The type β transforming growth factor (TGF-β) family of secreted factors encompasses a wide range of structurally related proteins that control the state of determination or differentiation in a wide variety of cell types. For all members of the family that have been studied at the protein level, the active moieties arise as dimers of the C-terminal ∼110 amino acid fragment derived from much longer precursor polypeptides. The hallmark of the family is a series of 7 completely conserved cysteine residues in the C-terminus; other conserved amino acid sequences generally cluster in the vicinity of 6 of these 7 cysteines. This report focuses on our current understanding of the genetic structure and developmental role of the decapentaplegic (dpp) gene in Drosophila, the only member of the TGF-β family thus far identified in invertebrates. The dpp polypeptide bears a sufficiently close relationship to two bone morphogenesis proteins (BMP-2A and BMP-2B) identified in mammals (Wozney et al. 1988, Science 242, 1528-1534) to warrant the suggestion that dpp and the BMP-2s are the descendants of a common ancestral gene. The protein-coding information for dpp is contained within a 6 kb DNA segment. An elaborate cis-regulatory apparatus, encompassing a >55 kb DNA segment, has evolved to control expression of the dpp gene, which is required for determination of dorsal ectoderm in the early embryo, for normal distal outgrowth of the adult appendages, and for sundry other developmental events, which are currently less well-defined. Studies of chimeric individuals and observations of transcript accumulation in situ have demonstrated that the dpp gene is expressed along the A/P boundary of the imaginai disks. A possible role of dpp in elaborating positional information in imaginai disk development is discussed.

The TGF-β family comprises a diverse set of 12 polypeptides having profound (generally inhibitory) effects on the growth and/or differentiation of many cell types. Typically, there are no obvious common physiological themes which would have predicted the relationship of these polypeptides to one another, but rather their homology has been revealed through searches of available protein data bases. In the four years following the reported sequencing of the prototypical member of the family, now called TGF-β1 (Derynck el al. 1985), the family has grown to include one polypeptide identified in the amphibian Xenopus laevis [a vegetal pole specific maternal transcript, Vgl (Weeks and Melton, 1987)], 9 other polypeptides identified in humans and other mammals [a Vgl cross-homologous gene, Vgr-1 (Lyons et al. 1989), TGF-β2 (Marquardt el al. 1987), TGF-β3 (Dijke et al. 1988), the inhibin aand ft subunits (Mason et al. 1985; Mayo et al. 1986), Müllerian inhibiting substance, MIS (Cate et al. 1986), three bone morpho genesis proteins BMP-2A, BMP-2B and BMP-3 (Wozney et al. 1988)], and the decapentaplegic product in Drosophila [dpp (Padgett et al. 1987)].

The mammalian polypeptides are processed to generate secreted dimeric proteins which are thought to bind to cell surface receptors on target cells (Cheifetz et al.1986). The first functional demonstration of a biologically active receptor for TGF-β1 and TGF-β2 has recently been reported (Boyd and Massagué, 1989). No definitive information is available on signal transduction systems which are activated in target cells in response to ligand-receptor binding. Several members of the TGF-β family have been implicated as key factors in determinative or differentiative decisions, and therefore, this family of intercellular signalling molecules is of special interest to developmental biologists. In this report, I will briefly review the current state of information regarding the TGF-β family, and will then focus on our current view of the one member of the family thus far identified in invertebrates: the decapentaplegic gene in Drosophila.

In general, the members of the TGF-β family have been isolated on the basis of specific in vitro assays: growth inhibition, differentiation inhibition, asymmetric distributions of transcripts. Upon further study, however, it has been found that at least some of these polypeptides have profound and singular effects on the growth or differentiation of many other cell types. Furthermore, by in situ localization of transcripts or by immunohistochemistry, it has been shown that at least some TGF-β family gene products are localized in many discrete sites during vertebrate development. However, in most cases, specific in vivo roles remain uncertain.

TGF-β1, TGF-β2 and TGF-β3

TGF-β1 was discovered as one of two proteins (the other being TGF-α, which is a member of the EGF family of growth factors) present in conditioned medium of transformed mammalian cell lines which would cause phenotypic transformation of the immortalized NRK (normal rat kidney) cell line, such that it would now exhibit anchorage-independent growth in soft agar (Roberts et al. 1985). Since its initial discovery based on this phenotypic transformation assay, TGF-α1 activity and responsiveness have been found in many normal cell types in vivo and in cell culture. It is especially abundant in bone and platelets (Roberts et al. 1981; Assoian et al. 1985). Its physiological functions in vivo have not been established. In addition, its effects in vitro have proven quite wide ranging. Under many conditions of treatment, it turns out to be growth inhibitory rather than growth promoting (Roberts et al. 1985). In other circumstances, it leads to the differentiation of particular cell lines in culture (Ignotz and Massagué, 1985). In part, the effects of TGF-β1 seem to depend upon the profile of other factors which are added along with it (Assoian et al. 1984). Many of the actions of TGF-β1 can be explained as the indirect consequences of changes in the architecture of the extracellular matrix (Ignotz and Massagué, 1986; Allen-Hoffmann et al. 1988; Montesano and Orci, 1988).

Recent observations have uncovered two proteins (TGF-β2 and TGF-β3) which are close relatives of TGF-β1. Both TGF-β1 and TGF-β2 homodimers have been identified. They are somewhat different in their binding to putative TGF-β receptors (Cheifetz el al. 1987), suggesting that they might have different biological activities. TGF-β3 was recovered on the basis of its cross-homology to TGF-β1 and TGF-β2 (Dijke et al.1988). No information is available on its structure or biological effects.

Inhibin/activin

Inhibin, a secretion found in ovarian fluid, suppresses the release of follicle-stimulating hormone (FSH) by the pituitary gland (Ramasharma et al. 1984), thereby interfering with ovulation. It is an α/ βheterodimer in which both subunits bear structural homologies to the TGF-βs (Mason et al. 1985). In pigs, two genes encoding related βchains have been found (Mason et al. 1985). β-chain homodimers, or heterodimers of these β-chains, have been found to have the opposite effect of promoting FSH release; these proteins have been termed ‘activins’ (Ling et al. 1986; Vale et al. 1986).

Mullerian inhibiting substance

Müllerian inhibiting substance (MIS) is a homodimer secreted by the fetal testis that contributes to conversion of the indifferent urogenital system into the male reproductive system, by causing regression of the Müllerian ducts (Picon, 1969). It is thus a key factor in the determination of the somatic sexual phenotype of the mammalian fetus. Molecular analysis has revealed it to be a divergent member of the TGF-β family (Cate et al. 1986).

