Fibroblast growth factor receptor 1 (Fgfr1) plays pleiotropic roles during embryonic development, but the mechanisms by which this receptor signals in vivo have not previously been elucidated. Biochemical studies have implicated Fgf receptor-specific substrates (Frs2, Frs3) as the principal mediators of Fgfr1 signal transduction to the MAPK and PI3K pathways. To determine the developmental requirements for Fgfr1-Frs signaling, we generated mice (Fgfr1ΔFrsFrs) in which the Frs2/3-binding site on Fgfr1 is deleted. Fgfr1ΔFrsFrs embryos die during late embryogenesis, and exhibit defects in neural tube closure and in the development of the tail bud and pharyngeal arches. However, the mutant receptor is able to drive Fgfr1 functions during gastrulation and somitogenesis, and drives normal MAPK responses to Fgf. These findings indicate that Fgfr1 uses distinct signal transduction mechanisms in different developmental contexts, and that some essential functions of this receptor are mediated by Frs-independent signaling.

Members of the receptor tyrosine kinase (RTK) superfamily elicit similar cellular, transcriptional and biochemical responses when stimulated in cultured cells (Fambrough et al.,1999; Schlessinger,2000), yet direct specific, non-redundant cellular responses in developing organisms. It has been proposed that qualitative differences in signal transduction underlie RTK functional specialization. Such differences could be in ligand sensitivity, effector pathway use or the strength or duration of signaling. This model has gained support from in vivo domain swap experiments, which have demonstrated biologically significant differences in RTK signal transduction (Klinghoffer et al., 2001; Madhani,2001; Hamilton et al.,2003). Analyses of mice harboring signaling domain mutations have also shown that effector pathways are differentially required within the RTK superfamily, and that many signaling events observed in biochemical studies are dispensable or required only for a subset of the in vivo functions of receptors (Partanen et al.,1998; Tallquist et al.,2000; Maina et al.,2001; Klinghoffer et al.,2002; Tallquist et al.,2003).

The mammalian fibroblast growth factor (Fgf) signaling network consists of four high affinity RTKs and at least 22 ligands(Ornitz, 2000; Böttcher and Niehrs,2005). Mouse knockout studies have identified numerous roles of Fgfs, and demonstrated that all essential embryonic roles are mediated by two receptors, Fgfr1 and Fgfr2(Deng et al., 1994; Yamaguchi et al., 1994; Deng et al., 1996; Arman et al., 1998; Yu et al., 2000; Böttcher and Niehrs,2005). We have focused our studies on Fgfr1, which is required for postimplantation growth, mesodermal migration and patterning during gastrulation, and for somitogenesis. Fgfr1nullmouse embryos also exhibit posterior truncations and neural tube closure defects (Deng et al., 1994; Yamaguchi et al., 1994). Studies of hypomorphic and conditional alleles have identified additional roles of this receptor in node regression, neural stem cells, olfactory bulb development, and patterning of the anteroposterior axis, central nervous system and pharyngeal arches (Partanen et al., 1998; Tropepe et al.,1999; Xu et al.,1999; Hébert et al.,2003; Trokovic et al.,2003a; Trokovic et al.,2003b).

Though many roles of Fgfr1 are known, the mechanisms by which this receptor signals in vivo have not yet been elucidated. Active Fgfrs can directly engage relatively few proteins, namely Crk, Grb14, Shb, PLCγ and Frs2,3(hereafter referred to collectively as `Frs')(Mohammadi et al., 1991; Wang et al., 1996; Kouhara et al., 1997; Larrson et al., 1999; Reilly et al., 2000; Cross et al., 2002). Biochemical studies have implicated Frs adaptors as the key mediators of Fgfr signal transduction. These proteins interact constitutively with the intracellular juxtamembrane regions of Fgfr1 and Fgfr2, unlike signaling proteins downstream of other RTKs that are recruited after receptor activation. Fgfr activation leads to phosphorylation of Frs tyrosine residues to which Grb2 and SHP2 are subsequently recruited, initiating PI3K and MAPK signaling (Wang et al., 1996; Kouhara et al., 1997; Xu et al., 1998; Ong et al., 2000; Hadari et al., 2001). This adaptor-mediated mechanism may distinguish the kinetics, amplitude or subcellular localization of Fgfr signal transduction from other RTK signaling events.

Frs adaptors are conserved among vertebrates, and expression patterns suggest that the two mammalian isoforms perform non-redundant functions in vivo (McDougall et al., 2001; Gotoh et al., 2004). Loss-of-function studies have not been reported for Frs3, but Frs2-/- mouse embryos die ∼E7.5 with defects in extra-embryonic development (Hadari et al., 2001; Gotoh et al.,2005). Mouse chimera and Xenopus knockdown studies have further implicated Frs2 in mesoderm development and convergent extension,respectively (Akagi et al.,2002; Gotoh et al.,2005). These studies are informative with respect to Frs2 roles,but they are difficult to interpret in terms of specific growth factor pathways because of the promiscuity of Frs adaptors. In addition to Fgfrs, Frs can interact with active neurotrophin receptors and are phosphorylated downstream of other RTKs (Rabin et al.,1993; Ong et al.,2000). Furthermore, both adaptors associate strongly with the cell cycle regulatory protein p13suc1, and may thus play a role in cell cycle regulation or progression (Rabin et al., 1993; Ong et al.,1996).

To assess the developmental requirements for Frs-mediated signaling downstream of Fgfr1, we have generated mice lacking the Frs-binding site on this receptor (Fgfr1ΔFrs). We found that Frs-mediated signaling is dispensable for Fgfr1 functions during gastrulation and somitogenesis, but is required for Fgfr1 roles in neurulation, tail bud and pharyngeal arch development. Furthermore, we found that primary embryonic cells do not require this signaling interaction to elicit strong and sustained MAPK responses to Fgf, and that MAPK activation in vivo is not grossly affected by the Fgfr1ΔFrsmutation. These results indicate that Frs adaptors are not the exclusive effectors of Fgfr1 signal transduction in vivo.

Targeting vector construction

Targeting vectors were generated in pPGKneoF2L2DTA, a plasmid containing a DTA selection cassette and a neomycin cassette flanked by nested loxP and FRT sites. For knock-in (KI) vectors (Fgfr1wtKI,Fgfr1ΔFrs), the short arm of homology, an XbaI/HindIII fragment spanning introns 6-7 of Fgfr1, was inserted upstream of the 5′ loxP site in the same orientation as the neomycin cassette. Intron 7-partial cDNA cassettes were generated through two rounds of PCR SOEing (splicing by overhang extension,oligos are given in Table S1 in the supplementary material)(Pogulis et al., 2000). In the first SOEing reaction, nucleotides 1276-1356 (amino acids 407-433) were deleted from a mouse Fgfr1 cDNA construct to generate an Fgfr1ΔFrscDNA. In the second round, partial Fgfr1wt, Fgfr1ΔFrs cDNAs (exons 8-17) were spliced to the 3′ end of intron 7, preserving the endogenous intron 7/exon 8 junction. All PCR products incorporated into targeting vectors were generated using Pfu (Stratagene) and were verified by sequencing. Intron 7-partial cDNA cassettes were digested with Tth111I and ligated to the long arm of homology, a ∼4.1 kb Tth111I/PstI genomic fragment. Resultant fragments were cloned between the 3′ loxP site and DTA(Fig. 1B).

For the Fgfr1floxex4 vector, the short arm of homology,exon 4 fragment, and part of the 3′ long arm of homology were PCR amplified from wild type 129Sv genomic DNA (oligos are given in Table S1 in the supplementary material). The exon 4 fragment was cloned between loxP sites, upstream of the FRT-flanked neomycin cassette in pPGKneoF2L2DTA. The short arm was inserted upstream of the 5′ loxP site; the intron 5 PCR fragment was ligated to a genomic HindIII/EcoRV fragment to complete the long arm, which was inserted between the second loxP site and DTA(Fig. 1C).

Generation of mouse lines

AK7 ES cells were electroporated with linearized targeting vectors, G418 selected and PCR screened for integration at the Fgfr1 locus (data not shown). Targeting was verified by Southern analysis with the following probes: Fgfr1KI-5′ external(NcoI/XmaI), internal (SpeI/NciI),3′ external (PstI/HindIII); Fgfr1Δex4-5′ external (PCR product from 129Sv genomic DNA, oligos are given in Table S1 in the supplementary material); internal (NotI/NcoI).

Correctly targeted ES cells were injected into wild type C57BL/6J blastocysts to generate chimeras. The neomycin cassette caused early embryonic lethality of KI homozygotes and was excised through breeding to Meox2-Cre or FlpeR mice (Farley et al.,2000; Tallquist and Soriano,2000). Complete excision of neomycin (Fgfr1KI,Fgfr1floxex4) and exon 4 (Fgfr1floxex4)was verified by Southern analysis (data not shown). Tail biopsies were routinely PCR genotyped (oligos are given in Table S1 in the supplementary material). In KI homozygotes, embryonic RT-PCR confirmed that targeting did not disrupt alternative splicing of the extracellular domain or splicing of upstream exons into the cDNA (data not shown).

Whole-mount in situ hybridization

Whole-mount in situ hybridization was performed using standard procedures.

