Signalling by members of the Hedgehog family of secreted proteins plays a central role in the development of vertebrate and invertebrate embryos. In Drosophila, transduction of the Hedgehog signal is intimately associated with the activity of protein kinase A and the product of the segment polarity gene patched. We have cloned a homologue of patched from the zebrafish Danio rerio and analysed the spatiotemporal regulation of its transcription during embryonic development in both wild-type and mutant animals. We find a striking correlation between the accumulation of patched1 transcripts and cells responding to sonic hedgehog activity both in the neurectoderm and mesoderm, suggesting that like its Drosophila counterpart, patched1 is regulated by sonic hedgehog activity. Consistent with this interpretation, mis-expression of sonic hedgehog results in ectopic activation of patched1 transcription. Using dominant negative and constitutively active forms of the protein kinase A subunits, we also show that expression of patched1 as well as of other sonic hedgehog targets, is regulated by protein kinase A activity. Taken together, our findings suggest that the mechanism of signalling by Hedgehog family proteins has been highly conserved during evolution.

In Drosophila, cell patterning is controlled by the segment polarity genes, a molecularly heterogeneous group that includes the components of two signal transduction pathways (Klingensmith and Nusse, 1994; Ingham, 1995). One of these pathways is responsible for transducing the activity of the secreted protein Hedgehog, itself encoded by a member of the segment polarity class (Lee et al., 1992; Mohler and Vani, 1992; Tabata et al., 1992; Tabata and Kornberg, 1994). Several genes closely related to hedgehog (hh) have now been described in various vertebrate species, the best characterised of these being sonic hedgehog (shh) (Echelard et al., 1993; Krauss et al., 1993; Riddle et al., 1993; Roelink et al., 1994). The spatiotemporal deployment of shh is highly conserved from fish to mouse: expression is initiated during gastrulation in the embryonic shield or node and persists in the midline mesoderm, the notochord precursor, as the main body axis extends. Subsequently, expression of shh is activated in the floorplate cells of the ventral neural tube that overlie the notochord, where it continues to be expressed throughout somitogenesis (reviewed by Fietz et al., 1994).

Experimental manipulations have identified both the notochord and floorplate as sources of inducing activities that control the patterning of the neural tube and somites (van Straaten et al., 1989; Placzek et al., 1990, 1991; Yamada et al., 1991; Pourquié et al., 1993; Goulding et al., 1994) and several lines of evidence suggest that shh is a major component of such activities. First, in the zebrafish, loss of midline signalling in various mutants is closely correlated with loss of shh expression (Krauss et al., 1993; Ekker et al., 1995; Macdonald et al., 1995). Second, ectopic expression of shh in zebrafish (Krauss et al., 1993; Barth and Wilson, 1995; Macdonald et al., 1995), as well as in mouse and frog (Echelard et al., 1993; Ruiz i Altaba et al., 1995), leads to the inappropriate expression of floorplate and/or ventral brain markers. Such ectopic expression has also been shown to result in the inappropriate activation of the sclerotomal and myotomal markers pax1 and myoD, respectively, in the developing somites of the chick (Johnson et al., 1994) as well as of myoD in paraxial mesoderm of the fish (Weinberg et al., 1996). Finally, and most definitively, recombinant Shh protein is itself capable of inducing floorplate and motor neuron differentiation in neural plate explants (Marti et al., 1995; Roelink et al., 1995) and sclerotomal or myotomal differentiation in explants of presomitic mesoderm (Fan et al., 1995; Munsterberg et al., 1995).

At later stages of development, transcription of shh is activated in the posterior mesenchyme of the developing limb buds, a region corresponding to the signalling centre known as the zone of polarising activity (ZPA). Mis-expression of shh in the anterior of the limb bud results in digit duplications similar to those induced by heterotopic grafts of ZPA material (Riddle et al., 1993; Chang et al., 1994), suggesting that shh mediates the signalling activity of this region of the vertebrate limb.

Although the importance of shh during vertebrate embryogenesis is thus well established, little is known about the way in which the activity of the Shh protein is transduced. In Drosophila, by contrast, genetic analysis has identified a number of putative components of the hh-signalling pathway, such as the cAMP-dependent protein kinase (PKA) (Jiang and Struhl, 1995; Lepage et al., 1995; Li et al., 1995; Pan and Rubin, 1995; Strutt et al., 1995) as well as the products of several segment polarity genes, including patched (ptc) (Hidalgo and Ingham, 1990; Ingham et al., 1991; Ingham and Hidalgo, 1993; Capdevila et al., 1994), a novel multipass membrane spanning protein (Hooper and Scott, 1989; Nakano et al., 1989). Inactivation of either Ptc or PKA in the developing limbs of Drosophila has remarkably similar consequences; in both cases, various hh target genes are inappropriately activated, suggesting that Ptc and PKA normally act to suppress the Hh response pathway. Whilst PKA most likely acts by phosphorylating other components of this pathway, the molecular mechanism of Ptc activity has remained enigmatic. One suggestion, based upon its predicted topology and membrane localisation, is that it functions as a receptor for Hh, binding of the latter antagonising Ptc activity and thus relieving the repression of different target genes (Ingham et al., 1991). Although consistent with the spatial deployment of the two proteins and the interactions between them as deduced by genetic analysis, this model has so far received no direct biochemical support.

Whatever the identity of the Hh receptor, it seems likely to have been highly conserved through evolution since vertebrate hh homologues are capable of activating the response pathway when expressed in transgenic Drosophila (Krauss et al., 1993; Chang et al., 1994; Ingham and Fietz, 1995). Moreover, the discovery that the response of tissue explants to Shh activity can be attenuated by drugs that activate PKA (Fan et al., 1995; Hynes et al., 1995) suggests that at least some of the intracellular components of the pathway may have also been conserved. In this study, we have explored the extent of this conservation further by cloning a zebrafish homologue of ptc and analysing its expression during embryogenesis. In Drosophila, ptc is itself one of the principal targets of hh activity, its transcription being upregulated in hh-responding cells. Using midline mutants and overexpression strategies, we have investigated the relationship between shh activity and transcription of the zebrafish ptc1 gene. Our results support a role for shh and ptc1 in the specification of both neural and mesodermal cell fates in the zebrafish; in addition, we show that both aspects of shh signalling are mediated by PKA activity.

In situ hybridisation

In situ hybridisation was performed as described by Oxtoby and Jowett (1993) with the following modifications: hybridisation were performed at 70°C and post-hybridisation washes were as suggested by Henrique et al. (1995). Stained embryos were dehydrated through an ethanol: butanol series, embedded in Fibrowax and sectioned (610 μm). Double stainings were performed essentially as described (Jowett and Lettice, 1994).

Probes used for in situ hybridisations were synthesized using the following templates: shh, (Krauss et al., 1993); pax-[b] (Krauss et al., 1991); nk2.2, (Barth and Wilson, 1995); MyoD (Weinberg et al., 1996). Unincorporated DIG-UTP was removed by centrifugation through a Nuc50 column (Kodak). Specimens were analysed using a Zeiss Axioplan microscope and photographed with Kodak Ektachrome 64T film. Images were scanned on a Sprintscan 35 slidescanner and processed using Adobe Photoshop software.

zebrafish stocks

Wild-type Danio rerio were bred from a founder population obtained from the Goldfish Bowl, Oxford. The flhn1 strain was obtained from T. Jowett (Newcastle University, UK). The cyc and ntl strains were obtained from C. Kimmel (University of Oregon, USA). Fish were maintained in a constant recirculating system at 28°C on a 14 hours light/10 hours dark cycle.

Cloning and sequence analysis

Two pairs of primers (rev2, rev4 and genie1, genie2, see below) were used separately to perform PCR starting from random-primed cDNA samples synthesised using either bud stage or somitogenesis stage zebrafish RNA.

Rev2: ggacgaattcTSYTCNRGCCARTGCAT

Rev4: ggacgaattcYTNGANTGYTTYTGGGA

Genie1: ggacgaattccGAYGGNATNATNAAYC

Genie2: ggacgaattcRTAYTGYTCCCARAANA

(large lettering corresponds to sequences conserved between mouse and different insect ptc sequences, small lettering corresponds to an EcoRI site and arbitrary flanking sequence introduced to facilitate subsequent cloning of PCR products). Amplification was performed using 35 cycles of 94°C 30 seconds; 50°C 30 seconds; 72°C 90 seconds. Fragments of the expected size were subcloned into Bluescript KS and sequenced. One fragment was obtained with the Genie1 and Genie2 primers (called Genie) and two with the Rev2 and Rev4 primers (called Rev1 and Rev3). Two primers identical to Rev1 and Rev3 fragments, respectively, a primer complementary to the Genie fragment and the Genie2 primer were used in different pairwise combinations to perform PCR starting from the same original cDNA samples. Two fragments of 1.8 and 2 kb were amplified using the Rev1and Genie-derived primers and the Rev3-derived and Genie2 primers respectively, which after sequencing proved to be homologous to both mouse and Drosophila ptc.

2×106 plaques of a λgt11 library made from 33-hour-old zebrafish embryos (K. Zinn) were screened at low stringency according to standard procedures using the 1.8 and 2 kb fragments as probes. We did not obtain any positive clones with the 2 kb probe using either the 33 hour library or libraries from other stages. 20 positive plaques from the screen using the 1.8 kbp fragment were purified and analysed by PCR using oligonucleotides against the 1.8 kb probe and against the lambda vector. The longest clone, λ105, was sequenced on both strands using either Pharmacia or Applied Biosysytems automated sequencers with primers from the cloning vector and internal primers. Sequence analysis was performed using Geneworks (Intelligenetics) and GCG software packages.