Bone morphogenesis proteins

These proteins were identified as products that have the ability to induce bone growth, using the ability to induce infiltration and differentiation of osteoblasts in demineralized bone as an assay system (Wang et al. 1988). Two of the proteins, BMP-2A and BMP-2B, are highly conserved (and hence their names), while BMP-3 is more divergent (Wozney et al. 1988). No information is available on the in vivo roles of these three proteins.

The vegetal pole specific transcript, Vgl

The sequencing of Vgl cDNAs has revealed that the protein encoded by Vgl transcripts is a member of the TGF-β family (Weeks and Melton, 1987). Vgl RNA was first identified as the only transcript that was highly concentrated at the vegetal pole of unfertilized Xeno-pus oöcytes (Melton, 1987; Pondel and King, 1988). As the embryo begins development, the maternal Vgl RNA becomes incorporated into vegetal cells, which will eventually form the endoderm. At this stage of gastrulation, endodermal cells induce ectoderm to form mesoderm. ATGF-β-like molecule has been implicated in this induction process (Slack et al. 1987; Kimelman and Kirschner, 1987; Rosa et al. 1987), using TGF-β1 and TGF-β2 polypeptides from mammalian sources. Given the distribution of Vgl RNA, it has been suggested that Vgl protein is a strong candidate to be the natural mesoderm-inducing factor (Weeks and Melton, 1987). Characterizations of Vgl and of mesoderminducing factor are the subjects of other reports in this volume.

Recently, using Vgl probes, Lyons et al. (1989) were able to isolate a cross-homologous mammalian gene from a mouse cDNA library. This gene, called Vgr-1 (Vgl-related), is expressed in numerous tissues throughout development. No information is available on the structure of the Vgr-1 protein, or on its biological role.

In every case in which the protein has been characterized, TGF-β family members are active as secreted dimers. The subunits of these dimers are about 110–130 amino acids in length and derive from longer precursors, ranging in size from approximately 300–600 amino acids. These precursors share several structural features. They all have N-terminal secretion signal sequences and several internal potential N-glycosylation sites; both of these features are consistent with the secreted nature of the polypeptides. No transmembrane domains are apparent in the protein sequences. The precursors all contain several basic residue pairs (arginine-arginine, arginine-lysine, lysine-lysine), at least some of which are the sites of cleavage in the processing of the precursor (e.g. Derynck el al. 1985). The amino acid conservations that characterize the TGF-β family are confined to the C-terminal 110 amino acid residues. As noted above, it is this C-terminal region that dimerizes to form the mature secreted product. In general, the extensive regions TV-terminal to the mature product are poorly conserved between different family members. Nonetheless, if the same member is examined in different vertebrate species, strong homologies in more N-terminal regions are identified as well (e.g. Derynck et al. 1986). In the case of TGF-/J1, a cleavage product of the N-terminal domain remains noncovalently associated with the C-terminal dimer in a latent form of the secreted molecule (Lyons et al. 1988).

With the recent expansion of the membership of the TGF-β family, it has been possible to discern subgroups on the basis of conserved or divergent structural features of the C-terminal portion of the polypeptides. The conservations among the members of the family generally reside in three domains surrounding six of the seven conserved cysteines present in each polypeptide. TGF-β1, TGF-β2, TGF-β3 and inhibin-βhave two additional conserved cysteines, placing these molecules in a different subfamily from the others. MIS and inhibin-o-show only a low level of conservation to the other polypeptides and to each other (∼20–25%). BMP-2A, BMP-213, BMP-3, Vgl, Vgr-1 and dpp exhibit at least 45% conservation (at the level of amino acid identities), and appear to be in their own subfamily. Below, we will discuss evidence that BMP-2A, BMP-2B and dpp are very closely related, and may actually represent the dipteran and mammalian derivatives of the same ancestral gene. The relationships of the different TGF-β family members to one another are demonstrated by a comparison of primary amino acid sequences of their C-termini (Fig. 1).

Fig. 1.

A comparison of the C-terminal sequences (∼100 amino acids) of the decapentaplegic polypeptide and several of the vertebrate members of the TGF-β family. The dpp polypeptide from Drosophila melanogaster is presented in the top line of the figure. Other members of the family, described in the text are listed below, in approximate rank order of amino acid similarity. The Vgl sequence is from Xenopus laevis. All other sequences are from mammals. Origins of sequences are given in the text. Amino acid positions that are blacked out (▪) are identical to the dpp amino acid. Positions that are shaded (▫) represent gaps shared with a gap in the dpp sequence. Thus, the more a sequence is blacked out or shaded, the more similar it is to the dpp polypeptide. A # symbol identifies the 7 completely conserved cysteines in the C-termini of all TGF-β family members. The asterisk at the end of each sequence indicates the presence of a stop codon at aligned positions for each polypeptide.

Fig. 1.

A comparison of the C-terminal sequences (∼100 amino acids) of the decapentaplegic polypeptide and several of the vertebrate members of the TGF-β family. The dpp polypeptide from Drosophila melanogaster is presented in the top line of the figure. Other members of the family, described in the text are listed below, in approximate rank order of amino acid similarity. The Vgl sequence is from Xenopus laevis. All other sequences are from mammals. Origins of sequences are given in the text. Amino acid positions that are blacked out (▪) are identical to the dpp amino acid. Positions that are shaded (▫) represent gaps shared with a gap in the dpp sequence. Thus, the more a sequence is blacked out or shaded, the more similar it is to the dpp polypeptide. A # symbol identifies the 7 completely conserved cysteines in the C-termini of all TGF-β family members. The asterisk at the end of each sequence indicates the presence of a stop codon at aligned positions for each polypeptide.

In summary, in the course of the four years since the first of these sequences - TGF-β1-was published, this vertebrate polypeptide family has come to include a diverse and physiologically unrelated set of factors with a range of very intriguing developmental effects. The homology of the decapentaplegic (dpp) polypeptide, which will be described in the next section, demonstrates that at least one ancestor of this gene family predates the divergence of arthropods from vertebrates, and provides yet another set of important developmental functions controlled by a representative of the TGF-β family.

We have now characterized several cDNAs corresponding to 5 different classes of overlapping dpp transcripts (Padgett et al. 1987; St. Johnston et al. 1989; Padgett et al. in preparation). Each transcript appears to encode the same polypeptide. We have not addressed the questions of possible heterogeneity among the transcripts in terms of stability or translational efficiency. The structures of the cDNAs that correspond to these transcripts are depicted in Fig. 2. These five transcript classes have three exons apiece, and they differ only in their first (j.e. 5’) untranslated exon. These untranslated exons appear to be initiated by different promoters and range in size from about 150 bases up to 2 kb. Each transcript class utilizes the same splice acceptor site in the second exon. There is only a single pattern of splicing between the second and third exons.