RNA preparation and semi-quantitative RT-PCR

Embryonic RNA was prepared using Trizol (GibcoBRL) and reverse transcribed as described previously (Chen and Soriano,2003). Diluted cDNA pools were amplified using the following primer pairs: FGFR1.52/FGFR1.36, FGFR2ex11.51/FGFR2ex13.31,Timm_cDNA+/Timm_cDNA- (see Table S1 in the supplementary material). Triplicate PCR reactions were cycled within a linear range of amplification and products were detected and quantified using Sybr Gold (Molecular Probes) and NIH ImageQuant software. Fgfr1 and Fgfr2 RT-PCR products were normalized to Timm (mitochondrial membrane transport protein)levels.

Immunohistochemistry

Whole-mount phospho-MAPK IHC was performed as described previously(Corson et al., 2003), with minor modifications.

For thin section immunohistochemistry, embryos were fixed overnight (4%PFA), embedded in paraffin wax and sectioned. After rehydration, slides were pretreated in (3% H2O2, 10% methanol, 1×PBS). Antigen retrieval was performed by microwaving slides in 10 mM citrate buffer(pH 6.0) (phosphohistone H3, Upstate Biotechnology #06-570) or 10 mM EDTA (pH 8.0) (pMAPK, Cell Signaling #9101S). Samples were blocked in 5% serum/PBS and incubated with antibody overnight. Immunocomplexes were detected using biotinylated secondary antibodies with ABC Elite and DAB substrate kits (Vectastain).

Nile Blue and TUNEL analysis

Freshly isolated embryos were stained in Nile Blue (0.003% in DMEM, 0.5%FBS) for 30-60 minutes (37°C) and washed in PBS. For TUNEL analysis,embryos were fixed, embedded and sectioned as for immunohistochemistry. Rehydrated sections were incubated in (0.3% Triton-X-100, 1x PBS), digested with proteinase K, and incubated in TdT reaction mix for 1 hour, 37°C [30 mM Tris pH 7.5, 140 mM cacodylate, 1 mM CoCl2, 10 μM dig-dUTP,0.3 U/μL TdT (Roche)]. Reactions were stopped in PBS and TUNEL-positive nuclei were visualized by IHC with anti-Dig-AP Fab (Roche) and NBT/BCIP detection.

Skeletal preparations

Skeletons were prepared and stained with Alcian Blue and Alizarin Red using a standard protocol.

Mouse embryonic cell (MEC) isolation and culture

E12.5-14.5 embryos were trypsinized (5′, 37°C), gently disaggregated with a Pasteur pipette, and plated on gelatin in DMEM, 15% FBS. To generate immortalized lines, p1 MECs were infected with an E6LTTNLoxP retrovirus transducing the SV40 Large T antigen(Berghella et al., 1999). Infected cells were selected with 500 μg/ml G418 and resistant colonies were pooled.

Immunoprecipitation (IPs) and western blotting

MECs were seeded at 1.5-1.8×104 cells/cm2,serum starved (24-36 hours, 0.5% FBS) and stimulated with heparin and Fgf(aFgf and bFgf gave similar results; Research Diagnostics). Cells were then washed with cold PBS and lysed in HNTG (20 mM HEPES pH 7.9, 150 mM NaCl, 1%Triton X-100, 10% glycerol, 1.5 mM MgCl2, 1 mM EGTA, 1 μg/ml aprotinin, 1 mM PMSF, 10 mM NaPPi, 0.2 mM activated Na2VO3, 50 mM NaF). IPs were performed overnight in HNTG; immunocomplexes were pulled down using protein A- or G-PLUS agarose(Santa Cruz) and washed in HNTG or (500 mM NaCl, 10 mM Tris pH 7.5, 2 mM EDTA,1% NP40). SDS-PAGE and western blotting were performed according to standard protocols, with phospho-specific and -nonspecific antibody blots blocked in 1%BSA and 3% milk, respectively.

Antibodies: pMAPK (Cell Signaling); Frs2 (Santa Cruz H-91); Fgfr2 (mouse monoclonal, Research Diagnostics); RasGAP (70.3; gift of Andrius Kazlauskas; Valius et al., 1993);phosphotyrosine (clone 4G10, Upstate Biotechnology); pAkt (Cell Signaling);Crk (BD Transduction Labs); and actin (clone AC-15, Sigma).

Generation of Fgfr1ΔFrs and Fgfr1wtKI mice

We used a partial cDNA knock-in approach to generate mice lacking the Frs binding site on Fgfr1 without disrupting alternative splicing of exons encoding the extracellular domain (Fig. 1) (Werner et al.,1992; Yan et al.,1992; Johnson and Williams,1993). This deletion mutation has previously been shown to abrogate Fgfr1 binding to both Frs2 and Frs3 without disrupting other receptor functions and interactions (Xu et al.,1998). As a control, we used the same approach to generate wild-type knock-in mice (Fgfr1wtKI; the untargeted allele is designated Fgfr1+).

Fig. 1.

Fgfr1 targeting. (A) Alleles introduced by gene targeting. (B) Partial cDNA knock-in approach: Fgfr1 exons 8-17, encoding the transmembrane and cytoplasmic domains, were replaced with a pseudoexon encoding the corresponding region of Fgfr1wtKI or Fgfr1ΔFrs. Partial cDNAs were spliced at the 5′ end into exon 8 and at the 3′ end into exon 17, upstream of the endogenous polyadenylation sequence. (C) To generate the null allele, the first exon common to all reported Fgfr1 splice variants (exon 4) was flanked with loxP sites and excised in vivo by Meox2-Cre(Tallquist and Soriano, 2000). RT-PCR from embryonic RNA confirmed the generation of a stop codon soon after the exon3-exon5 splice junction (data not shown). (D) Southern blots verifying correct targeting of all alleles in ES cell clones. Abbreviations:A, ApaI; H, HindIII; KI, knock-in; Nd, NdeI; Nh, NheI; P, PstI; Rv, EcoRV; T, Tth111I; X, XbaI; ext, external; int, internal. (E) Relative Fgfr1 mRNA levels in homozygous embryos, assessed by semi-quantitative RT-PCR. Each point is the mean of triplicate reactions for a single embryo (±s.d.).

Fig. 1.

Fgfr1 targeting. (A) Alleles introduced by gene targeting. (B) Partial cDNA knock-in approach: Fgfr1 exons 8-17, encoding the transmembrane and cytoplasmic domains, were replaced with a pseudoexon encoding the corresponding region of Fgfr1wtKI or Fgfr1ΔFrs. Partial cDNAs were spliced at the 5′ end into exon 8 and at the 3′ end into exon 17, upstream of the endogenous polyadenylation sequence. (C) To generate the null allele, the first exon common to all reported Fgfr1 splice variants (exon 4) was flanked with loxP sites and excised in vivo by Meox2-Cre(Tallquist and Soriano, 2000). RT-PCR from embryonic RNA confirmed the generation of a stop codon soon after the exon3-exon5 splice junction (data not shown). (D) Southern blots verifying correct targeting of all alleles in ES cell clones. Abbreviations:A, ApaI; H, HindIII; KI, knock-in; Nd, NdeI; Nh, NheI; P, PstI; Rv, EcoRV; T, Tth111I; X, XbaI; ext, external; int, internal. (E) Relative Fgfr1 mRNA levels in homozygous embryos, assessed by semi-quantitative RT-PCR. Each point is the mean of triplicate reactions for a single embryo (±s.d.).

Fgfr1wtKI/wtKI and Fgfr1Δex4/Δex4 phenotypes on a high percentage C57BL/6J background

Following two or more backcross generations to the C57BL/6J background, we recovered nearly the expected ratio of Fgfr1wtKI/wtKIanimals at E16.5, and a number of homozygotes survived to adulthood with normal fertility (Table 1). We therefore selected this background (>75% contribution from C57BL/6J) for our studies. On this genetic background, Fgfr1wtKI/wtKIembryos are often smaller than littermates and exhibit digit 3/4 fusions (data not shown). Importantly, however, Fgfr1ΔFrsFrs phenotypes discussed below are either absent or notably less penetrant/severe in Fgfr1wtKI/wtKI embryos (see Fig. S1 in the supplementary material). Phenotypic differences between the knock in lines were not due to differences in Fgfr1 expression: Fgfr1 mRNA levels are similar in untargeted, Fgfr1wtKI/wtKI and Fgfr1ΔFrsFrs embryos on the high percentage C57 background (Fig. 1E).

Table 1.

Genotypes of offspring from heterozygote knock-in (KI) intercrosses at late embryonic and postnatal stages

AgeLinen+/+ (%)+/KI (%)KI/KI (%)
P4-P7 wtKI #1 72 24 (33) 36 (50) 12 (17) 
 wtKI #2 31 9 (29) 20 (65) 2 (6) 
 ΔFrs 60 18 (30) 42 (70) 
E16.5 wtKI #2 24 7 (29) 13 (54) 4 (17) 
 wtKI #3 17 4 (24) 10 (59) 3 (18) 
 ΔFrs 22 4 (18) 17 (77) 1 (5) 
AgeLinen+/+ (%)+/KI (%)KI/KI (%)
P4-P7 wtKI #1 72 24 (33) 36 (50) 12 (17) 
 wtKI #2 31 9 (29) 20 (65) 2 (6) 
 ΔFrs 60 18 (30) 42 (70) 
E16.5 wtKI #2 24 7 (29) 13 (54) 4 (17) 
 wtKI #3 17 4 (24) 10 (59) 3 (18) 
 ΔFrs 22 4 (18) 17 (77) 1 (5) 

wtKI #1-3 are KI mouse lines derived from independently targeted ES cell clones.