The nucleotide sequence of the zebrafish ptc1 gene described here is deposited in the EMBL database under the accession number: X98883.

Plasmids

The open reading frames of a dominant negative mutant cDNA of the regulatory subunit of Drosophila melanogaster cAMP-dependent protein kinase (Li et al., 1995) and of a constitutively active subunit of the mouse PKA catalytic subunit (Orellana and McKnight, 1992) were subcloned into the CS2 expression plasmids where cDNA transcription is driven by the ubiquitous CMV enhancer/promoter region (Turner and Weintraub, 1994) to create pCS2dnPKA and pCS2cPKA, respectively. The open reading frame of shh was subcloned into the CDM8 expression vector (Seed, 1987).

Embryo injections

Injections were performed on 2to 4-cell-stage embryos using backfilled capillaries (Clarks Microelectrical Instruments, Reading) and a pressure-pulsed Narishige microinjector. RNA, synthesized in vitro from linearized plasmids p64Tshh (Krauss et al., 1993), pCS2dn PKA and pCS2cPKA, respectively, was injected at empirically determined concentrations which were of approximately 100 μg/ml. Plasmid DNA, purified using Qiagen columns, was injected at a concentration of approximately 100 μg/ml.

Isolating sequences homologous to Drosophila patched from zebrafish

Vertebrate homologues of ptc were first identified in mouse using a polymerase chain reaction (PCR) ptc homologues from distantly related insect species (Goodrich et al., 1996). Two pairs of degenerate oligonucleotide primers based upon sequences conserved between the mouse and insect genes were used to amplify related sequences from zebrafish cDNA (see Materials and Methods). Two fragments of approximately 320 bp and one fragment of around 330 bp, whose sequences proved to be related to both mouse and insect ptc sequences, were amplified. Oligonucleotides based on sequences within these three PCR fragments were then used to amplify two fragments of approximately 1.8 kb and 2.0 kb and each of these fragments was used as a probe to screen a 33-hour-old embryonic zebrafish cDNA library. One full-length cDNA, designated ptc1, was isolated and is described in this paper. Similar clones corresponding to the other ptc homologue, ptc2, have yet to be isolated and this gene will not be considered further here.

The deduced amino acid sequence of ptc1 shows 64% identity to both the mouse and chicken ptc genes and 39% identity to Drosophila ptc. Comparison of the different Ptc sequences reveals that eight cysteine residues are conserved in all four species. The Drosophila protein has seven potential Nglycosylation sites of which three are conserved in mouse and chick but only one is present in zebrafish. Hydropathy analysis predicts that each protein contains 12 potential transmembrane domains (see Fig. 1A). There are two regions in the amino acid sequences that are particularly divergent; in the first of these, the vertebrate sequences have insertions relative to the fly sequence, between the sixth and seventh putative transmembrane domains. This region is also relatively divergent between zebrafish and mouse although it is well conserved between chick and mouse (Marigo et al., 1996). The second significant divergence occurs towards the C-terminus, which in zebrafish is considerably shorter than in either Drosophila or mouse; the region absent in zebrafish shows little conservation between the latter two species (see Fig. 1A). Database searches have also revealed homology to a ptc-related sequence identified by the C. elegans genome project (Wilson et al., 1994); the zebrafish ptc1 gene is as similar to this gene as it is to Drosophila ptc (Fig. 1B).

Fig. 1.

Comparison of the amino acid sequences of ptc genes from different organisms. (A) Sequence alignment of the Drosophila and mouse ptc proteins with zebrafish ptc1. Putative transmembrane domains are indicated by black lines above the sequence. Orange shading indicates identities between all three species. Blue shading indicates identities between the two vertebrate proteins. Putative N-glycosylation sites in Drosophila are indicated by pink lines and conserved cysteine residues by asterisks. Zebrafish Ptc1 is significantly shorter at the C terminus than Ptc proteins from other species. We checked the sequence of the cDNA in the region of the stop codon and the position of the stop codon, by amplifying this region from cDNA prepared from zebrafish embryos at somitogenesis stages. (B) Percentage identity and similarity between patched genes from different organisms. Comparisons were done using ‘GCG Bestfit’. The first number is the percentage identity and the number in brackets is the percentage similarity

Fig. 1.

Comparison of the amino acid sequences of ptc genes from different organisms. (A) Sequence alignment of the Drosophila and mouse ptc proteins with zebrafish ptc1. Putative transmembrane domains are indicated by black lines above the sequence. Orange shading indicates identities between all three species. Blue shading indicates identities between the two vertebrate proteins. Putative N-glycosylation sites in Drosophila are indicated by pink lines and conserved cysteine residues by asterisks. Zebrafish Ptc1 is significantly shorter at the C terminus than Ptc proteins from other species. We checked the sequence of the cDNA in the region of the stop codon and the position of the stop codon, by amplifying this region from cDNA prepared from zebrafish embryos at somitogenesis stages. (B) Percentage identity and similarity between patched genes from different organisms. Comparisons were done using ‘GCG Bestfit’. The first number is the percentage identity and the number in brackets is the percentage similarity

Spatial expression pattern of ptc1 during zebrafish embryogenesis

The expression pattern of ptc1 during zebrafish embryogenesis was analysed by in situ hybridisation and compared to that of shh. Transcription of ptc1 is first apparent at around 70% epiboly in the presumptive mesodermal cell layer in two stripes of cells flanking the axial mesoderm, which at this stage already shows robust expression of shh. As the convergence extension movements of gastrulation proceed, low-level ptc1 expression is apparent throughout the presomitic mesoderm while the two stripes of high-level expression, corresponding to the so-called adaxial cells (Thisse et al., 1993), extend along the entire axis; in addition, high levels of ptc1 transcript are detectable in the neurectoderm overlying the axial mesoderm and in the ventral part of the future brain (Fig. 2A). With the onset of somitogenesis, high-level expression persists in the tissues surrounding the notochord in the trunk as well as in the ventral neuroectoderm of the brain (Fig. 2C). Distinct patches of high-level mesodermal expression become apparent adjacent to the notochord as the somites form (Fig. 2M), a pattern similar to that of α-tropomyosin at the same stage (Thisse et al., 1993). During later stages of somitogenesis when shh expression is lost from the notochord, expression of ptc1 persists throughout the ventral neural tube, except in the floorplate, which at this stage still expresses shh at high levels. Notably, expression in the mesoderm also persists lateral and ventral to the notochord (Fig. 2G,M-P). In the brain, ptc1 expression extends dorsally in the diencephalon, paralleling the dorsal extension of shh expression in the same region and shows a modulation in the hindbrain, which becomes more defined at later stages (Fig. 2G; see also Fig. 5C).

Fig. 2.

Comparison of the ptc1 and shh expression patterns during wild-type embryogenesis. Transcripts were revealed by in situ hybridization with antisense RNA probes to ptc1 (A,C,E,G,I,K,M,O,Q,S,U) or shh (B,D,F,H,J,L,N,P,R,T,V). (A,B) Dorsal views of embryos at the end of gastrulation. The arrowhead indicates a row of hypoblast cells adjacent to the axial mesoderm which express ptc1 at high levels. The asterisk indicates the anterior boundary of the ptc1 low-level expression domain in the presomitic mesoderm. (C,D) Lateral views of 10-somite (C) and 14somite (D) stage embryos. Arrowheads denote ptc1 expression in somites. Asterisks delimit the expression of ptc1 in the neurectoderm all along the anteroposterior axis of the embryo. (E,F) 18-somite stage embryos. (G,H) 26-somite stage. Asterisks indicate the dorsal extension of ptc1 and shh expression domains in the diencephalon. Brackets indicate upregulation of ptc1 in rhombomeres 2, 4 and 6, a modulation of expression that is unrelated to shh expression (see also Fig. 5c). (I,J) Embryos at 48 hours of development. Asterisks indicate the position of the floor plate of the neural tube. Arrowheads indicate expression in the gut. golden mutant embryos, which are defective in pigmentation, were used to facilitate visualisation of ptc1 and shh expression. (K,L) 8-somite stage embryos. Dorsal view of flat preparations showing the complementarity between the paraxial expression of ptc1 and the axial expression of shh in the notochord and the tailbud. (M,N) Transverse sections through the trunk of embryos at 24 hours of development. (N) Sections at two different levels along the axis, the section on the right-hand side being more posterior. (O,P) Horizontal sections through the trunk of embryos at 24 hours of development. (Q,R) Dorsal view of the trunk of 36-hour-old embryos showing the expression of ptc1 and shh in the posterior region of the developing fins. (S,T) Detail of embryos shown in I and J, respectively. Asterisks indicate the position of the floor plate of the neural tube. Arrowheads indicate expression in the foregut. Expression in the hindbrain, when observed dorsally, was found to be restricted to periventricular cells in the midline (data not shown). (U,V) Frontal views of embryos shown in I and J, respectively. Arrowheads indicate expression in the foregut surrounding the developing mouth. Abbreviations : e, eye; fp, floor plate; hb, hindbrain; n, notochord; te, tectum; t, telencephalon; tb, tail bud; y, yolk.