Fig. 2.

The molecular map of the dpp gene. This map summarizes information presented by St. Johnston et al. (1989). The positions of the major segments of the dpp gene (shv-Hin-disk) and of the subdivisions of the shv and disk regions are depicted, and are based on the molecular mapping of mutant breakpoints. Double-headed arrows above the molecular map indicate uncertainties in the positions of these boundaries. Distal and proximal indicate the orientation of the dpp gene relative to the chromosome 2 (left arm) telomere and centromere, respectively. Below the molecular map are noted the positions of two tRNATyr genes within dpp (Suter et al. 1990). A standard molecular scale, from 70–120 kb, is depicted. The positions of the two exons common to all dpp transcript species, falling within the Hin region, are shown. The open boxes within these exons represents the extent of the open reading frame of the dpp transcripts, and the black box represents 3’ untranslated material (Padgett et al. 1987; Padgett et al. in preparation). At least 5 different 5’ exons have been identified in different dpp cDNAs.

Fig. 2.

The molecular map of the dpp gene. This map summarizes information presented by St. Johnston et al. (1989). The positions of the major segments of the dpp gene (shv-Hin-disk) and of the subdivisions of the shv and disk regions are depicted, and are based on the molecular mapping of mutant breakpoints. Double-headed arrows above the molecular map indicate uncertainties in the positions of these boundaries. Distal and proximal indicate the orientation of the dpp gene relative to the chromosome 2 (left arm) telomere and centromere, respectively. Below the molecular map are noted the positions of two tRNATyr genes within dpp (Suter et al. 1990). A standard molecular scale, from 70–120 kb, is depicted. The positions of the two exons common to all dpp transcript species, falling within the Hin region, are shown. The open boxes within these exons represents the extent of the open reading frame of the dpp transcripts, and the black box represents 3’ untranslated material (Padgett et al. 1987; Padgett et al. in preparation). At least 5 different 5’ exons have been identified in different dpp cDNAs.

The first AUG in the open reading frame (ORF) of each transcript is located near the beginning of the common second exon. The ORF extends through the middle of the third exon. If translation initiated with the first AUG, the polypeptide that was produced would be 588 amino acids in length (Padgett et al. 1987). We think it likely that the first AUG of the ORF does indeed represent the actual translation initiation codon because (a) it conforms to the consensus sequence for Drosophila initiation codons and (b) unlike other AUGs in the ORF, it is followed by a consensus N- terminal signal secretion sequence (Watson, 1984), as expected for a secreted TGF-)5-like polypeptide.

Several potential glycosylation sites can be found within the dpp ORF, consistent with the view that the dpp polypeptide is itself a secreted protein. Several dibasic residues are found within the ORF, suggesting that the primary polypeptide product might be proteolytically cleaved to generate a smaller mature product. The positions of these sites are diagrammed in Fig. 3.

Fig. 3.

A schematized view of the dpp polypeptide. The full-length product of the dpp ORF is a 588 amino acid protein containing an N-terminal signal peptide and a C-terminal region exhibiting high conservation with other members of the TGF-β family. Positions of dibasic residues, some of which may serve as cleavage sites for serine proteases, and N- glycosylation sites are shown as indicated. While the dpp polypeptide is the longest member of the family, the basic structure of the molecule parallels the structures of the other family members.

Fig. 3.

A schematized view of the dpp polypeptide. The full-length product of the dpp ORF is a 588 amino acid protein containing an N-terminal signal peptide and a C-terminal region exhibiting high conservation with other members of the TGF-β family. Positions of dibasic residues, some of which may serve as cleavage sites for serine proteases, and N- glycosylation sites are shown as indicated. While the dpp polypeptide is the longest member of the family, the basic structure of the molecule parallels the structures of the other family members.

The C-terminal ∼110 amino acids of the dpp polypeptide show extensive similarity to the comparable regions of the vertebrate TGF-β family members (Fig. 1). Three arginine-arginine dipeptides reside just A-terminal to this region of high conservation; these potential proteolytic cleavage sites are thus positioned analogously to the known cleavage sites in most of the other family members. The dpp polypeptide contains the same seven cysteines in its putative C-terminal fragment as are shared by the vertebrate family members. The locations of the similarities between the dpp polypeptide and the vertebrate members of the TGF-β family conform to the regions of similarity among the vertebrate family members themselves. These conservations reside in three domains surrounding six of the seven C-terminal cysteines. We presume this pattern of conservation reflects a common structural motif in this family of molecules.

With the recent report of the primary sequences of human BMP-2A and BMP-2B (Wozney et al. 1988), we now have two very strong candidates for the vertebrate analogues of dpp. The C-termini of BMP-2A and BMP-2B are each 75 % identical in amino acid sequence to the corresponding region of the dpp polypeptide. Moreover, within the N-terminal precursor region, there is 25–30% amino acid identity. The only other polypeptides in the family which show any significant conservation in the N-termini are between BMP-2A and BMP-2B themselves, and among TGF-β1, TGF-β2 and TGF-β3. Based on these considerations, we think it very likely that the dpp gene of Drosophila and the BMP-2A and BMP-2B genes of humans are the modern highly conserved derivatives of a gene which was present in an ancestor common to the arthopod and vertebrate lineages.

The organization of the dpp locus

Numerous observations on the structure and function of the dpp gene have coalesced into a clear picture of its organization. In brief, the dpp gene (map location 2–4.0; 22F1,2) is a rather large genetic unit (>55kb), which produces a single TGF-β-homologous polypeptide (St. Johnston et al. 1990; Padgett et al. in preparation). This polypeptide is encoded by at least 5 overlapping transcripts (St. Johnston et al. 1990; Padgett et al. in preparation). These transcripts are initiated by different promoters and contain different 5’ untranslated exons. Each transcript has its own temporal (and presumably spatial) pattern of expression (Gelbart et al. 1985; St. Johnston and Gelbart, 1987; St. Johnston et al. 1989; Posakony et al. 1990). The bulk of the genetic information contained in the >55 kb of dpp DNA consists of c/s-regulatory elements controlling the timing, location and quantity of dpp transcription (Gelbart et al. 1985; St. Johnston et al. 1990; Blackman et al. in preparation). These c/s-regulatory elements must drive a very complex pattern of expression, as the dpp gene is required for numerous developmental events, including determination of dorsal ectoderm along the dorsoventral axis of the early embryo (Irish and Gelbart, 1987; St. Johnston and Gelbart, 1987), proper morphogenesis of the adult appendages derived from the imaginai disks (Spencer et al. 1982; Posakony et al. 1990; Spencer et al. in preparation), and other, less well-defined events necessary for proper development of the larva and adult (Segal and Gelbart, 1985).