Because the Fgfr1wtKI/wtKI phenotype is background dependent (data not shown), we revisited the null phenotype on the (>75%)C57BL/6J background. We generated mice harboring a null allele(Fgfr1Δex4, Fig. 1A,C) and backcrossed them more than two generations to C57BL/6J. Compared with published Fgfr1null embryos, which were analyzed on different mixed backgrounds,Fgfr1Δex4ex4embryos on the (>75%) C57BL/6J background survive slightly later in embryogenesis (E11.5 vs. E7.5-9.5). However, they are always developmentally arrested prior to E9.5 and exhibit phenotypes similar to published Fgfr1null embryos, including a developmental delay,mesodermal migration and patterning defects, craniorachischisis and posterior truncations (Fig. 2; Fig. 3B)(Deng et al., 1994; Yamaguchi et al., 1994). Fgfr1Δex4ex4 embryos also have enlarged hearts with disrupted looping (data not shown).

Fgfr1-Frs signaling is not required during gastrulation or somitogenesis

Fgfr1ΔFrs/+ animals are viable and indistinguishable from littermate controls, but Fgfr1ΔFrsFrs mice were never recovered postnatally (Table 1). Timed matings indicated that Fgfr1ΔFrsFrs embryos survive much later in embryogenesis than do Fgfr1Δex4ex4 embryos, with a drop in viability only after E15.5 (Fig. 2A).

Fgfr1 roles in early mesodermal development were rescued in∼80% of Fgfr1ΔFrsFrsembryos. During gastrulation, Fgfr1 is required for mesodermal migration through the primitive streak. In Fgfr1Δex4ex4 (and published null) embryos, this migration is impaired and cells accumulate in the streak. Consequently, there is an expansion of axial mesoderm and reduction of paraxial mesoderm, which remains disorganized and fails to form somites (Fig. 2B,C)(Deng et al., 1994; Yamaguchi et al., 1994; Ciruna et al., 1997; Ciruna and Rossant, 2001). We examined the development of axial and paraxial mesoderm in Fgfr1ΔFrsFrs embryos by Shh and Meox1 in situ hybridizations, respectively. Although Fgfr1ΔFrsFrs embryos are developmentally delayed by ∼1 day, most mutants exhibit normal expression of both mesodermal markers following gastrulation(Fig. 2D-G). This indicates that mesodermal migration is rescued in these embryos. Furthermore, whereas Fgfr1Δex4ex4 embryos never form somites, Fgfr1ΔFrsFrsembryos undergo normal somitogenesis, as evidenced by the segmented pattern of Meox1 staining.

Fig. 2.

Rescue of early Fgfr1 functions by Fgfr1ΔFrs.(A) Viability (percent of expected) of homozygotes recovered from Fgfr1ΔFrs/+ and Fgfr1Δex4/+ intercrosses. (B-G)Sonic hedgehog (Shh, E9.5) and Meox1 (E10.5) in situ hybridization. Arrows indicate width of the Shh expression domain(more than two cell diameters in Fgfr1Δex4ex4 compared with one cell diameter in control embryos). (C) The paraxial mesoderm population is reduced and disorganized in Fgfr1Δex4ex4 embryos, so Meox1 staining is faint and diffuse. (Color development time required to visualize any positive signal was notably longer in this embryo than in F,G.)

Fig. 2.

Rescue of early Fgfr1 functions by Fgfr1ΔFrs.(A) Viability (percent of expected) of homozygotes recovered from Fgfr1ΔFrs/+ and Fgfr1Δex4/+ intercrosses. (B-G)Sonic hedgehog (Shh, E9.5) and Meox1 (E10.5) in situ hybridization. Arrows indicate width of the Shh expression domain(more than two cell diameters in Fgfr1Δex4ex4 compared with one cell diameter in control embryos). (C) The paraxial mesoderm population is reduced and disorganized in Fgfr1Δex4ex4 embryos, so Meox1 staining is faint and diffuse. (Color development time required to visualize any positive signal was notably longer in this embryo than in F,G.)

Fgfr1-Frs signaling is essential during neurulation and tail bud development

Thirty to 50% of Fgfr1ΔFrsFrs embryos isolated E10.5-11.5 exhibited a defect in spinal neural tube closure(Fig. 3A,C). This was never observed in Fgfr1wtKI/wtKI embryos, and differs from the neurulation phenotype observed in Fgfr1Δex4ex4 embryos in which the neural tube remains open along the entire rostrocaudal axis(craniorachischisis; Fig. 3B). The spinal neural tube (caudal to approximately somite 8) closes by the elevation and fusion of neural folds in a progressive rostral-to-caudal manner(Shum and Copp, 1996; Copp et al., 2003). In some Fgfr1ΔFrsFrs embryos, the spinal neural tube remained completely open at E10.5, which could indicate a mere delay in the closure process (Fig. 3C, right). In others, however, the spinal neural tube was closed at discrete points but open in the intervening regions(Fig. 3C, left). Later in development, some Fgfr1ΔFrsFrs embryos completed neural tube closure and others exhibited varying degrees of spina bifida (Fig. 3D-F, data not shown). In one embryo (E17.5), spina bifida was accompanied by loss of spinal cord marginal layer tissue and caudal meningomyelocele (extrusion of the spinal cord and meninges from the vertebral column; data not shown).

Fig. 3.

Neural tube defects. (A-C) Hoxb1 whole-mount in situ hybridization, labeling dorsal neural tube in the spinal region. Edges of open neuroepithelium are outlined in B. In C, the open cranial region (left embryo)is a procedural artifact. (D-F) Hematoxylin and Eosin stained sections through spinal neural tubes of E10.5 littermates. Embryo in E completed closure and embryo in F exhibited the intermittent neural tube closure phenotype. (G) Mitotic and apoptotic indices of neural tube sections(E10.5), assessed at similar spinal levels by phosphohistone H3 immunostaining and TUNEL. Each data point is the mean of positive/total cells on 12-16 sections per embryo, two embryos/genotype. Error bars indicate ±1 s.d.; P values were determined by Student's t-test.

Fig. 3.

Neural tube defects. (A-C) Hoxb1 whole-mount in situ hybridization, labeling dorsal neural tube in the spinal region. Edges of open neuroepithelium are outlined in B. In C, the open cranial region (left embryo)is a procedural artifact. (D-F) Hematoxylin and Eosin stained sections through spinal neural tubes of E10.5 littermates. Embryo in E completed closure and embryo in F exhibited the intermittent neural tube closure phenotype. (G) Mitotic and apoptotic indices of neural tube sections(E10.5), assessed at similar spinal levels by phosphohistone H3 immunostaining and TUNEL. Each data point is the mean of positive/total cells on 12-16 sections per embryo, two embryos/genotype. Error bars indicate ±1 s.d.; P values were determined by Student's t-test.

Histological analysis revealed altered morphology of the canal through mutant neural tubes in the spinal region. This was observed in mutant embryos that completed closure, as well as in those that retained regions of open neural tube (Fig. 3E,F). During neurulation, spinal neural folds bend towards the midline through the formation of neuroepithelial hingepoints; this morphogenetic process occurs by different mechanisms at different axial levels. In the upper- and mid-spinal regions, the neuroepithelium bends at one medial hingepoint and the lateral edges fold towards the dorsal midline, generating a slit-shaped neural canal(e.g. Fig. 3D). In the lower spinal region, two dorsolateral hingepoints (DLHP) form in addition to the medial hingepoint, resulting in a diamond-shaped canal(Shum and Copp, 1996). In Fgfr1ΔFrsFrsembryos, the canal is diamond shaped at all spinal levels(Fig. 3E,F), suggesting that this second morphogenetic mechanism is used to close the entire spinal neural tube. Cranial neuroepithelial folding appeared normal in all Fgfr1ΔFrsFrs embryos (data not shown).

Normally, notochordal Shh inhibits DLHP formation at mid- and upper spinal levels, while signals from the dorsal ectoderm are thought to play an inductive role (Copp et al.,2003). In Fgfr1ΔFrsFrsembryos, the notochord appears normal, extends through the spinal region, and expresses Shh (Fig. 2E, Fig. 3D-F). Although ectodermal-inducing signals have not yet been identified, it has been proposed that DLHP formation involves localized proliferation or apoptosis in the dorsolateral neuroepithelium (Copp et al.,2003). We analyzed proliferation and apoptosis in Fgfr1ΔFrsFrs neural tubes and found that although the mitotic index is not significantly altered in these embryos, they do exhibit an increase in cell death within the spinal neural tube (Fig. 3G). However,as apoptosis was not specifically localized to dorsolateral regions (data not shown), this probably reflects a role of Fgfr1-Frs signaling separate from its role in neurulation.