Fig. 2.

Comparison of the ptc1 and shh expression patterns during wild-type embryogenesis. Transcripts were revealed by in situ hybridization with antisense RNA probes to ptc1 (A,C,E,G,I,K,M,O,Q,S,U) or shh (B,D,F,H,J,L,N,P,R,T,V). (A,B) Dorsal views of embryos at the end of gastrulation. The arrowhead indicates a row of hypoblast cells adjacent to the axial mesoderm which express ptc1 at high levels. The asterisk indicates the anterior boundary of the ptc1 low-level expression domain in the presomitic mesoderm. (C,D) Lateral views of 10-somite (C) and 14somite (D) stage embryos. Arrowheads denote ptc1 expression in somites. Asterisks delimit the expression of ptc1 in the neurectoderm all along the anteroposterior axis of the embryo. (E,F) 18-somite stage embryos. (G,H) 26-somite stage. Asterisks indicate the dorsal extension of ptc1 and shh expression domains in the diencephalon. Brackets indicate upregulation of ptc1 in rhombomeres 2, 4 and 6, a modulation of expression that is unrelated to shh expression (see also Fig. 5c). (I,J) Embryos at 48 hours of development. Asterisks indicate the position of the floor plate of the neural tube. Arrowheads indicate expression in the gut. golden mutant embryos, which are defective in pigmentation, were used to facilitate visualisation of ptc1 and shh expression. (K,L) 8-somite stage embryos. Dorsal view of flat preparations showing the complementarity between the paraxial expression of ptc1 and the axial expression of shh in the notochord and the tailbud. (M,N) Transverse sections through the trunk of embryos at 24 hours of development. (N) Sections at two different levels along the axis, the section on the right-hand side being more posterior. (O,P) Horizontal sections through the trunk of embryos at 24 hours of development. (Q,R) Dorsal view of the trunk of 36-hour-old embryos showing the expression of ptc1 and shh in the posterior region of the developing fins. (S,T) Detail of embryos shown in I and J, respectively. Asterisks indicate the position of the floor plate of the neural tube. Arrowheads indicate expression in the foregut. Expression in the hindbrain, when observed dorsally, was found to be restricted to periventricular cells in the midline (data not shown). (U,V) Frontal views of embryos shown in I and J, respectively. Arrowheads indicate expression in the foregut surrounding the developing mouth. Abbreviations : e, eye; fp, floor plate; hb, hindbrain; n, notochord; te, tectum; t, telencephalon; tb, tail bud; y, yolk.

The relationship between the ptc1 and shh expression domains in the developing brain was analysed directly using double-labelling techniques to visualise both transcripts simultaneously. At all stages, high-level expression of ptc1 occurs in a domain delimiting the expression of shh. By 24 hours, expression of shh no longer occupies the most ventral part of the rostral forebrain and ptc1 expression can now be detected ventrally to shh-expressing cells (Fig. 3A). High-level expression of ptc1 is also detected around the lumen of the neural tube.

Fig. 3.

Simultaneous localisation of ptc1 and shh transcripts in the head. Whole embryos were hybridized simultaneously with riboprobes to ptc1 and shh revealing ptc1 expression in red staining and that of shh in blue. Lateral views are shown with anterior to the left; yolk and eyes were removed to improve brain visualisation. (A) Embryo at 36 hours of development. (B) Embryo at 48 hours of development. Abbreviations: fp, floor plate; hb, hindbrain; hy, hypothalamus; t, telencephalon.

Fig. 3.

Simultaneous localisation of ptc1 and shh transcripts in the head. Whole embryos were hybridized simultaneously with riboprobes to ptc1 and shh revealing ptc1 expression in red staining and that of shh in blue. Lateral views are shown with anterior to the left; yolk and eyes were removed to improve brain visualisation. (A) Embryo at 36 hours of development. (B) Embryo at 48 hours of development. Abbreviations: fp, floor plate; hb, hindbrain; hy, hypothalamus; t, telencephalon.

By 36 hours, additional sites of ptc1 transcription appear in the first branchial arch (not shown) and the posterior mesenchyme of the fin buds (Fig. 2Q), in both cases close to domains of shh expression (Fig. 2R; see also Krauss et al., 1993). By 48 hours, both genes exhibit a complex expression pattern in the brain with a persistence of high-level ptc1 expression adjacent to shh-expressing cells. High-level expression of ptc1, is, however, also observed some distance away from cells expressing shh in a number of places, including most notably a domain of intense expression in the hindbrain (Figs 2S,T, 3B). Expression of both genes is now obvious in the foregut (Fig. 2S-V) as well as in more posterior domains.

Disruption of midline signalling affects both neural and mesodermal expression of ptc1

The relationship between the patterns of expression of shh and ptc1 described above is highly reminiscent of that between their Drosophila counterparts (Taylor et al., 1993; Tabata and Kornberg, 1994) suggesting that, as in Drosophila, transcription of ptc1 may be induced in response to hh signalling. Since no mutation of shh is currently available in the zebrafish, we took advantage of a number of mutants that eliminate shh expression at different levels along the anteroposterior axis to investigate this possibility.

Mutation of the cyclops (cyc) gene disrupts the specification of the prechordal plate mesoderm (Thisse et al., 1994) and concomitantly the induction of the overlying neurectoderm, resulting in defective midline signalling in the developing brain which gives rise to the cyclopic phenotype (Hatta et al., 1994; Ekker et al., 1995; Macdonald et al., 1995). In addition, cyc mutants lack a floorplate, apparently due to a defect in the response of cells to the inductive signal from the underlying notochord (Hatta et al., 1991). In line with these phenotypic effects, expression of shh is completely absent from both the prechordal plate mesoderm and the overlying neurectoderm that gives rise to the ventral floor of the brain (Krauss et al., 1993) while, posterior to the midbrain, expression is normal in the axial mesoderm but absent from the ventral neural tube.

Expression of ptc1 in cyc embryos mirrors these changes in shh expression precisely: thus by the time the body axis is fully extended, no ptc1 transcripts are detectable anterior to the domain of expression of pax[b] (Fig. 4A), which marks the future midbrain (Krauss et al., 1991), and the forebrain and midbrain remain devoid of ptc1 expression throughout the rest of embryogenesis (Fig. 4B). The only exception is a short dorsoventral stripe in the abnormal neural fold of the diencephalon that appears at around 24 hours of development surrounding a stripe of shh-expressing cells that appears at the same stage (Fig. 4B-D). Posterior to the midbrain/hindbrain boundary, ptc1 is expressed normally in the somites and ventral neural tube (Fig. 4B), except that it persists along the ventral midline of the latter, presumably reflecting the failure of the floorplate to differentiate.

Fig. 4.

Expression of ptc1 and shh in cyclops mutants. ptc1 expression is detected in blue (A,B) or in red (C,D) and shh in blue (C,D). (A) Frontal view of embryos at the 1to 2-somite stages. Expression of ptc1 is shown in a wild-type sib on the left and in a cyclops mutant on the right. To provide a landmark along the anteroposterior axis, embryos were simultaneously hybridized with a probe for pax-2 highlighting the position of the future posterior midbrain (asterisk). (B) ptc1 expression in a 24-hour-old cyclops mutant. The arrowhead indicates re-expression of ptc1 at high levels in a dorsoventral stripe in the diencephalon. (C) 24-hour-old cyclops embryos. ptc1 expression is revealed in red, shh in dark blue. (D) 24hour-old wild-type embryo.

Fig. 4.

Expression of ptc1 and shh in cyclops mutants. ptc1 expression is detected in blue (A,B) or in red (C,D) and shh in blue (C,D). (A) Frontal view of embryos at the 1to 2-somite stages. Expression of ptc1 is shown in a wild-type sib on the left and in a cyclops mutant on the right. To provide a landmark along the anteroposterior axis, embryos were simultaneously hybridized with a probe for pax-2 highlighting the position of the future posterior midbrain (asterisk). (B) ptc1 expression in a 24-hour-old cyclops mutant. The arrowhead indicates re-expression of ptc1 at high levels in a dorsoventral stripe in the diencephalon. (C) 24-hour-old cyclops embryos. ptc1 expression is revealed in red, shh in dark blue. (D) 24hour-old wild-type embryo.

Fig. 5.

Comparison of ptc1 and shh expression in floating head and no tail mutant embryos. Expression of ptc1 (A,C,E) and shh (B,D,F) in flh mutant embryos. (A,B) Embryos at the 5-somite stage. (C,D) Embryos at 24 hours of development. Brackets and bars indicate modulation of ptc1 in different rhombomeres which is unrelated to shh expression. (E,F) Lateral view of the trunk of embryos shown in C,D. Expression of ptc1 and shh in ntl mutant embryos. (G) ntl mutant embryos at 24 hours of development. ptc1 expression is revealed in red, shh in blue. (H) Lateral view of the trunk of a 24-hour-old ntl embryo in which ptc1 expression is revealed in blue.

Fig. 5.

Comparison of ptc1 and shh expression in floating head and no tail mutant embryos. Expression of ptc1 (A,C,E) and shh (B,D,F) in flh mutant embryos. (A,B) Embryos at the 5-somite stage. (C,D) Embryos at 24 hours of development. Brackets and bars indicate modulation of ptc1 in different rhombomeres which is unrelated to shh expression. (E,F) Lateral view of the trunk of embryos shown in C,D. Expression of ptc1 and shh in ntl mutant embryos. (G) ntl mutant embryos at 24 hours of development. ptc1 expression is revealed in red, shh in blue. (H) Lateral view of the trunk of a 24-hour-old ntl embryo in which ptc1 expression is revealed in blue.

