. L.
and
Zimm
,
G. G.
(
1992
).
The Genome of Drosophila melanogaster
.
Academic Press, Inc., San Diego
.
Maloof
,
J. N.
and
Kenyon
,
C.
(
1998
).
The Hox gene lin-39 is required during C. elegans vulval induction to select the outcome of Ras signaling
.
Development
125
,
181
190
.
Mann
,
R. S.
and
Chan
,
S. K.
(
1996
).
Extra specificity from extradenticle: the partnership between HOX and PBX/EXD homeodomain proteins
.
Trends Genet
.
12
,
259
262
.
Martinez-Arias
,
A.
(
1986
).
The Antennapedia gene is required and expressed in parasegments 4 and 5 of the Drosophila embryo
.
EMBO J
.
5
,
135
141
.
McCormick
,
A.
,
Core
,
N.
,
Kerridge
,
S.
and
Scott
,
M.
(
1995
).
Homeotic response elements are tightly linked to tissue-specific elements in a transcriptional enhancer of the teashirt gene
.
Development
121
,
2799
2812
.
McGinnis
,
N.
,
Ragnhildstveit
,
E.
,
Veraksa
,
A.
and
McGinnis
,
W.
(
1998
).
A cap ‘n’ collar protein isoform contains a selective Hox repressor function
.
Development
125
,
4553
4564
.
McGinnis
,
W.
,
Jack
,
T.
,
Chadwick
,
R.
,
Regulski
,
M.
,
Bergson
,
C.
,
McGinnis
,
N.
and
Kuziora
,
M. A.
(
1990
).
Establishment and maintenance of position-specific expression of the Drosophila homeotic selector gene Deformed
.
Adv. Genet
.
27
,
363
402
.
McGinnis
,
W.
and
Krumlauf
,
R.
(
1992
).
Homeobox genes and axial patterning
.
Cell
68
,
283
302
.
Merrill
,
V. K. L.
,
Diederich
,
R. J.
,
Turner
,
F. R.
and
Kaufman
,
T. C.
(
1989
).
A genetic and developmental analysis of mutations in labial, a gene necessary for proper head formation in Drosophila melanogaster
.
Dev. Biol
.
135
,
376
391
.
Merrill
,
V. K. L.
,
Turner
,
F. R.
and
Kaufman
,
T. C.
(
1987
).
A genetic and developmental analysis of mutations in the Deformed locus in Drosophila melanogaster
.
Dev. Biol
.
122
,
379
395
.
Mohler
,
J.
,
J. W.
Mahaffey
,
E.
Deutsch
, and
Vani
,
K.
(
1995
).
Control of Drosophila head segment identity by the bZIP homeotic gene cnc
.
Development
121
,
237
247
.
Myers
,
R. M.
,
T.
Maniatis
, and
Lerman
,
L. S.
(
1987
).
Detection and localization of single base changes by denaturing gradient gel electrophoresis
.
Methods Enzymol
.
155
,
501
527
.
Neckameyer
,
W. S.
and
White
,
K.
(
1993
).
Drosophila tyrosine hydroxylase is encoded by the pale locus
.
J. Neurogenet
.
8
,
189
199
.
Newby
,
L. M.
and
Jackson
,
F. R.
(
1996
).
Regulation of a specific circadian clock output pathway by lark, a putative RNA-binding protein with repressor activity
.
J. Neurobiol
.
31
,
117
128
.
Nüsslein-Volhard
,
C.
,
Wieschaus
,
E.
and
Kluding
,
H.
(
1984
).
Mutations affecting the pattern of the larval cuticle in Drosophila melanogaster. I. Zygotic loci on the second chromosome
.
Roux’s Arch. Dev. Biol
.
193
,
267
282
.
Pai
,
C.-Y.
,
Kuo
,
T.-S.
,
Jaw
,
T. J.
,
Kurant
,
E.
,
Chen
,
C.-T.
,
Bessarab
,
D.
,
Salzberg
,
A.
and
Sun
,
Y. H.
(
1998
).
The Homothorax homeoprotein activates the nuclear localization of another homeoprotein, Extradenticle, and suppresses eye development in Drosophila
.
Genes Dev
.
12
,
435
446
.
Pattatucci
,
A. M.
,
Otteson
,
D. C.
and
Kaufman
,
T. C.
(
1991
).
A functional and structural analysis of the Sex-combs reduced locus of Drosophila melanogaster
.
Genetics
129
,
423
441
.
Peifer
,
M.
and
Wieschaus
,
E.
(
1990
).
Mutations in the Drosophila gene extradenticle affect the way specific homeo domain proteins regulate segmental identity
.
Genes Dev
.
4
,
1209
1223
.
Pinsonneault
,
J.
,
Florence
,
B.
,
Vaessin
,
H.
and
McGinnis
,
W.
(
1997
).
A model for extradenticle function as a switch that changes Hox proteins from repressors to activators
.
EMBO J
.
16
,
2032
2042
.