Neural tube closure defects are often associated with posterior truncations, and it has been postulated that the curvature of the body axis generated during tail bud outgrowth provides tension in the neural folds that facilitates closure (Copp et al.,1982; Copp, 1985; Gofflot et al., 1997; Peeters et al., 1998; Finnell et al., 2003). Over 80% of Fgfr1ΔFrsFrsembryos exhibit posterior truncations of varying severity(Fig. 4A-C, see Fig. S1A in the supplementary material). Truncations persist through embryogenesis and in severe cases are accompanied by a truncation of the notochord at the level of the hindlimb and failure to bifurcate caudal limb fields and organs(Fig. 4D,E, data not shown). In previous Fgfr1 reports, posterior truncations were associated with ectopic posterior neuroectoderm and suggested to be secondary to the mesodermal defects (Ciruna et al.,1997; Ciruna and Rossant,2001). However, in Fgfr1ΔFrsFrs embryos, we observed truncations in the absence of these other phenotypes(Fig. 2, data not shown),indicating that Fgfr1-Frs signaling directly impacts tail bud development.

All tissue layers of the tail bud form during secondary gastrulation from remnants of the primitive streak and node. In most Fgfr1ΔFrsFrs embryos, Shh in situ hybridization and Hematoxylin and Eosin staining of embryo sections indicated full extension of the notochord, and hence, correct positioning of tail bud progenitors, at the onset of secondary gastrulation(Fig. 2E, data not shown). We investigated whether the Frs site deletion directly affected patterning or outgrowth of the tail bud. Mutant analysis and expression studies have identified genes expressed and required in the tail bud during secondary gastrulation (Roelink and Nusse,1991; Takada et al.,1994; Greco et al.,1996; Gofflot et al.,1997; Copp et al.,2003). Except in the most severely truncated embryos, tail bud gene expression (T, Wnt3a, Fgf8, Hoxb1) and proliferation appeared normal in Fgfr1ΔFrsFrstail buds (Fig. 4F-I, data not shown). However, TUNEL analysis revealed an increase in cell death throughout caudal tissues of Fgfr1ΔFrsFrs embryos(Fig. 4I). Cell death may hinder cell migrations or reduce the number of tail bud progenitors during secondary gastrulation.

Fig. 4.

Posterior truncations and tail bud development. (A-C)Bright-field photographs of E13.5 littermates (caudal region).(D,E) Hematoxylin and Eosin stained sections through kidneys of wild-type and severely truncated Fgfr1ΔFrsFrs embryos; bar indicates relative magnification. (F) T (Brachyury) and(G,H) Wnt3a whole-mount in situ hybridization.(I) Quantitation of proliferation (pH3, P=0.33) and apoptosis(TUNEL, P=0.09) in E10.5 tail bud sections. Each data point represents the mean of three to eight sections through each of at least three embryos. Error bars indicate ±1 s.d.; P values were determined by Student's t-test.

Fig. 4.

Posterior truncations and tail bud development. (A-C)Bright-field photographs of E13.5 littermates (caudal region).(D,E) Hematoxylin and Eosin stained sections through kidneys of wild-type and severely truncated Fgfr1ΔFrsFrs embryos; bar indicates relative magnification. (F) T (Brachyury) and(G,H) Wnt3a whole-mount in situ hybridization.(I) Quantitation of proliferation (pH3, P=0.33) and apoptosis(TUNEL, P=0.09) in E10.5 tail bud sections. Each data point represents the mean of three to eight sections through each of at least three embryos. Error bars indicate ±1 s.d.; P values were determined by Student's t-test.

Fig. 5.

Pharyngeal arch (PA) development. (A-C) Phosphohistone H3 immunostained PA sections (E10.5). PA1 is labeled, arrowheads indicate PA2.(D,E) Fgf8 whole-mount in situ hybridization (PA region); arrows indicate PA2 Fgf8 expression domain lost in mutant embryos. (F-H) Migrating NCCs in the pharyngeal region of stage-matched embryos, visualized by Sox10 in situ hybridization. Arrows indicate limits of NCC migration into PA1, 2. (I-L) Nile Blue staining of cell death in the pharyngeal region (E10.5). I/J and K/L are stage-matched control/mutant embryo pairs.

Fig. 5.

Pharyngeal arch (PA) development. (A-C) Phosphohistone H3 immunostained PA sections (E10.5). PA1 is labeled, arrowheads indicate PA2.(D,E) Fgf8 whole-mount in situ hybridization (PA region); arrows indicate PA2 Fgf8 expression domain lost in mutant embryos. (F-H) Migrating NCCs in the pharyngeal region of stage-matched embryos, visualized by Sox10 in situ hybridization. Arrows indicate limits of NCC migration into PA1, 2. (I-L) Nile Blue staining of cell death in the pharyngeal region (E10.5). I/J and K/L are stage-matched control/mutant embryo pairs.

Fgfr1-Frs signaling in pharyngeal arch (PA) development

The second PA is hypoplastic in Fgfr1ΔFrsFrs embryos by E10.5, and histological analysis revealed a dramatic reduction in the number of mesenchymal cells populating this arch. This population normally includes both mesoderm and neural crest cells. Unlike cells within wild-type PAs,mesenchymal cells within mutant PA2 are neither proliferating nor closely apposed to the surrounding epithelia (Fig. 5A-C). We also observed a loss of Fgf8 expression in PA2,which has been shown to promote survival of PA mesenchyme(Fig. 5D,E)(Frank et al., 2002; Macatee et al., 2003). Although TUNEL analysis on sections did not reveal ectopic apoptosis of cells within Fgfr1ΔFrsFrs arches(data not shown), the loss of Fgf8 signals may contribute to the lack of PA2 mesenchymal proliferation or affect migration of neural crest cells into the arches of mutant embryos. In a previous study of Fgfr1 hypomorphs,PA2 was found to be hypoplastic because of a neural crest migration defect(Trokovic et al., 2003a). Likewise, we observed impaired neural crest cell migration into Fgfr1ΔFrsFrs PAs by Sox10 in situ hybridization (Fig. 5F-H) (Kuhlbrodt et al.,1998). Nile Blue staining revealed ectopic cell death along the path of neural crest migration (prior to PA entry; Fig. 5I-L).

At later stages, we noted some recovery of PA2 size and did not observe prominent craniofacial defects that would be expected if PA neural crest development failed entirely. We therefore analyzed the formation of PA crest derivatives to determine whether neural crest migration recovered later in development. Analysis of craniofacial skeletal elements at E15.5 indicated that PA2 NCC derivatives are selectively affected in Fgfr1ΔFrsFrs embryos. We did not observe major hypoplasia or malformation of the jaws (PA1-derived;data not shown) or PA1 derivatives in the middle ear(Fig. 6A-D). By contrast,cartilages derived from PA2, including the stapes and styloid process of the middle ear and the lesser horns of the hyoid, were missing or hypoplastic in mutant embryos (Fig. 6).

Fgf signaling responses in Fgfr1ΔFrsFrs primary cells

Phenotypic data demonstrated that Fgfr1ΔFrs is expressed and functional in vivo: we observed neither recapitulation of the null phenotype, as would be expected if the mutant receptor were non-functional,nor gain-of-function phenotypes indicative of ligand-independent activation. However, phenotypic data also indicated that Fgfr1 transduces Frs-independent signals in some developmental contexts. To investigate the nature of these signals, we derived mouse embryonic cells (MECs) from wild-type and mutant embryos, and used them in biochemical studies.

We first used immortalized MECs to verify that the mutant receptor signaled as expected based on previous reports. In these cells, we observed rapid and robust phosphorylation of Frs2 in wild type, but not Fgfr1ΔFrsFrs, cells stimulated with Fgf (see Fig. S2A, part i in the supplementary material). We also observed an Fgf-stimulated decrease in Frs2 protein level in wild-type but not mutant cells, concomitant with Frs2 phosphorylation (see Fig. 2A, part ii in the supplementary material). In MAPK response assays, mutant immortalized MECs exhibited reduced sensitivity to Fgf and drove weaker signaling responses as assessed by phosphorylation of MAPK (see Fig. S2B in the supplementary material). These data recapitulate findings of previous studies performed in immortalized Frs2-/- MECs(Hadari et al., 2001).

Fig. 6.

PA2 neural crest derivatives. Skeleton preparations of E15.5 middle ears (A,B,D) and hyoid cartilages (E,F). PA1, 2 derivatives are traced in C,G,H. PA1 neural crest derivatives, traced in blue: i, incus, m, malleus, ty, tympanic ring, mc,Meckel's cartilage. PA2 derivatives, traced in red, are hypoplastic or missing in mutants: (C) s, stapes, sp, styloid process; (G,H) lesser horns of hyoid. Ectopic cartilages were observed in mutant ears (arrows in B,D; yellow in C).

Fig. 6.

PA2 neural crest derivatives. Skeleton preparations of E15.5 middle ears (A,B,D) and hyoid cartilages (E,F). PA1, 2 derivatives are traced in C,G,H. PA1 neural crest derivatives, traced in blue: i, incus, m, malleus, ty, tympanic ring, mc,Meckel's cartilage. PA2 derivatives, traced in red, are hypoplastic or missing in mutants: (C) s, stapes, sp, styloid process; (G,H) lesser horns of hyoid. Ectopic cartilages were observed in mutant ears (arrows in B,D; yellow in C).