In contrast to cyc, mutation of the floating head (flh) gene has no effect on prechordal plate mesoderm but instead disrupts notochord specification; this leads to a premature loss of shh expression from the axial mesoderm posterior to the midbrain-hindbrain boundary and the resultant disruption of floor plate induction is reflected in patchy expression of shh along the ventral midline of the neural tube (Talbot et al., 1995; see Fig. 5D). In addition, muscle pioneer cells that derive from the adaxial cells adjacent to the notochord (Felsenfeld et al., 1991; Thisse et al., 1993) fail to differentiate (Halpern et al., 1995; Talbot et al., 1995). A similar effect on muscle pioneer differentiation is caused by mutation of the notail (ntl) gene, which also lacks a notochord. Unlike flh, however, ntl mutants do not lack the floorplate (Halpern et al., 1993). These differences in phenotype seem to reflect differences in the effects of each mutation on shh expression, which persists longer in the axial midline of ntl embryos (Krauss et al., 1993) than in that of flh embryos (Talbot et al., 1995).

In flh embryos at 24 hours, expression of ptc1, like that of shh, is completely normal in the brain and anterior neural tube (Fig. 5C,D). More caudally, however, expression is restricted to small clusters of cells distributed sporadically along the ventral neural tube in a manner reflecting the intermittent expression of shh (Fig. 5; compare C to D and E to F). In addition to the effects on neural expression, the broad stripes of ptc1 expression typical of wild-type somites are absent from flh homozygotes, with only scattered mesodermal cells expressing the gene in close association with expressing cells in the neural tube (Fig. 5E). Thus expression of ptc1 in the mesoderm, as in the neural tube, seems to depend upon proximity to a source of shh activity. To investigate this relationship further, we analysed the expression of ptc1 at earlier stages of development in flh embryos. At the bud stage, shh is still expressed along the axial midline in flh embryos, in cells of uncertain origin (Fig. 5B); at the same stage, ptc1 is expressed in a single stripe of cells along the midline instead of in the two stripes typical of wild type (Fig. 5A). A similar change in the early pattern of ptc1 expression is seen in ntl homozygotes (data not shown). However, at 24 hours, in ntl mutants high-level expression persists in distinct stripes in each somite adjacent to the floorplate that expresses shh (Fig. 5G,H).

Induction of ptc1 transcription by ectopic shh activity mirrors its effects on neural and adaxial-specific genes

The analysis of ptc1 expression in midline mutations suggests that expression of shh is necessary for the induction of high-level ptc1 transcription. To test whether shh is sufficient to induce such transcription, we next examined the effects of ectopic shh activity on ptc1 expression. Previous studies have shown that injection of synthetic shh mRNA into the 2to 4-cell-stage embryo results in the ectopic activation of genes such as axial and nk2.2 whose expression domains are normally restricted to the ventral region of the neural tube and brain (Krauss et al., 1993; Barth and Wilson, 1995). In the brain, the distribution of ptc1 transcripts around the shh expression domain closely mirrors the expression domain of nk2.2 (Fig. 6A,B). To determine whether ptc1 expression can similarly be ectopically induced by shh in the nervous system, embryos from the same cohort were injected with shh mRNA, fixed after 27 hours of development and hybridised with probes for ptc1 and nk2.2 As expected, nk2.2 is expressed ectopically in dorsal and lateral regions of the brain as well as in the eye rudiment of such embryos (Fig. 6D). Similarly, high-level expression of ptc1 is also detected ectopically in the diencephalon and midbrain of shh-injected embryos as well as in the eye rudiments (Fig. 6C). Notably, ectopic expression of ptc1 was never detected in the telencephalon. Similar effects on ptc1 expression were observed in small groups of cells following injection of a shh expression plasmid (data not shown) supporting the interpretation that the induction of transcription is a direct consequence of shh activity.

Fig. 6.

Induction of ptc1, myoD andnk2.2 expression by shh in the brain and early somitogenesis stage embryos. (A,C) Lateral views of ptc1 expression in the brain of 27-hour-old embryos. (A) Uninjected control embryo. (C) shh-injected embryo. (B,D) Lateral views of nk2.2 expression in the brain of 27-hour-old embryos. (B) Uninjected control embryo. (D) shh-injected embryo. Arrowheads in C, D show the telencephalon which is devoid of ectopic ptc1 and nk2.2 expression. (E,G) Dorsal views of ptc1 expression in flat preparations of embryos at early somitogenesis. (E) Uninjected embryo. (G) shh-injected embryos. No ectopic ptc1 expression was detected in the ectoderm. (F,H) Dorsal views of myoD expression in flat preparations of embryos at early somitogenesis. (F) Uninjected embryo. (H) shh-injected embryo.

Fig. 6.

Induction of ptc1, myoD andnk2.2 expression by shh in the brain and early somitogenesis stage embryos. (A,C) Lateral views of ptc1 expression in the brain of 27-hour-old embryos. (A) Uninjected control embryo. (C) shh-injected embryo. (B,D) Lateral views of nk2.2 expression in the brain of 27-hour-old embryos. (B) Uninjected control embryo. (D) shh-injected embryo. Arrowheads in C, D show the telencephalon which is devoid of ectopic ptc1 and nk2.2 expression. (E,G) Dorsal views of ptc1 expression in flat preparations of embryos at early somitogenesis. (E) Uninjected embryo. (G) shh-injected embryos. No ectopic ptc1 expression was detected in the ectoderm. (F,H) Dorsal views of myoD expression in flat preparations of embryos at early somitogenesis. (F) Uninjected embryo. (H) shh-injected embryo.

To examine the effects of ectopic shh activity in the mesoderm, embryos from the same injection cohorts were fixed at the onset of somitogenesis and hybridised with probes for myoD and ptc1. Like ptc1, expression of myoD is restricted to the adaxial cells in the presomitic mesoderm (Fig. 6F) and recent studies have suggested that its expression is regulated by shh (Weinberg et al., 1996). In a significant proportion of injected embryos, we found that the expression domains of both ptc1 (64%; n=112) and myoD (71%; n=90), are expanded so that they occupy most of the lateral mesoderm on one or both sides of the midline (Fig. 6G,H). In addition, ptc1 is ectopically expressed throughout much of the neurectoderm at this stage following shh injection (Fig. 6G).

Transcription of ptc1 is regulated by protein kinase A

In Drosophila, removal of PKA activity from cells mimics their response to hh signalling, activating the transcription of ptc and other effectors of hh activity, suggesting that PKA normally acts to repress expression of hh target genes (Jiang and Struhl, 1995; Lepage et al., 1995; Li et al., 1995; Pan and Rubin, 1995; Strutt et al., 1995). Since the relationship between hh signalling and ptc expression appears to be conserved from flies to fish, we wondered whether regulation of ptc1 and other shh targets might also depend upon PKA activity. PKA exists as an inactive heterodimer comprising a catalytic and a cAMPbinding regulatory subunit; binding of cAMP to the latter dissociates the heterodimer allowing the catalytic subunit to phosphorylate its target substrates. To modulate PKA activity, we made use of two previously described dominant mutations of PKA subunits. In the first of these, the cAMP-binding sites in the regulatory subunit are mutated such that it cannot be dissociated from the catalytic subunit (Li et al., 1995). Expression of this mutated subunit, which we refer to as dnPKA, thus results in the inactivation of the catalytic subunit and hence in the reduction or loss of PKA activity. In the second case, the catalytic subunit is mutated such that it cannot bind the regulatory subunit, thus rendering it constitutively active (Orellana and McKnight, 1992); we refer to this mutant form as cPKA. Amongst embryos injected at the 2to 4-cell stage with synthetic mRNA encoding the dnPKA mutant and fixed at early somitogenesis stages, more than half (55% n=120) exhibit high-level expression of ptc1 transcription throughout the lateral mesoderm on one or both sides of the midline (Fig. 7C). In addition, a similar proportion of embryos from the same injection cohort exhibit ectopic expression of myoD either unilaterally or bilaterally (Fig. 7D). At later stages (27 hours), injected embryos express both ptc1 (Fig. 7G) and nk2.2 (Fig. 7H) at ectopic locations in the diencephalon and mesencephalon. All of these effects are similar to, though somewhat less robust than, those induced by ectopic shh expression (compare Figs 6 and 7). Similarly, embryos injected with dnPKA mRNA also consistently show ectopic activation of pax[b] in the eye, again similar to, though less extensive than, that induced by ectopic shh expression (Fig. 8A-C).

Fig. 7.

Induction of ptc1, nk2.2 and myoD expression in the brain, and in early somitogenesis stage embryos, by inhibition of PKA activity. (A,C) Dorsal views of ptc1 expression in flat preparations of embryos at early somitogenesis. (A) Uninjected embryo.(C) dominant negative PKA-injected embryos. (B,D) Dorsal views of myoD expression in flat preparations of embryos at early somitogenesis. (B) Uninjected embryo. (D) dominant negative PKAinjected embryos. (E,G) Lateral views of ptc1 expression in the brain of 27-hour-old embryos. (E) Uninjected control embryo. (G) dnPKAinjected embryo. (F,H) Lateral views of nk2.2 expression in the brain of 27-hour-old embryos. (F) Uninjected control embryo. (H) dnPKAinjected embryo.