Poodry
,
C. A.
(
1980
).
Epidermis, morphology and development
. In
The Genetics and Biology of Drosophila
(ed.
M.
Ashburner
and
T. R. F.
Wright
), pp.
443
497
.
Academic Press
,
New York
.
Pradel
,
J.
and
White
,
R. A. H.
(
1998
).
From selectors to realizators
.
Int. J. Dev. Biol
.
42
,
417
421
.
Rauskolb
,
C.
,
Peifer
,
M.
and
Wieschaus
,
E.
(
1993
).
extradenticle, a regulator of homeotic gene activity, is a homolog of the homeobox-containing human proto-oncogene pbx1
.
Cell
74
,
1101
1112
.
Regulski
,
M.
,
McGinnis
,
N.
,
Chadwick
,
R.
and
McGinnis
,
W.
(
1987
).
Developmental and molecular analysis of Deformed: A homeotic gene controlling Drosophila head development
.
EMBO J
.
6
,
767
777
.
Restifo
,
L. L.
and
Merrill
,
V. K. L.
(
1994
).
Two Drosophila regulatory genes, Deformed and the Broad-Complex, share common functions in development of adult CNS, head, and salivary glands
.
Dev. Biol
.
162
,
465
485
.
Rieckhof
,
G. E.
,
Casares
,
F.
,
Ryoo
,
H. D.
,
Abu-Shaar
,
M.
and
Mann
,
R. S.
(
1997
).
Nuclear translocation of Extradenticle requires homothorax, which encodes an Extradenticle-related homeodomain protein
.
Cell
91
,
171
183
.
Robertson
,
H. M.
,
Preston
,
C. R.
,
Phillis
,
R. W.
,
Johnson-Schlitz
,
D. M.
,
Benz
,
W. K.
and
Engels
,
W. R.
(
1988
).
A stable genomic source of P-element transposase in Drosophila melanogaster
.
Genetics
118
,
461
470
.
Röder
,
V.
Vola
,
C.
, and
Kerridge
,
S.
(
1992
).
The role of the teashirt gene in trunk segmental identity in Drosophila
.
Development
115
,
1017
1033
.
Rogge
,
R.
,
Green
,
P. J.
,
Urano
,
J.
,
Horn-Saban
,
S.
,
Mlodzik
,
M.
,
Shilo
,
B. Z.
,
Hartenstein
,
V.
and
Banerjee
,
U.
(
1995
).
The role of yan in mediating the choice between cell division and differentiation
.
Development
121
,
3947
3958
.
Sato
,
T.
,
Hayes
,
P. H.
and
Denell
,
R. E.
(
1985
).
Homeosis in Drosophila: Roles and spatial patterns of expression of the Antennapedia and Sex combs reduced loci in embryogenesis
.
Dev. Biol
.
111
,
171
192
.
Séguin
,
A.
,
Laible
,
G.
,
Leyva
,
A.
,
Dixon
,
R. A.
and
Lamb
,
C. J.
(
1997
).
Characterization of a gene encoding a DNA-binding protein that interacts in vitro with vascular specific cis elements of the phenylalanine ammonia-lyase promoter
.
Plant Mol. Biol
.
35
,
281
291
.
Serin
,
G.
,
Joseph
,
G.
,
Ghisolfi
,
L.
,
Bauzan
,
M.
,
Erard
,
M.
,
Amalric
,
F.
and
Bouvet
,
P.
(
1997
).
Two RNA-binding domains determine the RNA-binding specificity of nucleolin
.
J. Biol. Chem
.
272
,
13109
13116
.
Siomi
,
H.
and
Dreyfuss
,
G.
(
1997
).
RNA-binding proteins as regulators of gene expression
.
Curr. Opin. Genet. Dev
.
7
,
345
353
.
Staehling-Hampton
,
K.
,
Ciampa
,
P. J.
,
Brook
,
A.
and
Dyson
,
N.
(
1999
)
A genetic screen for modifiers of E2F
in
Drosophila melanogaster. Genetics
153
,
275
287
.
Tautz
,
D.
and
Pfeifle
,
C.
(
1989
).
A non-radioactive in situ hybridization method for the localization of specific RNAs in Drosophila embryos reveals translational control of the segmentation gene hunchback
.
Chromosoma
98
,
81
85
.
van Dijk
,
M.
and
Murre
,
C.
(
1994
).
extradenticle Raises the DNA Binding Specificity of Homeotic Selector Gene Products
.
Cell
78
,
617
624
.
Wakimoto
,
B. T.
and
Kaufman
,
T. C.
(
1981
).
Analysis of larval segmentation in lethal genotypes associated with the Antennapedia gene complex in Drosophila melanogaster
.
Dev. Biol
.
81
,
51
64
.
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.
Wright
,
T. R. F.
(
1996
).
Phenotypic analysis of the Dopa decarboxylase gene cluster mutants in Drosophila melanogaster
.
J. Hered
.
87
,
175
190
.