Cellular transformation may affect signal transduction and thus confound results of biochemical studies. Therefore, to approximate the in vivo situation, we re-examined Fgf responses in low passage (p2) non-immortalized MECs. Surprisingly, Fgfr1ΔFrsFrs and wild-type p2 MECs responded with indistinguishable MAPK activation profiles, both in terms of response duration and Fgf dose sensitivity(Fig. 7A). We did not observe a notable PI3K pathway response, assayed by Akt phosphorylation, in cells of either genotype (Fig. 7A). In addition, neither Crk nor PLCγ were tyrosine phosphorylated (i.e. activated) in p2 MECs following Fgf stimulation(Fig. 7B, data not shown). Thus, these pathways are probably not mediating the observed MAPK response in Fgfr1ΔFrsFrs cells. Furthermore, compared with wild-type cells, the Frs2 phosphorylation response was severely diminished in mutant cells despite comparable levels of total Frs2 (Fig. 7C). We expect that residual Frs activation in Fgfr1ΔFrsFrs cells occurs indirectly, as it has previously been shown by yeast two hybrid that the amino acid 407-433 deletion completely abolishes the interaction of the Fgfr1 cytoplasmic domain with Frs2 (Xu et al.,1998).

We next investigated whether signaling through Fgfr2 could compensate for the lack of Fgfr1-Frs signaling by activating Frs2 (low level) and MAPK in Fgfr1ΔFrsFrs MECs. Fgfr2 expression levels are not altered in Fgfr1ΔFrsFrs (p2) MECs or embryos, as might be expected if this receptor was upregulated to compensate for reduced Fgfr1 signaling (Fig. 7D,E). Furthermore, we analyzed the activation state (tyrosine phosphorylation) of Fgfr2 in p2 MECs and found that although both wild-type and mutant MECs had basal levels of tyrosine phosphorylated Fgfr2, the amount of active Fgfr2 was not increased in cells of either genotype in response to Fgf. Surprisingly, the basal (-Fgf) level of activated Fgfr2 was elevated in Fgfr1ΔFrsFrs MECs, and these cells responded to Fgf with a decrease in Fgfr2 phosphorylation(Fig. 7D). These data are not consistent with a role of Fgfr2 in compensatory activation of Frs and MAPK,and instead suggest that Fgfr1-Frs signaling transregulates Fgfr2 activity.

The MAPK signaling responses in p2 MECs did not directly translate into Fgf responsiveness in a proliferation assay: only two out of four mutant MEC cultures proliferated in response to Fgf, whereas all four responded with comparable MAPK and PI3K responses (Fig. 7A,F, data not shown). The growth curves did, however, reflect the relative phenotypic severity of embryos from which mutant cells were derived(data not shown). Variation in mutant MEC proliferative responses is probably due to differences in cell physiology or developmental staging between severely and mildly affected embryos. Nonetheless, the ability of two mutant lines to proliferate in response to Fgf indicates that Fgfr1-Frs signaling is not essential in directing this cellular response.

Impact of the Fgfr1ΔFrs mutation on MAPK signaling in vivo

Previous reports have implicated the MAPK pathway as an effector of Fgf signaling in vivo (Corson et al.,2003). Our primary MEC data support this model and suggest that the Fgfr1-Frs signaling is not required for Fgf-induced MAPK activation. To determine the physiological relevance of the MEC results, we examined MAPK activation in Fgfr1-dependent developmental contexts by whole-mount immunohistochemistry. First, we analyzed stage-matched embryos at E8.5-9.5,when the caudal region of the embryo is still undergoing gastrulation and more rostral axial levels are undergoing somitogenesis and neurulation. We found that phospho-MAPK staining in gastrulating mesoderm and somites was not altered in mutant embryos (Fig. 8A-F). In addition, we observed a thin line of phospho-MAPK staining along the medial edges of the neural folds, where neuroepithelial fusion takes place. This domain of MAPK activation was present in both mutant and control embryos (arrows, Fig. 8C,F), despite the disruption of neural tube closure in Fgfr1ΔFrsFrs embryos. Just after neurulation, when we observed an increase in cell death in mutant neural tubes, MAPK is not activated in wild-type or mutant neural tubes (data not shown). Thus, the cell survival role of Fgfr1-Frs signaling in the neural tube is probably facilitated by a different downstream pathway.

Fig. 7.

p2 MEC responses to Fgf. (A) pMAPK and pAkt responses of Fgfr1ΔFrsFrs and control MECs stimulated with Fgf, assayed over a dose titration (Ai, 5 minute stimulations) and a 1 hour timecourse (Aii). RasGAP, loading control.(B,C) Crk and Frs2 activation responses to Fgf (50 ng/ml aFgf, 5μg/ml heparin, 5′), assayed by IP-phosphotyrosine (4G10) western blot. Blots were stripped and reprobed with anti-Crk or anti-Frs2 (lower panels). *No antibody IP control. (D) Fgfr2 responses to Fgf(50 ng/ml aFgf, 5 μg/ml heparin, 5′), assayed by 4G10 IP-Fgfr2 western blot (upper panel). Total cell lysate Fgfr2 (middle panel) and actin(loading control, lower panel) western blots show relative protein levels. Matching arrowheads indicate bands of the same size recognized by different antibodies within B-D. (E) Relative Fgfr2 mRNA levels in homozygous embryos, determined by semi-quantitative RT-PCR. (F) Growth curves of p2 MECs derived from four mutant and two wild-type embryos cultured in DMEM, 4% FBS, 50 ng/ml aFgf.

Fig. 7.

p2 MEC responses to Fgf. (A) pMAPK and pAkt responses of Fgfr1ΔFrsFrs and control MECs stimulated with Fgf, assayed over a dose titration (Ai, 5 minute stimulations) and a 1 hour timecourse (Aii). RasGAP, loading control.(B,C) Crk and Frs2 activation responses to Fgf (50 ng/ml aFgf, 5μg/ml heparin, 5′), assayed by IP-phosphotyrosine (4G10) western blot. Blots were stripped and reprobed with anti-Crk or anti-Frs2 (lower panels). *No antibody IP control. (D) Fgfr2 responses to Fgf(50 ng/ml aFgf, 5 μg/ml heparin, 5′), assayed by 4G10 IP-Fgfr2 western blot (upper panel). Total cell lysate Fgfr2 (middle panel) and actin(loading control, lower panel) western blots show relative protein levels. Matching arrowheads indicate bands of the same size recognized by different antibodies within B-D. (E) Relative Fgfr2 mRNA levels in homozygous embryos, determined by semi-quantitative RT-PCR. (F) Growth curves of p2 MECs derived from four mutant and two wild-type embryos cultured in DMEM, 4% FBS, 50 ng/ml aFgf.

Later, during early PA2 development, phospho-MAPK immunostained dispersed cells throughout the pharyngeal arch region in both mutant and control embryos(Fig. 8G,J). As development progressed, phospho-MAPK became concentrated in control embryos at the rostral edge of PA1 and the caudal edge of PA2(Fig. 8H). In mutant embryos at this stage, although PA1 immunostaining appeared normal, intensification of phospho-MAPK staining was not observed in caudal PA2(Fig. 8K). This domain corresponds to the Fgf8 expression domain that is lost in mutants; however,the difference in MAPK activation observed in whole mount may be secondary to the failure of NCCs to populate this arch. Premigratory NCCs and NCCs within PA3 stained strongly with the phosphoMAPK antibody(Fig. 8I,L, data not shown).

Together, these results demonstrate that MAPK activity is not globally disrupted in Fgfr1-dependent contexts in Fgfr1ΔFrsFrsembryos. Thus,either Fgfr1 does not contribute significantly to cellular phosphoMAPK levels in these contexts, or Fgfr1 signaling to MAPK is not affected by the Frs site deletion.

Developmental roles of Fgfr1-Frs signaling

Our results demonstrate that Fgfr1-Frs signaling is essential during embryogenesis but is required only in a subset of Fgfr1-dependent contexts. Abrogation of the Fgfr1-Frs interaction does not disrupt Fgfr1 functions during gastrulation or somitogenesis, but does result in late embryonic lethality and defects in the development of the tail bud, neural tube, and pharyngeal arches. The specific array of phenotypes observed in Fgfr1ΔFrsFrs embryos is distinct from that observed in Fgfr1wtKI homo- or hemizygotes, previously described hypomorphs with reduced Fgfr1 levels, or embryos harboring an Fgfr1PLCγ binding site mutation (Table 2, see Fig. S1 in the supplementary material)(Partanen et al., 1998; Xu et al., 1999; Trokovic et al., 2003a). This suggests that Fgfr1ΔFrsFrsphenotypes are not general consequences of reduced Fgfr1 activity, but are indeed due to disrupted signaling through the Frs interaction site. Interestingly, activating mutations in Fgfr1 that cause human craniosynostosis syndromes also have context-specific effects restricted mainly to craniofacial and limb development(Passos-Bueno et al., 1999). Thus, developmental contexts differ in their requirements for specific signaling pathways downstream of Fgfr1, as well as for regulation of Fgfr1 activity.

Table 2.