Fig. 7.

Induction of ptc1, nk2.2 and myoD expression in the brain, and in early somitogenesis stage embryos, by inhibition of PKA activity. (A,C) Dorsal views of ptc1 expression in flat preparations of embryos at early somitogenesis. (A) Uninjected embryo.(C) dominant negative PKA-injected embryos. (B,D) Dorsal views of myoD expression in flat preparations of embryos at early somitogenesis. (B) Uninjected embryo. (D) dominant negative PKAinjected embryos. (E,G) Lateral views of ptc1 expression in the brain of 27-hour-old embryos. (E) Uninjected control embryo. (G) dnPKAinjected embryo. (F,H) Lateral views of nk2.2 expression in the brain of 27-hour-old embryos. (F) Uninjected control embryo. (H) dnPKAinjected embryo.

Fig. 8.

Expression of pax[b] in wild-type, shh-injected and dnPKAinjected embryos. Expression of pax[b] in uninjected (A), shhinjected (B) and dnPKA-injected (C) 27-hour-old embryos. Note expansion of expression into the eyes in the injected embryos in comparison to the restricted expression in the optic stalk (arrowhead) of the normal embryo.

Fig. 8.

Expression of pax[b] in wild-type, shh-injected and dnPKAinjected embryos. Expression of pax[b] in uninjected (A), shhinjected (B) and dnPKA-injected (C) 27-hour-old embryos. Note expansion of expression into the eyes in the injected embryos in comparison to the restricted expression in the optic stalk (arrowhead) of the normal embryo.

By contrast, injection of embryos with mRNA encoding the cPKA mutant has the opposite effect on the expression of hh target genes. At 27 hours, expression of ptc1 and nk2.2 is almost totally eliminated from the brain and ventral neural tube of injected embryos, with only a stripe of expression of both genes persisting in the diencephalon (Fig. 9D,E). This stripe corresponds to the dorsal extension of the normal diencephalic expression domain, which also persists in cyc mutant embryos (compare with Fig. 4). Like the latter, most cPKA-injected embryos (60%, n=90) also exhibit varying degrees of cyclopia, involving fusion of the retina alone or both the retina and lens (Fig. 9F), a further indication that midline signalling is attenuated or eliminated by the unregulated activity of PKA.

Fig. 9.

Suppression of midline signalling by cPKA activity. (A-C) normal (uninjected embryos) (D-F) embryos injected with cPKA RNA at the 2to 4-cell stage. (A,D) were hybridised with probes for both pax[b] and ptc1. In D, most of the signal derives from the pax[b] probe including the stripe at the midbrain/hindbrain boundary (arrowhead) and the reduced staining in the optic stalks (arrowhead). The asterisk indicates the stripe of ptc1 expression in the diencephalon note that in the injected embryo this extends further ventrally compared to the wild-type embryo, indicating a dorsalisation of the brain. (B,E) Hybridised with a probe for nk2.2. Note the marked loss of expression following injection of the cPKA mRNA. (C,F) Frontal views of the heads of 27hour-old embryos, showing the partial fusion of the eyes in the cPKA-injected animal.

Fig. 9.

Suppression of midline signalling by cPKA activity. (A-C) normal (uninjected embryos) (D-F) embryos injected with cPKA RNA at the 2to 4-cell stage. (A,D) were hybridised with probes for both pax[b] and ptc1. In D, most of the signal derives from the pax[b] probe including the stripe at the midbrain/hindbrain boundary (arrowhead) and the reduced staining in the optic stalks (arrowhead). The asterisk indicates the stripe of ptc1 expression in the diencephalon note that in the injected embryo this extends further ventrally compared to the wild-type embryo, indicating a dorsalisation of the brain. (B,E) Hybridised with a probe for nk2.2. Note the marked loss of expression following injection of the cPKA mRNA. (C,F) Frontal views of the heads of 27hour-old embryos, showing the partial fusion of the eyes in the cPKA-injected animal.

To explore the relationship between shh signalling and PKA further, we co-injected mRNAs encoding shh and cPKA and analysed the eyes of resultant embryos at 27 hours. Injection of shh RNA alone leads to a high proportion of embryos with rudimentary eyes (77% n=183) as previously described (Krauss et al., 1993; Ekker et al., 1995; MacDonald et al., 1995). Amongst embryos co-injected with both shh and cPKA RNAs, only 22% (n=70) exhibited such reduction of the eyes. Moreover, the frequency of cyclopia was much reduced (11%) relative to that induced by injecting cPKA alone (60%), suggesting that the activities of shh and PKA are mutually antagonistic.

ptc1 is a target of shh signalling

Signalling by Hh family proteins is used repeatedly and in various contexts during animal development. In Drosophila, signalling by hh in the embryonic segments as well as in the imaginal discs, retina and ovary is intimately associated with the activity of the segment polarity gene ptc (Ingham and Hidalgo, 1993; Ma et al., 1993; Capdevila et al., 1994; Heberlein et al., 1995; Tabata et al., 1995; Forbes et al., 1996a,b). In each instance, the two genes are expressed in complementary domains (Taylor et al., 1993; Tabata and Kornberg, 1994; Capdevila et al., 1994; Forbes et al., 1996b), the activity of ptc suppressing the transcription of hh target genes wherever it is expressed. Thus when ptc activity is eliminated, cells behave as though they have received the hh signal (Ingham and Hidalgo, 1993; Capdevila et al., 1994; Lepage et al., 1995; Tabata et al., 1995), even in the total absence of hh function (Ingham and Hidalgo, 1993). These properties have led to the notion that hh acts by antagonising the activity of ptc (Ingham et al., 1991; Ingham and Hidalgo, 1993; Basler and Struhl, 1994; Tabata and Kornberg, 1994) and, because ptc encodes a transmembrane protein, it has been suggested that it may act as a receptor for the secreted Hh protein (Ingham et al., 1991). Since the signalling activity of Hh family proteins has been highly conserved during evolution (Krauss et al., 1993; Chang et al., 1994; Ingham and Fietz, 1995), we anticipated that proteins involved in their reception at the cell surface would similarly be well conserved. The isolation of ptc homologues from zebrafish reported here, as well as from mouse (Goodrich et al., 1996) and chicken (Marigo et al., 1996), is in line with this expectation. Whether or not there is a direct interaction between these proteins and the respective Hh family proteins, however, remains a moot point. Although the membrane association of Drosophila Ptc (Taylor et al., 1993) is consistent with its hypothesised role as a receptor, the predicted topology of both the invertebrate and vertebrate Ptc proteins is more typical of that of ion channels or transporter proteins (Hooper and Scott, 1989; Nakano et al., 1989; Goodrich et al., 1996).

A defining feature of Drosophila ptc is its transcriptional up regulation in response to hh signalling. The highest levels of ptc expression are typically found in cells immediately adjacent to those expressing hh (Hooper and Scott, 1989; Nakano et al., 1989; Taylor et al., 1993; Tabata and Kornberg, 1994; Capdevila et al., 1994; Forbes et al., 1996b) and this pattern of expression depends upon hh activity (Hidalgo and Ingham, 1990; Capdevila et al., 1994) whilst ectopic expression of hh induces inappropriate levels of ptc transcription (Ingham, 1993; Basler and Struhl, 1994; Capdevila and Guerrero, 1994; Diaz-Benjumea and Cohen, 1994; Tabata and Kornberg, 1994; Ingham and Fietz, 1995; Forbes et al., 1996b). High-level expression of ptc is thus an indication of a response to hh signalling that is independent of cell or tissue type. Although the functional significance of this regulatory relationship is currently unclear, one possibility is that Ptc activity acts as a sink for the Hh protein; in this case, its upregulation in response to Hh would effectively limit the range of the Hh signal. Alternatively, the increase in ptc expression could act as a feedback mechanism that attenuates the response of a cell to the hh signal.

Our analysis of ptc1 expression in the zebrafish embryo does not allow us to distinguish between these or other possibilities; however, it does provide compelling evidence that this regulatory interaction has itself been conserved during evolution and is therefore likely to be fundamental to the mechanism of hh signalling. Thus as in flies, the highest levels of ptc1 transcription are invariably associated with shh-expressing cells; for instance, in the pectoral fin buds and the foregut high-level ptc1 expression is found immediately adjacent to cells expressing shh, while in the CNS, the levels of ptc1 transcript are at their highest in the ventral region of the neural tube adjacent to the shh-expressing cells of the axial mesoderm and floorplate. A similar relationship exists between the expression domains of ptc homologues and shh both in the chick (Marigo et al., 1996) and mouse (Goodrich et al., 1996). However, in each of these organisms, additional sites of elevated ptc expression are observed in tissues associated with the expression of two other hh family members, Indian (Ihh) and Desert (Dhh) hedgehog (Goodrich et al., 1996; Marigo et al., 1996; Bitgood et al., 1996). In zebrafish, the expression of tiggy-winkle hedgehog (twhh), the only other hh family gene characterised to date in this organism, although initiated slightly earlier, is entirely included within the shh expression domain (Ekker et al., 1995). Thus, in contrast to amniotes, there is no expression of ptc1 that can be specifically associated with the activity of a hh family gene other than shh. On the contrary, except in the fin buds, where twhh is not expressed, it is possible that at least some aspects of the transcriptional regulation of ptc1 may in part be mediated by twhh. In midline mutants that lack shh expression at specific positions along the body axis, we observe changes in the pattern of ptc1 transcription consistent with a regulatory interaction between the two genes. In cyc mutants, high-level ptc1 expression is completely absent from the brain, consistent with the lack of shh expression in the ventral floor of the brain typical of this mutant (Krauss et al., 1993; Ekker et al., 1995; MacDonald et al., 1995). Since cyc embryos also lack twhh expression along the entire midline (Ekker et al., 1995), we cannot rule out the possibility that this gene may also be necessary (and indeed sufficient) for ptc1 transcription in the brain. However, the essentially normal expression of ptc1 along the ventral neural tube of cyc mutants, suggests that in this part of the body at least, shh activity is sufficient for the regulation of ptc1 transcription. In flh mutants by contrast, shh expression disappears from the axial mesoderm at a relatively early stage, many of the cells along the ventral neural tube failing to differentiate into floorplate presumably as a consequence (Talbot et al., 1995). In line with this widespread loss of shh expression, transcription of ptc1 is also severely reduced, being maintained only around the few scattered islands of cells that still express shh. Taken with our finding that transcription of ptc1 can be induced ectopically in the neural tube by mis-expression of shh, these data strongly suggest that shh is both sufficient and necessary for transcriptional activation of ptc1 transcription in the neural tube.

