The Trk family of tyrosine kinases encodes receptors for nerve growth factor-related neurotrophins. Here we present a developmental expression study of trkC, which encodes a receptor for neurotrophin-3 (NT-3). Like the related genes, trk and trkB, trkC is expressed primarily in neural lineages although the pattern is complex and includes non-neuronal cells. Direct comparison with trk and trkB developmental expression patterns permits the following observations. (1) trkC is expressed in novel neural tissues where other Trk genes are silent. (2) Some tissues appear to coexpress trkB and trkC receptors in the embryo and in the adult. (3) trkC expression can be detected in the gastrulating embryo. These data provide insights into the role of Trk-family receptors and nerve growth factor-related neurotrophins during development and suggest that, in addition to regulating neuronal survival and differentiation, the neurotrophin/Trk receptor system may have broader physiological effects. Finally, interspecific mouse backcrosses have been used to map the location of each of the Trk genes on mouse chromosomes. Alignment with available chromosomal maps identify possible linkage between the Trk genes and known neurological mutations.

The Trk gene family encodes receptor tyrosine kinases (RTKs) that include trk, trkB and trkC (for review see: Chao, 1992). The best characterized of these genes is human trk, which was originally described as a dominantly acting oncogene (Martin-Zanca et al., 1986) and recognized to encode a RTK (Martin-Zanca et al., 1989). However, the normal biological role of gp145prototrk remained enigmatic until expression studies provided the first clues regarding possible function (Martin-Zanca et al., 1990). Biochemical analysis demonstrated that the gp140prototrk RTK interacts directly with Nerve Growth Factor (NGF; Kaplan et al., 1991a,b; Klein et al., 1991), is required for NGF function (Loeb et al., 1991) and forms part of the high affinity NGF receptor (Klein et al., 1991; Hempstead et al., 1991). Furthermore, the NGF-related neurotrophins have been identified as ligands for the Trk-related RTK-encoding genes trkB and trkC (Berkemeier et al., 1991; Ip et al 1992; Soppet et al., 1991; Squinto et al., 1991; Lamballe et al., 1991; Tsoulfas et al., 1993). The interaction between neurotrophins and their respective receptors stimulates receptor tyrosine kinase activity and elicits different biological responses depending on the cellular environment of receptor expression. For example, in the PC12 rat pheochromocytoma cell line, neurotrophic factors promote neuronal survival and differentiation upon binding to the different members of the Trk receptor family (Kaplan et al., 1991a; Loeb et al., 1991; Squinto et al., 1991; Tsoulfas et al., 1993; D. Soppet and L. F. P., unpublished data). In contrast, stimulation of Trk receptors by neurotrophins induces proliferation in transfected NIH-3T3 fibroblasts (Glass et al., 1991; Barbacid et al., 1991). Other examples of RTKs exhibiting pleiotropic effects depending on the particular site of expression are the let-23 and c-kit genes. In Caenorhabdi tis elegans, genetic evidence indicates that the let-23 gene, which encodes a putative RTK of the epidermal growth factor (EGF) receptor subfamily, can control two opposing pathways in vulval differentiation and functions in at least five others tissues (Aroian and Sternberg, 1991). In the mouse, the c-kit RTK acts differently in cells originating from diverse embryonic lineages such as primordial germ cells, melanocytes and hematopoietic stem cells (Geissler et al., 1981, 1988). Similarly, findings that NGF has specific effects in the rat seminiferous epithelium (Parvinen et al., 1992) and B-lymphocytes (Otten et al., 1989) suggest that, in addition to its neurotrophic actions, this factor and presumably its related neurotrophins, brain-derived neurotrophic factor (BDNF; Leibrock et al., 1989), neurotrophin-3 (NT-3; Enfors et al., 1990; Hohn et al., 1990; Jones and Reichardt, 1990; Kaisho et al., 1990; Maisonpierre et al., 1990a; Rosenthal et al., 1990), Xenopus NT4 (xNT-4, Hallböök et al., 1991; Ip et al., 1992) and human NT-5 (hNT-5; Berkemeier et al., 1991), may have broader physiological effects.

The rat trkC locus is complex, encoding multiple distinct receptors (Tsoulfas et al., 1993). These trkC isoforms bind NT-3, a widely expressed neurotrophin (Maisonpierre et al., 1990b), in chemical cross-linking and equilibrium binding analyses (Lamballe et al., 1991, Tsoulfas et al., 1993).

In this study, we have investigated the expression of trkC in the mouse embryo and adult CNS. We find trkC transcripts throughout the nervous system but also in non-neural tissues. Furthermore, we contrast the expression of the trkC gene with that of trk and trkB. Finally, we have mapped the chromosomal location of the three Trk loci and show that they are unlinked and map in the vicinity of previously existing neurological mutations. This information may lead to a direct association of Trk genes with the phenotypes elicited by known spontaneous mutations in the mouse.