Comparison of Fgfr1 mutant phenotypes

Fgfr1-dependent contextFgfr1redFgfr1PLCγFgfr1ΔFrsFgfr1wtKI
Gastrulation     
Somitogenesis     
Node regression    
Tail bud development  (x) 
Neural tube closure   
AP patterning   
PA NCC development  (x) 
Digit development  
Fgfr1-dependent contextFgfr1redFgfr1PLCγFgfr1ΔFrsFgfr1wtKI
Gastrulation     
Somitogenesis     
Node regression    
Tail bud development  (x) 
Neural tube closure   
AP patterning   
PA NCC development  (x) 
Digit development  

Fgfr1red, hypomorphic alleles with reduced Fgfr1 expression; X, disrupted in homozygous embryos; (x) disrupted in Fgfr1wtKI/wtKI embryos but at lower penetrance and/or severity than in Fgfr1ΔFrs/Δfrs embryos. See text for references.

Fig. 8.

Phospho-MAPK immunohistochemistry. Whole-mount phospho-MAPK immunohistochemistry of stage-matched embryos: (A-C) E8.5, (D-F)E9.5. High magnification views of (B,E) caudal gastrulating regions and (C,F)more rostral areas undergoing somitogenesis and neurulation. Arrowheads and arrows indicate phospho-MAPK-positive somites and neural folds, respectively.(G-L) G/J, H/K and I/L are stage-matched embryo pairs (E10.5).(G,H,J,K) Pharyngeal arch 1, 2 region; (I) dorsal and (L) lateral views of premigratory neural crest cells (arrows). Lateral view is shown in L because this embryo has an open neural tube.

Fig. 8.

Phospho-MAPK immunohistochemistry. Whole-mount phospho-MAPK immunohistochemistry of stage-matched embryos: (A-C) E8.5, (D-F)E9.5. High magnification views of (B,E) caudal gastrulating regions and (C,F)more rostral areas undergoing somitogenesis and neurulation. Arrowheads and arrows indicate phospho-MAPK-positive somites and neural folds, respectively.(G-L) G/J, H/K and I/L are stage-matched embryo pairs (E10.5).(G,H,J,K) Pharyngeal arch 1, 2 region; (I) dorsal and (L) lateral views of premigratory neural crest cells (arrows). Lateral view is shown in L because this embryo has an open neural tube.

Because Fgfr1ΔFrsFrsembryos exhibited rescue of Fgfr1null mesodermal phenotypes and early lethality, analysis of these signaling mutants enabled us to clarify later roles of Fgfr1 in neural tube and tail bud development. We thus identified a novel role of Fgfr1 in spinal neurulation. In contrast to craniorachischisis, which in Fgfr1Δex4ex4 embryos is probably secondary to mesodermal defects, Fgfr1ΔFrsFrs neural tube closure defects affect only the spinal region and are not accompanied by notochord or paraxial mesoderm defects. Our histological data demonstrate that Fgfr1-Frs signaling impacts the morphogenetic folding of the spinal neuroepithelium. Frs2 was previously implicated in convergent extension movements (Akagi et al., 2002),disruption of which can alter the size, shape and signaling properties of the notochord and neural plate, and thus result in neural tube closure defects. However, whereas previously described convergent extension mutants lack medial hinge points and exhibit notochord defects, Fgfr1ΔFrsFrsembryos form ectopic DHLPs and do not exhibit defects in mesoderm or medial bending,(Copp et al., 2003). Our results are therefore more consistent with a role of Fgfr1-Frs signaling in modulating neuroepithelial responsiveness to DLHP-inducing or -repressing signals.

The neural tube closure defect in Fgfr1ΔFrsFrs embryos may be exacerbated by the accompanying defect in tail bud development. Previously,posterior truncations in Fgfr1null animals were postulated to be secondary to mesoderm defects. However, the high penetrance of posterior truncations, together with the low incidence of mesodermal defects in Fgfr1ΔFrsFrs embryos suggest that Fgfr1 instead plays a direct, Frs-dependent role in tail bud development. The Frs site deletion did not affect patterning or proliferation in the tail bud during secondary gastrulation, but resulted in ectopic apoptosis throughout the caudal region of mutant embryos. Cell death could contribute to posterior truncations by depleting tail bud progenitors or hindering cell migration during secondary gastrulation.

In Fgfr1ΔFrsFrsembryos, Sox10-positive neural crest cells are hindered in their migration into the first and second pharyngeal arches, though at later stages,craniofacial defects in these embryos are restricted predominantly to PA2 NCC derivatives. This implies that the Sox10-positive population migrating towards PA1 recovers or is not essential for the development of some PA1 NCC derivatives. Specific disruption of PA2 and/or PA2 NCC development in Fgfr1ΔFrsFrs embryos contrasts with general Fgfr1 hypomorph phenotypes that affect NCC derivatives of both PA1 and PA2 (Trokovic et al., 2003a). Fgfr1-Frs signaling may be primarily required for PA2 patterning, or may participate in reciprocal signaling between the epithelium and NCCs that facilitates PA2 entry or survival of NCCs. In Fgfr1ΔFrsFrs embryos, an epithelial Fgf8 expression domain is lost in PA2, and the NCC migration defect is accompanied by an increase in cell death along the migration pathway. Preliminary results of conditional mutagenesis experiments indicate that ectopic cell death in the pharyngeal region reflects a NCC-nonautonomous effect of the Frs site deletion (R.V.H. and P.S.,unpublished).

Potential mechanisms of Fgfr1 signaling during development

It has long been presumed that Fgfrs signal primarily through the MAPK pathway, and previous biochemical studies implicated Frs adaptors as important mediators of MAPK activation downstream of Fgfr1. However, we have demonstrated that the Fgfr1-Frs interaction is dispensable for early developmental roles of Fgfr1 and for activation of the MAPK pathway,both in primary embryonic cells and in vivo. Wild-type and Fgfr1ΔFrsFrs primary cell responses to Fgf stimulation were indistinguishable in terms of both dose sensitivity and signaling kinetics, and MAPK phosphorylation appeared normal in several Fgfr1-dependent contexts in Fgfr1ΔFrsFrs embryos. Furthermore, proliferative responses of non-immortalized Fgfr1ΔFrsFrs cells suggest that this pathway is not essential in driving Fgf-induced proliferation,although cells from embryos with more severe phenotypes were growth impaired. The discrepancy between our biochemical data and those published previously probably reflects cell type-specific signaling mechanisms, effects of Frs2 disruption on receptors other than Fgfr1, or effects of cellular transformation on the ability of other pathways to signal downstream of Fgfr1.

Our work, as well as previous studies of Fgfr1-PLCγ signaling,demonstrates that individual signaling events downstream of Fgfr1 are required in a context-specific manner during development. In contexts where Fgfr1 functions are not compromised by the Frs site deletion, other pathways may drive cellular responses either independently of or additively with Frs adaptors. Aside from Frs2 and Frs3, only PLCγ, Crk, Grb14 and Sef have been shown to interact directly with activated Fgfr1, although Src, STAT and Shc are also activated downstream of Fgfr1 in some cell types(Mohammadi et al., 1991; Klint and Claesson-Welsh,1999; Larrson et al.,1999; Reilly et al.,2000; Tsang et al.,2002; Kovalenko et al.,2003). Sensitivity to the Frs site deletion could reflect the availability of alternate signaling pathways or the level of overall Fgfr1 signal required in a given context, rather than selective use of the Frs pathway. Thus, Fgfr1-Frs signaling may be used more broadly during development than is revealed by Fgfr1ΔFrsFrsphenotypes.

Fgfr1 and Fgfr2 can form heterodimers in vitro(Bellot et al., 1991). If such heterodimers were to form in vivo, Fgfr1ΔFrs signals could get shunted to the Frs pathway downstream of the Fgfr2 subunit in Fgfr1ΔFrsFrs embryos. This predicts that heterodimers would rescue Fgfr1ΔFrs function in contexts co-expressing Fgfr1 and Fgfr2, and that Fgfr2 would be activated by Fgf stimulation in primary cells. Our data do not support this model: Fgfr1ΔFrsFrsembryos exhibit defects in the tail bud and pharyngeal arches, where Fgfr1 and Fgfr2 are co-expressed, but not during gastrulation or somitogenesis when the receptors have distinct expression patterns(Orr-Urtreger et al., 1991; Yamaguchi et al., 1992; Walshe and Mason, 2000). Furthermore, we did not observe compensatory upregulation of Fgfr2mRNA in vivo, or ligand-dependent activation of Fgfr2 in primary cells. Instead, basal levels of active Fgfr2 are notably elevated in Fgfr1ΔFrsFrs cells,suggesting that Fgfr1-Frs signaling transregulates Fgfr2. This may coordinate Fgfr responses and/or serve a homeostatic role in vivo. Deregulation of Fgfr2 activity may contribute in a gain-of-function manner to Fgfr1ΔFrsFrs phenotypes. Resolution of this issue will require further biochemical analysis and crosses between Fgfr1ΔFrs and Fgfr2 mutant mouse lines.

We thank Janet Rossant for providing Fgfr1 genomic clones; Mitch Goldfarb for advice on the Frs site deletion; Philip Corrin, Jason Frazier and Marc Grenley for excellent technical assistance; and Raj Kapur, Henk Roelink and our laboratory colleagues for critical reading of the manuscript. This work was supported by an NSF predoctoral fellowship and the CMB Training Grant (NIH#GM07270) to R.V.H., and by grants HD 24875 and HD 25326 to P.S.