ptc1 regulation reflects a role of shh in somite patterning in the zebrafish

Whereas low level transcription of ptc1 is detectable throughout the presomitic mesoderm, its expression is significantly elevated in the adaxial cells that flank the midline mesoderm. Several lines of evidence suggest that this mesodermal expression of ptc1 depends upon shh rather than twhh activity. First, at this stage, expression of twhh is restricted to the neuroectoderm (Ekker et al., 1995) while shh is expressed at high levels throughout the axial mesoderm (Krauss et al., 1993). Second, in cyc embryos, which lack twhh expression along the entire midline (Ekker et al., 1995) but have normal levels of shh expression in the axial mesoderm, expression of ptc1 in the somitic mesoderm is normal. Conversely, the absence of shh expression along the axial midline caused by the flh mutation, is accompanied by a significant loss of ptc1 expression in the mesoderm. When considered with our finding that mis-expression of shh induces high-level transcription of ptc1 throughout the paraxial mesoderm, these data strongly suggest that the mesodermal expression of ptc1 is regulated by shh.

Interactions between the notochord and paraxial mesoderm have previously been well documented in higher vertebrates (Dietrich et al., 1993; Koseki et al., 1993; Pourquie et al., 1993; Goulding et al., 1994) and indeed the activity of Shh has been directly implicated in these interactions (Fan and TessierLavigne, 1994; Johnson et al., 1994; Fan et al., 1995). However, whereas in zebrafish, the induction of ptc1 expression in the adaxial cells appears to involve a very shortrange effect of shh, in the chick, Shh protein acts over relatively long distances (Fan and Tessier-Lavigne, 1994; Fan et al., 1995). Moreover, adaxial cells give rise to a myotomal lineage (Felsenfeld et al., 1991; Thisse et al., 1993) and indeed express the muscle-specific transcription factor myoD, which can itself be induced by shh activity (this report; Weinberg et al., 1996). By contrast, in the mouse and chick, Shh induces expression of the sclerotomal marker Pax1 (Fan and TessierLavigne, 1994; Johnson et al., 1994; Fan et al., 1995); in line with this, notochord grafts have been shown to induce differentiation of sclerotome at the expense of myotome, actually repressing MyoD expression (Goulding et al., 1994). Thus, it appears that the outcome of shh signalling on somite patterning differs between lower and higher vertebrates. The different organisation of the somites in amniotes and teleosts may thus be accounted for, at least in part, by a modification of the inductive signals that control their specification. Interestingly recent studies have revealed that, in chick, MyoD expression can be induced by Shh in presomitic mesoderm when presented in combination with signals from the dorsal neural tube (Munsterberg et al., 1995), an activity that may reflect its original role in the evolution of the vertebrate body plan.

The role of PKA in shh signalling

Intriguingly, as in Drosophila, we have found that the effects of ectopic shh activity on ptc1 expression can be mimicked by reducing or eliminating the activity of PKA. Although PKA has previously been implicated in signalling by Shh protein in the brain and somites of rats and chicks, respectively (Fan et al., 1995; Hynes et al., 1995), our finding that transcription of a ptc homologue in both the neuroectoderm and paraxial mesoderm is repressed by PKA provides the most direct indication to date of the extent to which the elements of the hh signalling pathway have been conserved. Similar effects are seen on the transcription of other targets of shh activity, both in the mesoderm, where myoD transcription is induced, and in the neural tube, where nk2.2 is inappropriately expressed. In addition, we consistently observe the ectopic activation of pax[b] in the developing eye, an effect that presages eye abnormalities similar to, though less extreme than, those induced by ectopic shh expression (unpublished observations). Analogous effects of a dnPKA on zebrafish eye development have also recently been described by Hammerschmidt et al. (1996).

The implication of PKA in the regulation of ptc1 is further supported by our finding that constitutive activity of PKA suppresses ptc1 transcription throughout the brain and neural tube. This effect is accompanied by a similar suppression of nk2.2 expression and by cyclopia, a condition clearly associated with a loss of midline signalling (Krauss et al., 1993; Hatta et al., 1994; Ekker et al., 1995; MacDonald et al., 1995). Together, all of these effects strongly support the notion that the shh signal acts by antagonising the repressive activity of PKA on various target genes, just as in Drosophila, hh counteracts the repressive effect of PKA on wg, dpp and ptc transcription. In Drosophila, it has been argued that PKA acts in parallel to the hh signalling cascade (Jiang and Struhl, 1995; Kalderon, 1995). In this regard, it is interesting that co-expression of shh and the cPKA mutant results in a mutual suppression of the two contrasting eye phenotypes induced by either activity alone. This result implies a balance between two opposing signals rather than a strictly linear relationship between the two, supporting the conclusions from the Drosophila studies. Further analyses will be required to elucidate the precise relationship between these activities.

We are especially grateful to T. Simmonds for running the fish facility, G. Clark and A. Davies for help with the automated sequencing, I. Goldsmith for oligonucleotide synthesis and our colleagues on the 4th floor at the ICRF for interest and advice. We thank T. Jowett and C. Kimmel for providing mutant fish strains, A. Barth, D. Henrique, T. Jowett, D. Kalderon, S. McKnight, D. Simmons and E. Weinberg for various clones and K. Zinn for providing the cDNA library. L. V. G. is an HHMI predoctoral fellow, R. L. J. is a fellow of the Damon Runyon-Walter Winchell Foundation Cancer Research Fund (DRG 1218) and M. P. S. is supported by the HHMI. The experimental part of this study was funded by the Imperial Cancer Research Fund and by grants from the Human Frontiers Science Programme and the European Community Human Capital and Mobility Programme to P. W. I.