Some mutations in the dpp gene interfere with the production of an active dpp polypeptide product. They are identified as lesions that affect all developmental events controlled by dpp (Irish and Gelbart, 1987). Null (amorphic) mutations are called dppHin alleles, because they are Haplo-insufficient that is, dppHin/ + animals are lethal, apparently because there is too little dpp product in the early embryo to produce proper determination along the dorsoventral axis. Most of these are likely to act by altering the structure of the dpp polypeptide, while the remainder probably derange the structure of all of the transcripts. Some dppHin alleles have lesions that can be identified by restriction mapping; these lesions cluster in a region of about 6 kb which we term the Hin region (St. Johnston et al. 1990). This region also contains the two exons encoding the open reading frame for the dpp polypeptide. Leaky (hypomorphic) mutations are not haplo-insufficient; because they map to the Hin region but are recessive, they are called dpphin-ralleles. All dpphin-r alleles have normal restriction maps in the Hin region. We think it likely that some of the dppHin and many of the dpphin-rlesions are point mutations altering the amino acid sequence of the dpp polypeptide.

Other dpp mutations affect only a subset of the diverse array of dpp functions. We think that these mutations (virtually all of which are inversions, translocations or deletions: Spencer et al. 1982; Segal and Gelbart, 1985; St. Johnston et al. 1990) represent lesions disrupting batteries of cfs-regulatory elements necessary for the proper spatial and temporal deployment of dpp transcripts. This is best exemplified by a consideration of the disk region, which is a >25 kb untranscribed region 3′ to the polyadenylation sites of all dpp transcripts.

Mutations in the disk region are recessive lesions that disrupt proper dpp expression in the imaginai disks, leading to the production of distally incomplete appendages. Several classes of such mutations have been described, with the most severe removing all but the most proximal regions of each appendage. All mutations of the disk region are recessive to wild-type.

The mildest alleles of the disk region are defective only in wing posture. These ‘heldout’ (dppd-ho) mutations cause the wings to be held out laterally and perpendicular to the body axis; this phenotype is associated with the loss of specific pattern elements (a group of 25 sensory structures called the sensilla campaniformia-25 or Sc25) on the dorsal base of the wing (Spencer et al. 1982). Most mutations of the disk region have been recovered as dppd-ho alleles (Fig. 4). When examined in homozygotes, most have a more severe array of mutant phenotypes (Spencer et al. 1982; Irish and Gelbart, 1987; Blackman et al. 1987). These mutations are generally associated with gross chromosomal rearrangements. Mutant homozygotes lack structures derived from distal portions of the adult appendages. Mild alleles (dppdisk-11) only affect the wing, halter and male genital derivatives. Intermediate alleles (dppdisk-11) affect all major appendages (wing, halter, leg, eye, antenna, proboscis, male and female genitalia), but sufficient appendages remain to permit dpp homozygotes to survive to adulthood. Severe alleles (dppdisk-11) cause such severe defects in all of the appendages that they engender death early in pupation. The defects in the adult structures are due to corresponding abnormalities in the imaginai disks (the precursors of the adult appendages). Distal structures of adult appendages are known to derive from central regions of the imaginai disks (Bryant, 1978). These central regions are defective in these dpp homozygotes. The amount of imaginai disk material lacking in these mutant individuals parallels the severity of the adult defects (Spencer et al. 1982).

Fig. 4.

The origins and types of mutations recovered in heldout mutagenesis screens. TOP: The basic scheme for generating dpp mutations as alleles of the original heldout (dppd-ho) lesion is shown. Several permutations of this screen, in which different mutagens (X-rays, hobo transposon mobilization, ethyl methanesulfonate [EMS]) were employed, extra copies of the Hin region were included to recover dppnull alleles, and/or rearrangements were included on the dppd-ho tester chromosome to obviate dpp transvection effects (Gelbart, 1982). BOTTOM: Specific examples of the three major types of dpp lesions recovered as heldout alleles are depicted below a molecular map of the dpp region (St. Johnston et al. 1990). On the molecular map, the subdivisions of the disk region are included. The boxes above the map indicate the position of relevant dpp exons. The two common exons are indicated in open and black boxes, as in Fig. 2. The striped boxes indicate the positions of some of the alternative 5’ exons containing untranslated information. The dppd-ho mutation is a ∼2.8kb deletion from molecular position 112–114.8 on the molecular map of dpp. It presumably deletes a cis- regulatory element necessary to elaborate the portion of the dorsal base of the wing which includes a group of mechanoreceptors called the sensilla campaniformia-25 (Sc25) (Spencer et al. 1982). Other lesions which delete this region of the gene also confer the heldout phenotype and lack the Sc25. The dppHm+6 lesion is a ∼0.5 kb deletion within the first of the two common exons, and hence within the dpp ORF. It behaves as a null allele of dpp (although it is transvection-sensitive) and hence was recovered as a heldout allele (Irish and Gelbart, 1987). ln(2L)dppd6 is an example of the largest class of heldout alleles recovered by X-irradiation. It is a cytologically visible inversion of the dppdisk-111class, and has one breakpoint at ∼102 on the molecular map. In(2L)dppd6 splits the dpp gene into two pieces. The fragment containing the shv-Hin region retains its residual activities of the shv, Hin and disk-V regions. The fragment containing the disk-HI, disk-II and disk-ho cis-regulatory regions is no longer adjacent to the dpp transcription unit, and is thus are inactive. Hence, this inversion was recovered as a heldout allele. Most dppdlk mutations recovered as heldout alleles are such rearrangement breakpoints falling between the disk-hocis- regulatory element and the dpp transcription unit.

Fig. 4.