Akagi, K., Park, E. K., Mood, K. and Daar, I. O.(
2002
). Docking protein SNT1 is a critical mediator of fibroblast growth factor signaling during Xenopus embryonic development.
Dev. Dyn.
223
,
216
-228.
Arman, E., Haffner-Krausz, R., Chen, Y., Heath, J. K. and Lonai,P. (
1998
). Targeted disruption of fibroblast growth factor(FGF) receptor 2 suggests a role for FGF signaling in pregastrulation mammalian development.
Proc. Natl. Acad. Sci. USA
95
,
5082
-5087.
Bellot, F., Crumley, G., Kaplow, J. M., Schlessinger, J., Jaye,M. and Dionne, C. A. (
1991
). Ligand-induced transphosphorylation between different FGF receptors.
EMBO J.
10
,
2849
-2854.
Berghella, L., de Angelis, L., Coletta, M., Berarducci, B.,Sonnino, C., Salvatori, G., Anthonissen, C., Cooper, R., Butler-Browne, G. S.,Mouly, V. et al. (
1999
). Reversible immortalization of human myogenic cells by site-specific excision of a retrovirally transferred oncogene.
Hum. Gene Ther.
10
,
1607
-1617.
Böttcher, R. T. and Niehrs, C. (
2005
). Fibroblast growth factor signalling during early vertebrate development.
Endocr. Rev.
26
,
63
-77.
Chen, W. V. and Soriano, P. (
2003
). Gene trap mutagenesis in embryonic stem cells.
Methods Enzymol.
365
,
367
-386.
Ciruna, B. and Rossant, J. (
2001
). FGF signaling regulates mesoderm cell fate specification and morphogenetic movement at the primitive streak.
Dev. Cell
1
,
37
-49.
Ciruna, B., Schwartz, L., Harpal, K., Yamaguchi, T. P. and Rossant, J. (
1997
). Chimeric analysis of fibroblast growth factor receptor-1 (Fgfr1) function: a role for FGFR1 in morphogenetic movement through the primitive streak.
Development
124
,
2829
-2841.
Copp, A. J. (
1985
). Relationship between timing of posterior neuropore closure and development of spinal neural tube defects in mutant (curly tail) and nurmal mouse embryos in culture.
J. Embryol. Exp. Morphol.
88
,
39
-54.
Copp, A. J., Seller, M. J. and Polani, P. E.(
1982
). Neural tube development in mutant (curly tail) and normal mouse embryos: the timing of posterior neuropore closure in vivo and in vitro.
J. Embryol. Exp. Morphol.
69
,
151
-156.
Copp, A. J., Greene, N. D. E. and Murdoch, J. N.(
2003
). The genetic basis of mammalian neurulation.
Nat. Rev. Genet.
4
,
784
-793.
Corson, L. B., Yamanaka, Y., Lai, K. M. and Rossant, J.(
2003
). Spatial and temporal patterns of ERK signaling during mouse embryogenesis.
Development
130
,
4527
-4537.
Cross, M. J., Lu, L., Magnusson, P., Nyqvist, D., Holmqvist, K.,Welsh, M. and Claesson-Welsh, L. (
2002
). The Shb adaptor protein binds to tyrosine 766 in the FGFR-1 and regulates the Ras/MEK/MAPK pathway via FRS2 phosphorylation in endothelial cells.
Mol. Biol. Cell
13
,
2881
-2893.
Deng, C.-X., Wynshaw-Boris, A., Shen, M. M., Daugherty, C.,Ornitz, D. M. and Leder, P. (
1994
). Murine FGFR-1 is required for early postimplantation growth and axial organization.
Genes Dev.
8
,
3045
-3057.
Deng, C., Wynshaw-Boris, A., Zhou, F., Kuo, A. and Leder, P.(
1996
). Fibroblast growth factor receptor 3 is a negative regulator of bone growth.
Cell
84
,
911
-921.
Fambrough, D., McClure, K., Kazlauskas, A. and Lander, E. S.(
1999
). Diverse signaling pathways activated by growth factor receptors induce broadly overlapping, rather than independent, sets of genes.
Cell
97
,
727
-741.
Farley, F. W., Soriano, P., Steffen, L. S. and Dymecki, S. M. (
2000
). Widespread recombinase expression using FLPeR(flipper) mice.
Genesis
28
,
106
-110.
Finnell, R. H., Gould, A. and Spiegelstein, O.(
2003
). Pathobiology and genetics of neural tube defects.
Epilepsia
44
,
S14
-S23.
Frank, D. U., Fotheringham, L. K., Brewer, J. A., Muglia, L. J.,Tristani-Firouzi, M., Capecchi, M. R. and Moon, A. M. (
2002
). An Fgf8 mouse mutant phenocopies human 22q11 deletion syndrome.
Development
129
,
4591
-4603.
Gofflot, F., Hall, M. and Morriss-Kay, G. M.(
1997
). Genetic patterning of the developing mouse tail at the time of posterior neuropore closure.
Dev. Dyn.
210
,
431
-445.
Gotoh, N., Laks, S., Nakashima, M., Lax, I. and Schlessinger,J. (
2004
). FRS2 family docking proteins with overlapping roles in activation of MAP kinase have distinct spatial-temporal patterns of expression of their transcripts.
FEBS Lett.
564
,
14
-18.
Gotoh, N., Manova, K., Tanaka, S., Murohashi, M., Hadari, Y.,Lee, A., Hamada, Y., Hiroe, T., Ito, M., Kurihara, T. et al.(
2005
). The docking protein FRS2alpha is an essential component of multiple fibroblast growth factor responses during early mouse development.
Mol. Cell. Biol.
25
,
4105
-4116.
Greco, T. L., Takada, S., Newhouse, M. M., McMahon, J. A.,McMahon, A. P. and Camper, S. A. (
1996
). Analysis of the vestigial tail mutation demonstrates that Wnt-3a gene dosage regulates mouse axial development.
Genes Dev.
10
,
313
-324.
Hadari, Y. R., Gotoh, N., Kouhara, H., Lax, I. and Schlessinger,J. (
2001
). Critical role for the docking-protein FRS2αin FGF receptor-mediated signal transduction pathways.
Proc. Natl. Acad. Sci. USA
98
,
8578
-8583.
Hamilton, T. G., Klinghoffer, R. A., Corrin, P. D. and Soriano,P. (
2003
). Evolutionary divergence of platelet-derived growth factor alpha receptor signaling mechanisms.
Mol. Cell. Biol.
23
,
4013
-4025.
Hébert, J. M., Lin, M., Partanen, J., Rossant, J. and McConnell, S. K. (
2003
). FGF signaling through FGFR1 is required for olfactory bulb morphogenesis.
Development
130
,
1101
-1111.
Johnson, D. E. and Williams, L. T. (
1993
). Structural and functional diversity in the FGF receptor multigene family.
Adv. Cancer Res.
60
,
1
-41.
Klinghoffer, R. A., Mueting-Nelsen-P. F., Faerman, A., Shani, M. and Soriano, P. (
2001
). The two PDGFRs display conserved signaling in vivo despite divergent embryological functions.
Mol. Cell
7
,
343
-354.
Klinghoffer, R. A., Hamilton, T. G., Hoch, R. and Soriano,P. (
2002
). An allelic series at the PDGFαR locus indicates unequal contributions of distinct signaling pathways during development.
Dev. Cell
2
,
103
-113.
Klint, P. and Claesson-Welsh, L. (
1999
). Signal transduction by fibroblast growth factor receptors.
Front. Biosci.
4
,
D165
-D177.
Kouhara, H., Hadari, Y. R., Spivak-Kroizman, T., Schilling, J.,Bar-Sagi, D., Lax, I. and Schlessinger, J. (
1997
). A lipid-anchored Grb2-binding protein that links FGF-receptor activation to the Ras/MAPK pathway.
Cell
89
,
693
-702.
Kovalenko, D., Yang, X., Nadeau, R. J., Harkins, L. K. and Friesel, R. (
2003
). Sef inhibits fibroblast growth factor signaling by inhibiting FGFR1 tyrosine phosphorylation and subsequent ERK activation.
J. Biol. Chem.
278
,
14087
-14091.
Kuhlbrodt, K., Herbarth, B., Sock, E., Hermans-Borgmeyer, I. and Wegner, M. (
1998
). Sox10, a novel transcriptional modulator in glial cells.
J. Neurosci.
18
,
237
-250.
Larrson, H., Klint, P., Landgren, E. and Claesson-Welsh, L.(
1999
). Fibroblast growth factor receptor-1 mediated endothelial cell proliferation is dependent on the src homology (SH2/SH3)domain-containing adaptor protein Crk.
J. Biol. Chem.
274
,
25726
-25734.
Macatee, T. L., Hammond, B. P., Arenkiel, B. R., Francis, L.,Frank, D. U. and Moon, A. M. (
2003
). Ablation of specific expression domains reveals discrete functions of ectoderm- and endoderm-derived FGF8 during cardiovascular and pharyngeal development.
Development
130
,
6361
-6374.
Madhani, H. D. (
2001
). Accounting for specificity in receptor tyrosine kinase signaling.
Cell
106
,
9
-11.