Barth
,
K. A.
and
Wilson
,
S.
(
1995
).
Expression of zebrafish nk2.2 is influenced by sonic hedgehog/vertebrate hedgehog-1 and demarcates a zone of neuronal differentiation in the embryonic forebrain
.
Development
121
,
1755
1768
.
Basler
,
K.
and
Struhl
,
G.
(
1994
).
Compartment boundaries and the control of Drosophila limb pattern by hedgehog protein
.
Nature
368
,
208
214
.
Bitgood
,
M. J.
,
Shen
,
L.
and
McMahon
,
A. P.
(
1996
).
Sertoli cell signalling by desert hedgehog regulates the male germline
.
Current Biology
6
,
298304
.
Capdevila
,
J.
,
Estrada
,
M. P.
,
Sanchez-Herrero
,
E.
and
Guerrero
,
I.
(
1994
).
The Drosophila segment polarity gene patched interacts with decapentaplegic in wing development
.
EMBO J
.
13
,
71
82
.
Capdevila
,
J.
and
Guerrero
,
I.
(
1994
).
Targeted expression of the signalling molecule decapentaplegic induces pattern duplications and growth alterations in Drosophila wings
.
EMBO J
.
13
,
4459
4468
.
Chang
,
D. T.
(
1994
).
Products, genetic linkage and limb patterning activity of a murine hedgehog gene
.
Development
120
,
3339
3353
.
Diaz-Benjumea
,
F. J.
and
Cohen
,
S. M.
(
1994
).
Cell-interaction between compartments establishes the proximal-distal axis of Drosophila legs
.
Nature
372
,
175
179
.
Dietrich
,
S.
,
Schubert
,
F. R.
and
Gruss
,
P.
(
1993
).
Altered Pax gene expression in murine notochord mutants: The notochord is required to initiate and maintain ventral identity in the somite
.
Mech. Dev
.
44
,
189
207
.
Echelard
,
Y.
,
Epstein
,
D. J.
,
St.-Jacques
,
B.
,
Shen
,
L.
,
Mohler
,
J.
,
McMahon
,
J. A.
and
McMahon
,
A. P.
(
1993
).
Sonic hedgehog, a member of a family of putative signalling molecules is implicated in the regulation of CNS and limb polarity
.
Cell
75
,
1417
1430
.
Ekker
,
S.C.
,
Ungar
,
A. R.
,
Greenstein
,
P.
,
von Kessler
,
D. P.
,
Porter
,
J.A.
,
Moon
,
R.T.
, and
Beachy
,
P.A.
(
1995
).
Patterning activities of vertebrate hedgehog proteins in the developing eye and brain
.
Current Biology
5
,
944955
.
Fan
,
C.-M.
and
Tessier-Lavigne
,
M.
(
1994
).
Patterning of Mammalian Somites by the Surface Ectoderm and the Notochord: Evidence for Sclerotome Induction by Sonic hedgehog/Vhh-1
.
Cell
79
,
1175
1186
.
Fan
,
C.-M.
,
Porter
,
J. A.
,
Chiang
,
C.
,
Chang
,
D. T.
,
Beachy
,
P. A.
and
Tessier-Lavigne
,
M.
(
1995
).
Long-Range Induction of Sclerotome by Sonic hedgehog: A Direct Role for the Amino Terminal Product of Autoproteolytic Cleavage and Modulation by the Cyclic AMP Signalling Pathway
.
Cell
81
,
457
465
.
Felsenfeld
,
A. L.
,
Curry
,
M.
and
Kimmel
,
C. B.
(
1991
).
The fub-1 mutation blocks initial myofibril formation in zebrafish muscle pioneer cells
.
Dev. Biol
.
148
,
23
36
.
Fietz
,
M.
,
Concordet
,
J.-P.
,
Barbosa
,
R.
,
Johnson
,
R.
,
Krauss
,
S.
,
McMahon
,
A. P.
,
Tabin
,
C.
and
Ingham
,
P. W.
(
1994
).
The hedgehog gene family in Drosophila and vertebrate development
.
Development
1994 Supplement, 43-51.
Forbes
,
A. J.
,
Nakano
,
Y.
,
Taylor
,
A. M.
and
Ingham
,
P. W.
(
1993
).
Genetic analysis of hedgehog signalling in the Drosophila embryo
.
Development
1993 Supplement, 115-124.
Forbes
,
A.J.
,
Lin
,
H.
,
Ingham
,
P. W.
, and
Spradling
,
A. C.
(
1996a
).
hedgehog is required for the proliferation and specification of somatic cells prior to egg chamber formation in Drosophila
.
Development
122
,
11251135
.
Forbes
,
A.J.
,
Spradling
,
A. C.
,
Ingham
,
P.W.
, and
Lin
,
H.
(
1996b
).
The role of segment polarity genes during early oogenesis in Drosophila
.
Development (in press)
.
Goodrich
,
L. V.
,
Johnson
,
R. L.
,
Milenokovic
,
L.
,
McMahon
,
J. A.
and
Scott
,
M. P.
(
1996
).
Conservation of the hedgehog/patched signalling pathway from flies to mice: induction of a mouse patched gene by hedgehog
.
Genes and Development
10
,
301
312
.
Goulding
,
M.
,
Lumsden
,
A.
and
Paquette
,
A. J.
(
1994
).
Regulation of pax-3 expression in the dermomyotome and its role in muscle development
.
Development
120
,
957
971
.
Halpern
,
M. E.
,
Ho
,
R.K.
,
Walker
,
C.
, and
Kimmel
,
C.
(
1993
).
Induction of Muscle Pioneers and Floor Plate is Distinguished by the zebrafish no tail mutation
.
Cell
75
,
99
111
.
Halpern
,
M. E.
,
Thisse
,
C.
,
Ho
,
R. K.
,
Thisse
,
B.
,
Riggleman
,
B.
,
Trevarrow
,
B.
,
Weinberg
,
E. S.
,
Postlethwait
,
J. H.
and
Kimmel
,
C. B.
(
1995
).
Cell-autonomous shift from axial to paraxial mesodermal development in zebrafish floating head mutants
.
Development
121
,
42574264
.
Hammerschmidt
,
H.
,
Bitgood
,
M. J.
and
McMahon
,
A. P.
(
1996
).
Protein kinase A is a common negative regulator of Hedgehog signalling in the vertebrate embryo
.
Genes Dev
.
10
,
647
658
.
Hatta
,
K.
,
Kimmel
,
C. B.
,
Ho
,
R. K.
and
Walker
,
C.
(
1991
).
The cyclops mutation blocks specification of the floor plate of the zebrafish central nervous system
.
Nature
350
,
339
341
.
Hatta
,
K.
,
Puschel
,
A. W.
and
Kimmel
,
C. B.
(
1994
).
Midline signalling in the primordium of the zebrafish anterior central nervous system
.
Proc. Natn. Acad. Sci., USA
91
,
2061
2065
.
Heberlein
,
U.
,
Singh
,
C. M.
,
Luk
,
A. Y.
, and
Donohoe
,
T. J.
(
1995
).
Growth and differentiation in the Drosophila eye coordinated by hedgehog
.
Nature
373
,
709
711
.
Henrique
,
D.
,
Adam
,
J. A.
,
Chitnis
,
A.
,
Lewis
,
J.
and
Ish-Horowicz
,
D.
(
1995
).
Expression of a Delta homologue in prospective neurons in the chick
.
Nature
375
,
787
790
.
Hidalgo
,
A.
and
Ingham
,
P.
(
1990
).
Cell patterning in the Drosophila segment: spatial regulation of the segment polarity gene patched
.
Development
110
,
291
302
.
Hooper
,
J.
and
Scott
,
M. P.
(
1989
).
The Drosophila patched gene encodes a putative membrane protein required for segmental patterning
.
Cell
59
,
751765
.
Hynes
,
M.
,
Porter
,
J. A.
,
Chiang
,
C.
,
Chang
,
D.
,
Tessier-Lavigne
,
M.
,
Beachy
,
P. A.
and
Rosenthal
,
A.
(
1995
).
Induction of midbrain dopaminergic neurons by Sonic hedgehog
.
Neuron
15
,
35
44
.
Ingham
,
P. W.
,
Taylor
,
A. M.
and
Nakano
,
Y.
(
1991
).
Role of the Drosophila patched gene in positional signalling
.
Nature
353
,
184
187
.
Ingham
,
P. W.
(
1993
).
Localised hedgehog activity controls spatially restricted transcription of wingless in the Drosophila embryo
.
Nature
366
,
560
562
.
Ingham
,
P. W.
and
Hidalgo
,
A.
(
1993
).
Regulation of wingless transcription in the Drosophila embryo
.
Development
117
,
283
291
.
Ingham
,
P.
(
1995
).
Signalling by hedgehog family proteins in Drosophila and vertebrate development
.
Current Opinion In Genetics and Development
5
,
492
498
.
Ingham
,
P. W.
and
Fietz
,
M. J.
(
1995
).
Quantitative Effects of hedgehog and decapentaplegic Activity on the Patterning of the Drosophila Wing
.
Current Biology
.
5
,
432
441
.
Jiang
,
J.
and
Struhl
,
G.
(
1995
).
Protein kinase A and Hedgehog signalling in Drosophila limb development
.
Cell
80
,
563
572
.
Johnson
,
R. L.
,
Laufer
,
E.
,
Riddle
,
R. D.
and
Tabin
,
C.
(
1994
).
Ectopic expression of Sonic hedgehog Alters Dorsal-Ventral Patterning of Somites
.
Cell
79
,
1165
1173
.
Jowett
,
T.
and
Lettice
,
L.
(
1994
).
Whole-mount in-situ hybridizations on zebrafish embryos using a mixture of digoxigenin-labelled and fluoresceinlabelled probes
.
Trends In Genetics
10
,
73
74
.
Kalderon
,
D.
(
1995
).
Responses to Hedgehog
.
Current Biology
5
,
580
582
.
Klingensmith
,
J.
and
Nusse
,
R.
(
1994
).
Signalling by wingless in Drosophila
.
Dev. Biol
.
166
,
396
414
.
Koseki
,
H.
,
Wallin
,
J.
,
Wilting
,
J.
,
Mizutani
,
Y.
,
Kispert
,
A.
,
Ebensperger
,
C.
,
Herrmann
,
B. G.
,
Christ
,
B.
and
Balling
,
R.
(
1993
).
A role for pax-1 as a mediator of notochordal signals during the dorsoventral specification of vertebrae