The origins and types of mutations recovered in heldout mutagenesis screens. TOP: The basic scheme for generating dpp mutations as alleles of the original heldout (dppd-ho) lesion is shown. Several permutations of this screen, in which different mutagens (X-rays, hobo transposon mobilization, ethyl methanesulfonate [EMS]) were employed, extra copies of the Hin region were included to recover dppnull alleles, and/or rearrangements were included on the dppd-ho tester chromosome to obviate dpp transvection effects (Gelbart, 1982). BOTTOM: Specific examples of the three major types of dpp lesions recovered as heldout alleles are depicted below a molecular map of the dpp region (St. Johnston et al. 1990). On the molecular map, the subdivisions of the disk region are included. The boxes above the map indicate the position of relevant dpp exons. The two common exons are indicated in open and black boxes, as in Fig. 2. The striped boxes indicate the positions of some of the alternative 5’ exons containing untranslated information. The dppd-ho mutation is a ∼2.8kb deletion from molecular position 112–114.8 on the molecular map of dpp. It presumably deletes a cis- regulatory element necessary to elaborate the portion of the dorsal base of the wing which includes a group of mechanoreceptors called the sensilla campaniformia-25 (Sc25) (Spencer et al. 1982). Other lesions which delete this region of the gene also confer the heldout phenotype and lack the Sc25. The dppHm+6 lesion is a ∼0.5 kb deletion within the first of the two common exons, and hence within the dpp ORF. It behaves as a null allele of dpp (although it is transvection-sensitive) and hence was recovered as a heldout allele (Irish and Gelbart, 1987). ln(2L)dppd6 is an example of the largest class of heldout alleles recovered by X-irradiation. It is a cytologically visible inversion of the dppdisk-111class, and has one breakpoint at ∼102 on the molecular map. In(2L)dppd6 splits the dpp gene into two pieces. The fragment containing the shv-Hin region retains its residual activities of the shv, Hin and disk-V regions. The fragment containing the disk-HI, disk-II and disk-ho cis-regulatory regions is no longer adjacent to the dpp transcription unit, and is thus are inactive. Hence, this inversion was recovered as a heldout allele. Most dppdlk mutations recovered as heldout alleles are such rearrangement breakpoints falling between the disk-hocis- regulatory element and the dpp transcription unit.

We have several lines of indirect evidence that the disk region is cú-regulatory in nature, (a) When duplications covering the Hin region, but not the disk region are used to rescue pseudopoint dppHin mutations past the Haplo-lethai embryonic period, dppHin mutations exhibit a phenotype indistinguishable from the most severe disk region phenotype, exhibited by dppdisk-Vhomozygotes, (b) The pattern of expression of the dpp transcripts emerging from the shv-Hin region is indistinguishable from the location of imaginai disk cells requiring normal disk region expression for proper appendage formation (Spencer et al. in preparation; Posakony et al. 1990). (c) We have no clear evidence of transcriptional activity within the disk region (St. Johnston et al. 1989). (d) The closer to the transcribed shv-Hin region that mutant breakpoints in the disk region fall, the more severe are their mutant phenotypes (Fig. 2). All of these mutations break within the disk region, and can be viewed as removing all disk region cis-regulatory DNA proximal to the breakpoint from the neighborhood of the dpp transcription unit (Fig. 4).

Based on these observations, we conclude that the disk region acts by controlling the expression patterns of shv-Hin transcripts in the imaginai disks. Similar considerations lead us to conclude that the Hin region must contain c/s-regulatory elements for the embryonic ectodermal expression of the dpp transcripts (St. Johnston and Gelbart, 1987; Irish and Gelbart, 1987; Hoffmann and Goodman, 1987), and the shv region must contain cis-regulatory elements driving still other spatial and temporal patterns of dpp expression (Segal and Gelbart, 1985; St. Johnston et al. 1990).

The nature of the dpp protein product, combined with the developmental effects of dpp mutations, has led us to speculate that dpp contributes to Drosophila development at the level of intercellular signalling of positional information (Padgett et al. 1987). By analysis of chimeric individuals containing marked dpp∼ tissue, we have made several observations (chiefly on wing disks and their derivatives) which bear on possible models of dpp’s developmental role (Spencer et al. in preparation; Posakony et al. 1990). (1) In mosaics that are mixtures of genotypically dpp+ and dppd-111 tissue, each imaginai disk solely determines the dpp phenotype of its consequent adult appendage. (2) In those mosaics in which a single imaginai disk contains a mixture of genotypically wild-type and mutant cells, the derivative adult appendages may be fully normal, even though 50 % or more of the pattern elements that would be lacking in wholly mutant appendages are of the mutant genotype. (3) Alternatively, such disks of mixed dpp genotype can sometimes give rise to appendages that are partially or fully mutant in phenotype. (4) The dpp phenotype of a genotypically mixed imaginai disk depends solely upon the genotype of the cells in the region of the disk just anterior to, and abutting, the anterior-posterior (A/P) compartment boundary. We term this region the ‘focus’ of dpp disk expression. (5) dpp transcripts accumulate in a pattern concordant with the position of the dpp focus, i.e. along the A/P compartment boundary.

Based on these results, we conclude that dpp cannot act as an exocrine function, secreted from one organ to act on distant target tissues. However, within an imaginai disk primordium, we cannot determine if the primary action of dpp is strictly local, or if it acts on target cells throughout the disk.

One surprising aspect of our observations on dpp expression within the imaginai disks is its dissimilarity to the phenotypes dpp elicits in adult appendages. While the basic phenotype engendered by dppdisk mutations are the loss of tissue in a center to peripheral (radial) direction within the disk, transcript accumulation and localization of the dpp focus occur as a stripe cutting across the entire disk at the level of the A/P compartment boundary. In other words, there is no apparent asymmetry of dpp expression along the radial axis of the imaginai disk. How then does dpp cause pattern deletions along the radial axis, leading to distally incomplete appendages?

One possible explanation is suggested by Meinhardt (1982). Based on theoretical considerations, he supposed that intersections of compartment boundaries (A/P and dorsal/ventral) would identify unique sets of cells, which would then serve as point sources of radial positional information. Given our observations on dpp expression in disks, I suggest that the strip of expression of dpp (which is constrained by the position of the A/P compartment boundary) serves as a reference line for a series of intersections which lead to the elaboration of radial positional information (detailed in Fig. 5). A gradient of dpp product within the disk may be one component of this positional signalling system, or alternatively, the localized expression of dpp along the A/P boundary may contribute to the activation of a different positional signalling molecule. In either case, it would be the integrated information from the intersection of dpp expression and the expression of other localized molecules within the disk that would lead to the elaboration of radial positional information. The wingless gene, which encodes an apparent secreted factor and which is expressed in a different but also highly localized pattern in all disks, is an intriguing candidate for another component of the positional signalling system (Cabrera et al. 1987; Rijsewijk et al. 1987; Baker, 1988).

Fig. 5.

A possible sequence of events in the elaboration of radial positional information within developing imaginai disks. In this scheme, first the disk primordium is divided into anterior (A) and posterior (P) compartments by the differential expression of the engrailed gene (Lawrence and Morata, 1976; Kornberg et al. 1985). dpp expression then occurs just anterior to the A/P boundary, forming a reference line running from the dorsal to the ventral margin of the disk proper. Another gene product is then locally activated in another orientation within the disk, such that it intersects the dpp reference line. The cells at the intersection point have the unique property of expressing both gene products. These cells might then elaborate a radially graded positional signal involving some third molecule, or radial information may be transmitted as direct gradients of the products of dpp and this other, undefined gene. In the last frame, another, developmentally later product intersecting the dpp reference line is shown, to indicate that multiple rounds of intersections and graded signals are likely to be necessary to elaborate the structures typical of an asymmetric primordium such as the wing disk. The basic idea for this model derives from the suggestions of Meinhardt (1982) and from the observations on dpp expression by Posakony et al. (1990).