Maina, F., Pante, G., Helmbacher, F., Andres, R., Porthin, A.,Davies, A. M., Ponzetto, C. and Klein, R. (
2001
). Coupling Met to specific pathways results in distinct developmental outcomes.
Mol. Cell
7
,
1293
-1306.
McDougall, K., Kubu, C., Verdi, J. M. and Meakin, S. O.(
2001
). Developmental expression patterns of the signaling adapters FRS-2 and FRS-3 during early embryogenesis.
Mech. Dev.
103
,
145
-148.
Mohammadi, M., Honegger, A. M., Rotin, D., Fischer, R., Bellot,F., Li, W., Dionne, C. A., Jaye, M., Rubinstein, M. and Schlessinger, J.(
1991
). A tyrosine-phosphorylated carboxy-terminal peptide of the fibroblast growth factor receptor (Flg) is a binding site for the SH2 domain of phospholipase C-γ1.
Mol. Cell. Biol.
11
,
5068
-5078.
Ong, S. H., Goh, K. C., Lim, Y. P., Low, B. C., Klint, P.,Claesson-Welsh, L., Cao, X., Tan, Y. H. and Guy, G. R.(
1996
). SUC1-associated neurotrophic factor (SNT) protein is a major FGF-stimulated tyrosine phosphorylated 90-kDa protein which binds to the SH2 domain of GRB2.
Biochem. Biophys. Res. Commun.
225
,
1021
-1026.
Ong, S. H., Guy, G. R., Hadari, Y. R., Laks, S., Gotoh, N.,Schlessinger, J. and Lax, I. (
2000
). FRS2 proteins recruit intracellular signaling pathways by binding to diverse targets on fibroblast growth factor and nerve growth factor receptors.
Mol. Cell. Biol.
20
,
979
-989.
Ornitz, D. M. (
2000
). FGFs, heparan sulfate and FGFRs: complex interactions essential for development.
BioEssays
22
,
108
-112.
Orr-Urtreger, A., Givol, D., Yayon, A., Yarden, Y. and Lonai,P. (
1991
). Developmental expression of two murine fibroblast growth factor receptors, flg and bek.
Development
113
,
1419
-1434.
Partanen, J., Schwartz, L. and Rossant, J.(
1998
). Opposite phenotypes of hypomorphic and Y766 phosphorylation site mutations reveal a function for Fgfr1 in anteroposterior patterning of mouse embryos.
Genes Dev.
12
,
2332
-2344.
Passos-Bueno, M. R., Wilcox, W. R., Jabs, E. W., Sertié,A. L., Alonso, L. G. and Kitoh, H. (
1999
). Clinical spectrum of fibroblast growth factor receptor mutations.
Hum. Mutat.
14
,
115
-125.
Peeters, M. C. E., Schutte, B., Lenders, M.-H. J. N., Hekking,J. W. M., Drukker, J. and van Straaten, H. W. M. (
1998
). Role of differential cell proliferation in the tail bud in aberrant mouse neurulation.
Dev. Dyn.
211
,
382
-389.
Pogulis, R. J., Vallejo, A. N. and Pease, L. R.(
2000
). Recombination and mutagenesis by overlap extension PCR. In
The Nucleic Acid Protocols Handbook
. (ed. R. Rapley), pp.
857
-873. Totowa (NJ): Humana Press.
Rabin, S. J., Cleghon, V. and Kaplan, D. R.(
1993
). SNT1 a differentiation-specific target of neurotrophic factor-induced tyrosine kinase activity in neurons and PC12 cells.
Mol. Cell. Biol.
13
,
2203
-2213.
Reilly, J. F., Mickey, G. and Maher, P. A.(
2000
). Association of fibroblast growth factor receptor 1 with the adaptor protein Grb14.
J. Biol. Chem.
275
,
7771
-7778.
Roelink, H. and Nusse, R. (
1991
). Expression of two members of the Wnt family during mouse development-restricted temporal and spatial patterns in the developing neural tube.
Genes Dev.
5
,
381
-388.
Schlessinger, J. (
2000
). Cell signaling by receptor tyrosine kinases.
Cell
103
,
211
-225.
Shum, A. S. W. and Copp, A. J. (
1996
). Regional differences in morphogenesis of the neuroepithelium suggest multiple mechanisms of spinal neurulation in the mouse.
Anat. Embryol.
194
,
65
-73.
Takada, S., Stark, K. L., Shea, M. J., Vassileva, G., McMahon,J. A. and McMahon, A. P. (
1994
). Wnt-3a regulates somite and tailbud formation in the mouse embryo.
Genes Dev.
8
,
174
-189.
Tallquist, M. D. and Soriano, P. (
2000
). Epiblast-restricted Cre expression in MORE mice: a tool to distinguish embryonic vs. extra-embryonic gene function.
Genesis
26
,
113
-115.
Tallquist, M. D., Klinghoffer, R. A., Heuchel, R.,Mueting-Nelsen, P. F., Corrin, P. D., Heldin, C.-H., Johnson, R. J. and Soriano, P. (
2000
). Retention of PDGFRβ function in mice in the absence of phosphatidylinositol-3′kinase and phospholipase Cγ signaling pathways.
Genes Dev.
14
,
3179
-3190.
Tallquist, M. D., French, W. J. and Soriano, P.(
2003
). Additive effects of PDGF receptor b signaling pathways in vascular smooth muscle cell development.
PLOS Biol.
1
,
288
-299.
Trokovic, N., Trokovic, R., Mai, P. and Partanen, J.(
2003a
). FGFR1 regulates patterning of the pharyngeal region.
Genes Dev.
17
,
141
-153.
Trokovic, R., Trokovic, N., Hernesniemi, S., Pirvola, U., Vogt Weisenhorn, D. M., Rossant, J., McMahon, A. P., Wurst, W. and Partanen, J.(
2003b
). FGFR1 is independently required in both developing mid-and hindbrain for sustained response to isthmic signals.
EMBO J.
22
,
1811
-1823.
Tropepe, V., Siblia, M., Ciruna, B. G., Rossant, J., Wagner, E. F. and van der Kooy, D. (
1999
). Distinct neural stem cells proliferate in response to EGF and FGF in the developing mouse telencephalon.
Dev. Biol.
208
,
166
-188.
Tsang, M., Friesel, R., Kudoh, T. and Dawid, I. B.(
2002
). Identification of Sef, a novel modulator of FGF signalling.
Nat. Cell Biol.
4
,
165
-169.
Valius, M., Bazenet, C. and Kauzlauskas, A.(
1993
). Tyrosines 1021 and 1009 are phosphorylation sites in the carboxy terminus of the platelet-derived growth factor receptor beta subunit and are required for binding of phospholipase Cg and a 64-kilodalton protein,respectively.
Mol. Cell. Biol.
13
,
133
-143.
Walshe, J. and Mason, I. (
2000
). Expression of FGFR1, FGFR2, and FGFR3 during early neural development in the chick embryo.
Mech. Dev.
90
,
103
-110.
Wang, J.-K., Xu, H., Li, H.-C. and Goldfarb, M.(
1996
). Broadly expressed SNT-like proteins link FGF receptor stimulation to activators of the Ras pathway.
Oncogene
13
,
721
-729.
Werner, S., Duan, D.-S. R., de Vries, C., Peters, K. G.,Johnson, D. E. and Williams, L. T. (
1992
). Differential splicing in the extracellular region of Fibroblast Growth Factor Receptor 1 generates receptor variants with different ligand-binding specificities.
Mol. Cell. Biol.
12
,
82
-88.
Xu, H., Lee, K. W. and Goldfarb, M. (
1998
). Novel recognition motif on fibroblast growth factor receptor mediates direct association and activation of SNT adapter proteins.
J. Biol. Chem.
273
,
17987
-17990.
Xu, X., Li, C., Takahashi, K., Slavkin, H. C., Shum, L. and Deng, C.-X. (
1999
). Murine fibroblast growth factor receptor 1α isoforms mediate node regression and are essential for posterior mesoderm development.
Dev. Biol.
208
,
293
-306.
Yamaguchi, T. P., Conlon, R. A. and Rossant, J.(
1992
). Expression of the fibroblast growth factor receptor FGFR-1/flg during gastrulation and segmentation in the mouse embryo.
Dev. Biol.
152
,
75
-88.
Yamaguchi, T. P., Harpal, K., Henkemeyer, M. and Rossant, J.(
1994
). fgfr-1 is required for embryonic growth and mesodermal patterning during mouse gastrulation.
Genes Dev.
8
,
3032
-3044.
Yan, G., Wang, F., Fukabori, Y., Sussman, D., Hou, J. and McKeehan, W. L. (
1992
). Expression and transforming activity of a variant of the heparin-binding fibroblast growth factor receptor (flg)gene resulting from splicing of the alpha exon at an alternate 3′-acceptor site.
Biochem. Biophys. Res. Comm.
183
,
423
-430.
Yu, C., Wang, F., Kan, M., Jin, C., Jones, R. B., Weinstein, M.,Deng, C.-X. and McKeehan, W. L. (
2000
). Elevated cholesterol metabolism and bile acid synthesis in mice lacking membrane tyrosine kinase receptor FGFR4.
J. Biol. Chem.
275
,
15482
-15489.