.
Development
119
,
649
660
.
Krauss
,
S.
,
Johansen
,
T.
,
Korzh
,
V.
and
Fjose
,
A.
(
1991
).
Expression of the zebrafish paired box gene pax[zf-b] during early neurogenesis
.
Development
113
,
1193
1206
.
Krauss
,
S.
,
Concordet
,
J.-P.
and
Ingham
,
P. W.
(
1993
).
A functionally conserved homolog of the Drosophila segment polarity gene hedgehog is expressed in tissues with polarising activity in zebrafish embryos
.
Cell
.
75
,
1431
1444
.
Lee
,
J. J.
,
von Kessler
,
D. P.
,
Parks
,
S.
and
Beachy
,
P. A.
(
1992
).
Secretion and localised transcription suggests a role in positional signalling for products of the segmentation gene hedgehog
.
Cell
70
,
777
789
.
Lepage
,
T.
,
Cohen
S. M.
,
Diaz-Benjumea
,
F. J.
and
Parkhurst
,
S. M.
(
1995
).
Signal transduction by cAMP-dependent protein kinase A in Drosophila limb patterning
.
Nature
373
,
711
715
.
Li
,
W.
,
J.
Ohlmeyer
,
T.
,
Lane
,
M. E.
and
Kalderon
,
D.
(
1995
).
Function of protein kinase A in Hedgehog signal transduction and Drosophila imaginal disc development
.
Cell
80
,
553
562
.
Ma
,
C.
,
Zhou
,
Y.
,
Beachy
,
P. A.
and
Moses
,
K.
(
1993
).
The segment polarity gene hedgehog is required for progression of the morphogenetic furrow in the developing Drosophila eye
.
Cell
75
,
927
938
.
Macdonald
,
R.
,
Barth
,
K. A.
,
Xu
,
Q.
,
Holder
,
N.
,
Mikkola
,
I.
and
Wilson
,
S. W.
(
1995
).
Midline signalling is required for Pax gene regulation and patterning of the eyes
.
Development
121
,
3267
3278
.
Marigo
,
V.
,
Scott
,
M. P.
,
Johnson
,
R. L.
,
Goodrich
,
L.
and
Tabin
,
C. J.
(
1996
).
Conservation in hedgehog signalling, Sonic hedgehog controls the expression of a chicken patched homolog in the developing limb
.
Development
122
,
1225
1233
Marti
,
E.
,
Bumcrot
,
D. A.
,
Takada
,
R.
and
McMahon
,
A. P.
(
1995
).
Requirement of, 19kDalton form of Sonic hedgehog for induction of distinct ventral cell types in vertebrate CNS explants
.
Nature
375
,
322
325
.
Mohler
,
J.
and
Vani
,
K.
(
1992
).
Molecular organisation and embryonic expression of the hedgehog gene involved in cell-cell communication in segmental patterning in Drosophila
.
Development
115
,
957
971
.
Munsterberg
,
A.
,
Kitajewski
,
J.
,
Bumcrot
,
D. A.
,
McMahon
,
A. P.
and
Lassar
,
A. B.
(
1995
).
Combinatorial signalling by Sonic hedgehog and Wnt family members induces myogenic bHLH gene expression in the somite
.
Genes Dev
.
9
,
2911
2922
.
Nakano
,
Y.
,
Guerrero
,
I.
,
Hidalgo
,
A.
,
Taylor
,
A. M.
,
Whittle
,
J. R. S.
and
Ingham
,
P. W.
(
1989
).
The Drosophila segment polarity gene patched encodes a protein with multiple potential membrane spanning domains
.
Nature
341
,
508
513
.
Orellana
,
S. A.
and
McKnight
,
G. S.
(
1992
).
Mutations in the catalytic subunit of cAMP-dependent protein kinase result in unregulated biological activity
.
Proc. Natn. Acad. Sci. USA
89
,
4726
30
.
Oxtoby
,
E.
and
Jowett
,
T.
(
1993
).
Cloning of the zebrafish krox-20 gene (krx20) and its expression during hindbrain development
.
Nuc. Acids Res
.
21
,
1087
1095
.
Pan
,
D.
and
Rubin
,
G. M.
(
1995
).
cAMP-dependent Protein Kinase and hedgehog Act Antagonistically in Regulating decapentaplegic Transcription in Drosophila Imaginal Discs
.
Cell
80
,
543
552
.
Placzek
,
M.
,
Tessier-Lavigne
,
M.
,
Yamada
,
T.
,
Jessel
,
T.
and
Dodd
,
J.
(
1990
).
Mesodermal control of neural cell identity: floor plate induction by the notochord
.
Science
250
,
985
988
.
Placzek
,
M.
,
Yamada
,
T.
,
Tessier-Lavigne
,
M.
,
Jessel
,
T.
and
Dodd
,
J.
(
1991
).
Control of dorsoventral pattern in vertebrate neural development: induction and polarising properties of the floor plate
.
Development
1991 Supplement, 105-122.
Pourquie
,
O.
,
Coltey
,
M.
,
Teillet
,
M. A.
,
Ordahl
,
C.
and
Ledouarin
,
N. M.
(
1993
).
Control of dorsoventral patterning of somitic derivatives by notochord and floor plate
.
Proc. Natn. Acad. Sci. USA
90
(
11
),
5242
5246
.
Riddle
,
R.
,
Johnson
,
R. L.
,
Laufer
,
E.
and
Tabin
,
C.
(
1993
).
Sonic Hedgehog mediates the polarizing activity of the ZPA
.
Cell
75
,
1401
1416
.
Roelink
,
H.
,
Augsburger
,
A.
,
Heemskerk
,
J.
,
Korzh
,
V.
,
Norlin
,
S.
,
Ruiz i Altaba
,
A.
,
Tanabe
,
Y.
,
Placzek
,
M.
,
Edlund
,
T.
,
Jessell
,
T. M.
and
Dodd
,
J.
(
1994
).
Floor plate and motor neuron induction by vhh-1, a vertebrate homolog of hedgehog expressed by the notochord
.
Cell
76
,
761
775
.
Roelink
,
H.
,
J.
Porter
,
C.
Chiang
,
Y.
Tanabe
,
D. T.
Chang
,
P. A.
Beachy
and
Jessell
,
T. M.
(
1995
).
Floor plate and motor neuron induction by different concentrations of the amino terminal cleavage product of Sonic Hedgehog autoproteolysis
.
Cell
81
,
445
455
.
Ruiz i Altaba
,
A.
,
Roelink
,
H.
and
Jessell
,
T. M.
(
1995
).
Restrictions to floorplate induction by hedgehog and winged helix genes in the neural tube of frog embryos
.
Mol. Cell Neuroscience
6
,
106
121
.
Seed
,
B.
(
1987
).
An LFA-3 cDNA encodes a phospholipid-linked membrane protein homologous to its receptor CD2
.
Nature
329
,
840
842
Strutt
,
D. I.
,
Wiersdorff
,
V.
and
Mlodzik
,
M.
(
1995
).
Regulation of furrow progression in the Drosophila eye by cAMP-dependent protein kinase A
.
Nature
373
,
705
709
.
Tabata
,
T.
,
Eaton
,
S.
and
Kornberg
,
T. B.
(
1992
).
The Drosophila hedgehog gene is expressed specifically in posterior compartment cells and is a target of engrailed regulation
.
Genes Dev
.
6
,
2635
2645
.
Tabata
,
T.
and
Kornberg
,
T. B.
(
1994
).
Hedgehog is a signalling protein with a key role in patterning Drosophila imaginal discs
.
Cell
76
,
89
102
.
Tabata
,
T.
,
Schwartz
,
C.
,
Gustavson
,
E.
,
Ali
,
Z.
and
Kornberg
,
T.
(
1995
).
Creating a Drosophila wing de-novo, the role of engrailed, and the compartment border hypothesis
.
Development
121
,
3359
3369
.
Talbot
,
W. S.
,
Trevarrow
,
B.
,
Halpern
,
M. E.
,
Melby
,
A. E.
,
Farr
,
G.
,
Postlethwait
,
J. H.
,
Jowett
,
T.
,
Kimmel
,
C. B.
and
Kimelman
,
D.
(
1995
).
A homeobox gene essential for zebrafish notochord development
.
Nature
378
,
150
157
.
Taylor
,
A. M.
,
Nakano
,
Y.
,
Mohler
,
J.
and
Ingham
,
P. W.
(
1993
).
Contrasting distributions of patched and hedgehog proteins in the Drosophila embryo
.
Mech. Dev
.
43
,
89
96
.
Thisse
,
C.
,
Thisse
,
B.
,
Schilling
,
T. F.
and
Postlethwait
,
J. H.
(
1993
).
Structure of the zebrafish snail1 gene and its expression in wild-type, spadetail and no tail mutant embryos
.
Development
119
,
1203
1215
.
Thisse
,
C.
,
Thisse
,
B.
,
Halpern
,
M. E.
and
Postlethwait
,
J. H.
(
1994
).
Goosecoid expression in neurectoderm and mesendoderm is disrupted in zebrafish cyclops gastrulas
.
Dev. Biol
.
64
,
420
429
.
Turner
,
D L
and
Weintraub
H.
(
1994
).
Expression of achaete-scute homolog3 in Xenopous embryos converts ectodermal cells to a neural fate
.
Genes Dev
.
8
,
1434
1447
.
van Straaten
,
H. W. M.
,
Hekking
,
J. W. M.
,
Beursgens
,
J. P. W. M.
,
Terwindt-Rouwenhorst
,
E.
and
Drukker
,
J.
(
1989
).
Effect of the notochord on proliferation and differentiation in the neural tube of the chick embryo
.
Development
107
,
793
803
.
Weinberg
,
E. S.
,
Allende
,
M. L.
,
Kelly
,
C. S.
,
Abdelhamid
,
A.
,
Murakami
,
T.
,
Andermann
,
P.
,
Doerre
,
O. G.
,
Grunwald
,
D. J.
and
Riggleman
,
B.
(
1996
).
Developmental regulation of zebrafish MyoD in wild type, no tail and spadetail embryos
.
Development
122
,
271
280
.
Wilson
,
R.
et al. , (
1994
).
2.2 Mb of contiguous nucleotide sequence from chromosome III of C. elegans
.
Nature
368
,
32
38
.
Yamada
,
T.
,
Placzek
,
M.
,
Tanaka
,
H.
,
Dodd
,
J.
and
Jessell
,
T. M.
(
1991
).
Control of cell patterning in the developing nervous system: polarizing activity of the floor plate and notochord
.
Cell
64
,
635
647
.