Fig. 5.

A possible sequence of events in the elaboration of radial positional information within developing imaginai disks. In this scheme, first the disk primordium is divided into anterior (A) and posterior (P) compartments by the differential expression of the engrailed gene (Lawrence and Morata, 1976; Kornberg et al. 1985). dpp expression then occurs just anterior to the A/P boundary, forming a reference line running from the dorsal to the ventral margin of the disk proper. Another gene product is then locally activated in another orientation within the disk, such that it intersects the dpp reference line. The cells at the intersection point have the unique property of expressing both gene products. These cells might then elaborate a radially graded positional signal involving some third molecule, or radial information may be transmitted as direct gradients of the products of dpp and this other, undefined gene. In the last frame, another, developmentally later product intersecting the dpp reference line is shown, to indicate that multiple rounds of intersections and graded signals are likely to be necessary to elaborate the structures typical of an asymmetric primordium such as the wing disk. The basic idea for this model derives from the suggestions of Meinhardt (1982) and from the observations on dpp expression by Posakony et al. (1990).

In essence, then, I am proposing that dpp contributes to the establishment of central positions within developing imaginai disks. The establishment of these central positions would then, directly or indirectly, lead to the elaboration of radial positional information. Perhaps the establishment of central positions, together with the pre-existence of peripheral positional values, could lead to the elaboration of intermediate positions through a process of intercalation. While there are some difficulties in fitting the range of dpp adult appendage phenotypes into such a model, it does have the attractive feature of relating to observations coming from so-called ‘competence’ experiments in imaginai disk development. As first noted by Schubiger (1974) in transplantation experiments in which immature leg disks were forced to metamorphose precociously, there is a regional hierarchy in the abilities of disk cells to differentiate adult cuticular structures. In the youngest disks at all competent to elaborate adult structures, structures from the very periphery and the very center of the disk differentiate. In somewhat older disks, larger swatches of peripheral and of central tissue differentiates. As the age of the disk increases, the size of the swatches increases until finally, in mature disks, the entire radial pattern is differentiated. A similar scenario holds for the wing disk (Bownes and Roberts, 1979). It is as if the cells forming center and periphery are first specified during development, and as if further specification occurs by intercalation. Curiously, the first structures from the leg and the wing that are removed by the mildest dpp alleles affecting those tissues are precisely the first structures competent to be differentiated from the central regions of those disks (the tarsal claws of the leg in dppd-111homozygotes, and the Sc25 of the dorsal base of the wing in dppd-ho homozygotes). Thus, dpp mutations and competence experiments appear to identify the same set of structures as central within the developing disk. Whether this relationship is of importance in understanding the molecular basis of positional specification within the developing disks must await future experimentation.

This work summarizes the excellent and invaluable intellectual and technical contributions of many investigators over a ten year period, most notably (in alphabetical order) Ronald Blackman, F. Michael Hoffmann, Vivian Irish, Richard W. Padgett, Leila Posakony, Laurel Raftery, Daniel Segal, Forrest Spencer and R. Daniel St. Johnston. This work has been supported by research grants from the Public Health Service and from the American Cancer Society.

Allen-Hoffmann
,
B. L
,
Crankshaw
,
C. L.
and
Mosher
,
D. F
(
1988
).
Transforming growth factor /J increases cell surface binding and assembly of exogenous (plasma) fibronectin by normal human fibroblasts
.
Molec. cell. Biol
.
10
,
4234
4242
.
Assoian
,
R. K.
, GROTENDORST. G. R.,
Miller
,
D. M.
and
Shorn
,
M. B.
(
1984
).
Cellular transformation by coordinated action of three peptide growth factors from human platelets
.
Nature, Land
.
309
,
804
806
.
Assoian
,
R. K.. KOMORIGA. A.
,
Meyers
,
C A.. MILLER. D. M
and
Sporn
,
M. B.
(
1985
).
Transforming growth factor-/! in human platelets: Identification of a major storage site, purification and characterization
.
J. biol. Chem
.
258
.
7155
7160
.
Baker
,
N. E.
(
1988
).
Transcription of the segment-polarity gene wingless in the imaginai discs of Drosophila, and the phenotype of a pupal-lethal ivg mutation
.
Development
102
.
489
497
.
Blackman
,
R. K.
,
Grimaila
,
R.
,
Koehler
,
M. M. D.
and
Gelbart
,
W. M.
(
1987
).
Mobilization of hobo elements residing within the decapentaplegic gene complex: Suggestion of a new hybrid dysgenesis system in Drosophila tnelanogaster
.
Cell
49
,
497
-
505
.
Bownes
,
M
and
Roberts
,
S.
(
1979
).
Acquisition of differentiative capacity in imaginai wing discs of Drosophila tnelanogaster
,
J. Embryol. exp. Morph
.
49
,
103
113
Boyd
,
F. T.
and MASSAGUé. j
. (
1989
).
Transforming growth factor-β inhibition of epithelial cell proliferation linked to the expression of a 53-kDa membrane receptor
.
J. biol Chem
264
.
2272
2278
.
Bryant
,
P. J
(
1978
).
Pattern formation in imaginai discs
. In
Genetics and Biology of Drosophila, vol. 2c
(ed.
M.
Ashburner
and
T.R.F.
Wright
), pp.
230
335
,
London
:
Academic Press
Cabrera
,
C. V.
,
Alonso
,
M. C.
,
Johnston
,
P.
,
Phillips
,
R. G.
and
Lawrence
,
P. A.
(
1987
).
Phenocopies induced with antisense RNA identify the wingless gene
.
Cell
50
,
659
663
.
Cate
,
R. L.
,
Mattaliano
,
R. J.
,
Hession
,
C.
,
Tizard
,
R.
,
Farber
,
N. M.
,
Cheung
,
A.
,
Ninfa
,
E. G.
,
Frey
,
A. Z.
,
Gash
,
D. J.
,
Chow
,
E. P.
,
Fisher
,
R. A.
,
Bertonis
,
J. M.
,
Torres
,
G.
,
Wallner
,
B. P.
,
Ramachandran
,
K. L.
,
Ragin
,
R. C.
,
Managanaro
,
T. F.
,
Maclaughlin
,
D. T.
and
Donahoe
,
P. K.
(
1986
).
Isolation of the bovine and human genes for Müllerian inhibiting substance and expression of the human gene in animal cells
.
Cell
45
,
685
698
.
Cheifetz
,
S.
,
Like
,
B.
and
Massagué
,
J.
(
1986
).
Cellular distribution of type I and II receptors for transforming growth factor-β
.
J. biol. Chem
.
261
,
9972
9978
.
Cheifetz
,
S.
,
Weatherbee
,
J. A.
,
Tsang
,
M. L.-S.
,
Anderson
,
J. K.
,
Mole
,
J. E.
,
Lucas
,
R.
and
Massagué
,
J.
(
1987
).
The transforming growth factor-β system, a complex pattern of crossreactive ligands and receptors
.
Cell
48
,
409
415
.
Derynck
,
R.
,
Jarrett
,
J. A.
,
Chen
,
E. Y.
,
Eaton
,
D. H.
,
Bell
,
J. R.
,
Assoian
,
R. K.
,
Roberts
,
A. B.
,
Sporn
,
M. B.
and
Goeddel
,
D. V.
(
1985
).
Human transforming growth factor-β complementary DNA sequence and expression in normal and transformed cells
.
Nature, Land
.
316
,
701
705
.
Derynck
,
R.
,
Jarrett
,
J. A.
,
Chen
,
E. Y.
and
Goeddel
,
D. V.
(
1986
).
The murine transforming growth factor-β precursor
.
J. biol. Chem
.
261
,
4377
4379
.
Duke
,
P. T.
,
Hansen
,
P.
,
Iwata
,
K. K.
,
Pieler
,
C.
and
Foulkes
,
J. G.
(
1988
).
Identification of another member of the transforming growth factor type P gene family
.
Proc. natn. Acad. Sci. U.S.A
.
85
,
4715
4719
.
Gelbart
,
W. M.
(
1982
).
Synapsis-dependent allelic complementation at the decapentaplegic gene complex in Drosophila melanogaster
.
Proc. natn. Acad. Sci. U.S.A
.
79
,
2636
2640
.
Gelbart
,
W. M.
,
Irish
,
V. F.
,
St Johnston
,
R. D.
,
Hoffmann
,
F. M.
,
Blackman
,
R. K.
,
Segal
,
D.
,
Posakony
,
L. M.
and
Grimaila
,
R.
(
1985
).
The decapentaplegic gene complex in Drosophila melanogaster
.
Cold Spring Harbor Symp. quant. Biol
.
50
,
119
125
.
Hoffmann
,
F. M.
and
Goodman
,
W.
(
1987
).
Identification in transgenic animals of Drosophila decapentaplegic sequences required for embryonic dorsal pattern formation
.
Genes & Devel
.
1
,
615
625
.
Ignotz
,
R. A.
and
Massagué
,
J.
(
1985
).
Type ft transforming growth factor controls the adipogénie differentiation of 3T3 fibroblasts
.
Proc. natn. Acad. Sci. U.S.A
.
82
,
8530
8534
.
Ignotz
,
R. A.
and
Massagué
,
J.
(
1986
).
Transforming growth factor-β stimulates the expression of fibronectin and collagen and their incorporation into the extracellular matrix
.
J. biol. Chem
.
261
,
4337
4345
.
Irish
,
V. F.
and
Gelbart
,
W. M.
(
1987
).
The decapentaplegic gene is required for dorsal-ventral patterning in Drosophila embryos
.
Genes á Devel
.
1
,
868
879
.
Kimelman
,
D.
and
Kirschner
,
M.
(
1987
).
Synergistic induction of mesoderm by FGF and TGF-/J and the identification of an mRNA coding for FGF in the early xenopus embryo
.
Cell
51
,
869
877
.
Kornberg
,
T.
,
Siden
,
I.
,
O’farrell
,
P.
and
Simen
,
M.
(
1985
).
The engrailed locus of Drosophila-. In situ localization of transcripts reveals compartment-specific expression
.
Cell
40
,
45
53
.
Lawrence
,
P. A.
and
Morata
,
G.
(
1976
).
Compartments in the wing of Drosophila: A study of the engrailed gene
.
Devi Biol
.
50
,
321
337
.
Ling
,
N.
, YING. S-Y.,
Ueno
,
N.
,
Shimasaki
,
S.
,
Esch
,
F.
,
Hotta
,
M.
and
Guillemin
,
R.
(
1986
).
Pituitary FSH is released by a heterodimer of the β-subunits from the two forms of inhibin
.
Nature, Lond
.
321
,
779
782
.
Lyons
,
K.
,
Graycar
,
J. L.
,
Lee
,
A.
,
Hashmi
,
S.
,
Lindquist
,
P. B.
,
Chen
,
E. Y.
,
Hogan
,
B. L. M.
and
Derynck
,
R.
(
1989
).
Vgr-1, a mammalian gene related to Xenopus Vg-I, is a member of the transforming growth factor β gene superfamily
.
Proc. natn. Acad Sci. U.S.A
.
86
,
4554
4558
.
Lyons
,
R. M.
,
Keski-Oja
,
J.
and
Moses
,
H. L.
(
1988
).
Proteolyticactivation of latent transforming growth factor-/? from fibroblast-conditioned medium
.
J. Cel! Biol
.
106
,
1659
1665
.
Marquardt
,
H.
,
Lioubin
,
M. N.
and
Ikeda
,
T.
(
1987
).
Complete amino acid sequence of human transforming growth factor type p2
.
J. biol. Chem
.
262
,
12127
12131
.
Mason
,
A. J.
,
Hayflick
,
J. S.
,
Ling
,
N.
,
Esch
,
F.
,
Ueno
,
N.
,
Ying
,
S-Y.
,
Guillemin
,
R.
,
Naill
,
H.
and
Seeburg
,
R. H.
(
1985
).
Complementary DNA sequences of ovarian follicular fluid inhibin show precursor structure and homology with transforming growth factor-/!
.
Nature, Lond
.
318
,
659
663
.
Mayo
,
K. E.
,
Cerelu
,
G. M.
,
Spiess
,
J.
,
Rivier
,
J.
,
Rosenfeld
,
M. G.
,
Evand
,
R. M.
and
Vale
,
W.
(
1986
).
Inhibin A-subunit cDNAs from porcine ovary and human placenta
.
Proc. natn. Acad. Sci. U.S.A
.
83
,
5849
5853
.
Meinhardt
,
H.
(
1982
).
Generation of structures in a developing organism
. In
Developmental Order: Its Origin and Regulation
(ed.
S.
Subtelny
and
P.B.
Green
), pp.
439
461
.
New York
:
Alan Liss
.
Melton
,
D. A.
(
1987
).
Translocation of a localized maternal mRNA to the vegetal pole of Xenopus oocytes
.
Nature, Lond
.
328
,
80
82
.
Montesano
,
R.
and
Orci
,
L.
(
1988
).
Transforming growth factor βJ stimulates collagen-matnx contraction by fibroblasts: Implications for wound healing
.
Proc. natn. Acad. Sci. U.S.A
.
85
,
4894
4897
.
Padgett
,
R. W.
,
St Johnston
,
R. D.
and
Gelbart
,
W. M.
(
1987
).
The decapentaplegic gene complex of Drosophila encodes a protein homologous to the transforming growth factor-β gene family
.
Nature, Lond
.
325
,
81
84
.
Picon
,
R.
(
1969
).
Action du testicule foetal sur le development in vitro des canaux de Müller chez le rat
.
Arch. Anat. Micro. Morph, exp
.
58
,
1
19
.
Pondel
,
M. D.
and
King
,
M. L.
(
1988
).
Localized maternal mRNA related to transforming growth factor β mRNA is concentrated in a cytokeratin-enriched fraction from Xenopus oocytes
.
Proc. natn. Acad Sci. U.S.A
.
85
,
7612
7616
.
Posakony
,
L. M.
,
Raftery
,
L. A.
,
St Johnston
,
R. D.
and
Gelbart
,
W. M.
(
1990
).
Wing formation in Drosophila melanogaster requires decapentaplegic gene function along the anterior-posterior compartment boundary
.
Genes & Devel. (submitted)
.
Ramasharma
,
K.
,
Sairam
,
M. R.
,
Seidak
,
N. G.
,
Chretien
,
M.
,
Marjunath
,
P.
,
Schiller
,
P. W.
,
Yamashira
,
D.
and
Lai
,
C. H.
(
1984
).
Isolation, structure, and synthesis of a human seminal plasmis peptide with inhibin-like activity
.
Science
223
,
1199
1202
.
Rusewuk
,
F.
,
Schuermann
,
M.
,
Wagenaar
,
E.
,
Parren
,
P.
, WEIGEL. D. and
Nusse
,
R.
(
1987
).
The Drosophila homolog of the mouse mammary oncogene int-1 is identical to the segment polarity gene wingless
.
Cell
50
,
649
657
.
Roberts
,
A. B.
,
Anzano
,
M. A.
,
Lamb
,
L. C.
and
Sporn
,
M. B.
(
1981
).
New class of transforming growth factors potentiated by epidermal growth factor: isolation from non-neoplastic tissues
.
Proc. natn. Acad. Sci. U.S.A
.
78
,
5339
5343
.
Roberts
,
A. B.
,
Anzano
,
M. A.
,
Wakefield
,
L. M.
,
Roche
,
N. S.
,
Stern
,
D. F.
and
Sporn
,
M. B.
(
1985
).
Type β transforming growth factor: A bifunctional regulator of cellular growth
.
Proc, natn. Acad. Sci. U.S.A
.
82
,
119
123
.
Rosa
,
F.
,
Roberts
,
A. B.
,
Danielpour
,
D.
,
Dart
,
L. L.
,
Sporn
,
M. B.
and
Dawid
,
I. B.
(
1987
).
Mesoderm induction in Amphibians: The role of TGF-β2-like factors
.
Science
239
,
783
785
.
Schubiger
,
G.
(
1974
).
Acquisition of differentiative competence in the imaginai leg disc of Drosophila. Wilhelm Roux’ Arch
.
EntwMech. Org
.
174
,
303
311
.
Segal
,
D.
and
Gelbart
,
W. M.
(
1985
).
Shortvein, A new component of the decapentaplegic gene complex in Drosophila melanogaster
.
Genetics
109
,
119
143
.
Slack
,
J. M. W.
,
Darlington
,
B. G.
,
Heath
,
J. K.
and
Godsave
,
S. F.
(
1987
).
Mesoderm induction in early Xenopus embryos by heparin-binding growth factors
Nature, Lond
.
326
,
197
200
.
Spencer
,
F. A..
Hoffmann
,
F. M.
and
Gelbart
,
W. M.
(
1982
).
Decapentaplegic: a gene complex affecting morphogenesis in Drosophila melanogaster
.
Cell
28
,
451
461
.
St Johnston.
R. D.
and GELBART. W. M
. (
1987
).
decapentaplegic transcripts are localized along the dorsal-ventral axis of the Drosophila embryo
.
EMBO J
.
6
,
2785
2791
.
St Johnston.
R. D.
,
Hoffmann
,
F. M.
,
Blackman.
R. K.
,
Segal.
D
,
Grimaila.
R
,
Padgett
,
R. W.
,
Irick
,
H. A.
and
Gelbart.
W. M.
(
1990
).
The molecular organization of the decapentaplegic gene in Drosophila melanogaster
.
Genes & Devel. (in press)
.
Suter
,
B.
,
Dorig.
R. E.
,
Hoffmann
,
F M.
,
Gelbart
,
W. M.
and
Kubli
,
E.
(
1990
).
A tRNA gene located in the decapentaplegic gene of Drosophila melanogaster is expressed stage specifically
.
Molec. Cell Biol, (submitted)
.
Vale
,
W.
Rivier.
J.
,
Vaughan
,
J.
,
Mcclintock
,
R.
,
Corrigan
,
A
,
Woo.
W
.
Karr.
D
and
Spiess.
J
(
1986
).
Purification and characterization of an FSH releasing protein from porcine ovarian follicular fluid
.
Nature. Lond
.
321
,
776
779
.
Wang
,
E. A.
,
Rosen.
V.
,
Cordes.
P.
Hewick
,
R. M.
Kriz.
M.-J.
,
Luxenberg.
D. P.
,
Sibley
,
B. S.
and
Wozney
,
J. M.
(
1988
)
Purification and characterization of novel bone-inducing factors
.
Proc. natn. Acad. Sci. U S.A
.
85
,
9484
9488
.
Watson
,
M. E. E.
(
1984
).
Compilation of published signal sequences
.
Nucl. Acids Res
.
12
.
5145
5163
.
Weeks
,
D. L.
and
Melton
,
D. A.
(
1987
).
A maternal mRNA localized to the vegetal hemisphere in Xenopus eggs codes for a growth factor related to TGF-/3
.
Cell
51
,
861
867
.
WOZNEY.
J. M.
,
ROSEN.
V.
,
Celeste.
A. J.
,
Mitsock
,
L M.. WHITTERS. M. J.
,
Kriz.
R. W.
,
Hewick
,
R. M.
and
Wang.
E. A.
(
1988
).
Molecular cloning of novel regulators of bone formation
.
Science
242
.
1528
1534
.