RNA preparation and RNAse protection analysis

RNA was extracted using RNAzol (Cinna/Biotecx) following the manufacturer’s recommendations. RNAse protection experiments were performed as previously described (Tessarollo et al.,1992) using the RPA kit (Ambion). A genomic trkC-specific probe that spans 196 nucleotides of the extracellular domain (aa 336-401 of the rat sequence; Tsoulfas et al., 1993) and including 64 downstream intronic nucleotides was used to generate an antisense RNA probe employed in RNAse protection analysis. A 250 base β-actin c-RNA probe (Alonso et al., 1986) was included in the same reaction as a means of assessing relative levels of RNA present in each hybridization.

In situ hybridization

In situ hybridization protocols were as described (Martin-Zanca et al. 1990) with the following modifications. Dissected embryos were fixed overnight in 4% paraformaldehyde, dehydrated with alcohols and xylenes, and embedded in paraffin. Embryos were sectioned at 5 μm thickness and mounted on gelatin-coated slides. Slides were deparaffinized in xylene and rehydrated in graded (100-30%) ethanol solutions. After fixing in 4% paraformaldehyde, the tissues were pretreated with proteinase K (20 μg/ml) (Boehringer Mannheim), refixed and immersed in triethanolamine buffer containing acetic anydrate and dehydrated. Sections were hybridized with antisense cRNA probes (5×105 cts/minute) in a buffer containing 50% formamide, 0.3 M NaCl, 20 mM Tris-Cl (pH 7.4), 1× Denhardt’s solution, 0.5 mg/ml yeast tRNA and 10 mM DTT at 50°C for 20 hours. After hybridization, washes were performed in 4× SSC and 10 mM DTT at 50°C. The slides were then incubated for 30 minutes at 37°C with RNase A (20 μg/ml) and RNase T1 (2 μg/ml) followed by a 30 minute incubation at 55°C in 50% formamide, 0.2× SSC, 10 mM DTT, washed twice for 30 minutes in 0.2× SSC, 1% sodium pyrophosphate (w/v), 10 mM DTT and dehydrated. The slides were dipped in Kodak emulsion NTB-2 and exposed for up to 10 days at 4°C. The slides were then developed in Kodak D-19, fixed as recommended by manufacturer and stained in 0.2% toluidine blue. All photomicroscopy was done on a Zeiss Axiophot microscope. Sense and anti-sense probes labeled with 35S were prepared by standard procedures (Krieg and Melton, 1987) by using UTP as the labeled nucleotide. The trkC antisense probe was synthesized from a ∽2 kb rat trkC cDNA (NRT-8) as described by Tsoulfas et al. (1993). The trk and trkB probes are described in detail by Martin-Zanca et al. (1990) and Klein et al. (1989), respectively.

Interspecific backcross mapping

Interspecific backcross progeny were generated by mating (C57BL/6J × Mus spretus)F1 females and C57BL/6J males as described (Copeland and Jenkins, 1991). A total of 205 N2 progeny were obtained; a random subset of these N2 mice were used to map the Trk loci (see text for details). DNA isolation, restriction enzyme digestion, agarose gel electrophoresis, Southern blot transfer and hybridization were performed as described (Jenkins et al., 1982). All blots were prepared with Zetabind nylon membrane (AMF-Cuno). Hybridization probes were labeled with [α32P] dCTP using a nick translation labeling kit (IL-9) (Boehringer Mannheim) or random priming (trk, trkB, trkC) (Amersham Corporation). Washing was done to a final stringency of 0.2-0.8 × SSCP, 0.1% SDS, 65°C.

The three mouse Trk loci, trk, trkB and trkC, have been designated Ntrk1, Ntrk2 and Ntrk3, respectively, to conform with the current human locus designations for these genes. The trk probe was a 1.5 kb EcoRI genomic fragment from the mouse trk extracellular domain (Martin-Zanca et al., 1990). The trk probe hybridization to a 6.4 kb fragment in BglII-digested C57BL/6J (B6) DNA and 5.0 and 1.3 kb fragments in M. spretus (S) DNA. The two M. spretus-specific fragments cosegregated and were followed in backcross mice. The trkB probe was a 938 bp corresponding to amino acid 1-308 of mouse trkB sequences (Klein et al., 1989). The trkB probe hybridized to 3.8, 2.5, 1.1, 0.9 and 0.35 kb fragments in TaqI-digested B6 DNA and 3.5, 3.1, 2.5, 1.4, 1.1 and 0.35 kb fragments in S DNA. The 3.5, 3.1 and 1.4 kb M. spre tus-specific fragments cosegregated and were followed in backcross mice. The trkC probe was a 288 bp cDNA corresponding to amino acid 282-378 of the mouse trkC gene (Tsoulfas et al., 1993). The TrkC probe hybridized to a 1.3 kb fragment in EcoRIdigested B6 DNA and a 3.6 kb fragment in S DNA. The 1.3 kb M. spretus-specific fragment was followed in backcross mice.

A description of most of the probes and RFLPs for loci used to position the Trk loci on the interspecific map have been reported. The loci in include: fibrinogen, gamma polypeptide (Fgg), connexin-40 (Cnx-40) and nerve growth factor beta (Ngfb) on chromosome 3 (Mucenski et al., 1988; Cox et al., 1991; Haefliger et al., 1992); neuroendocrine convertase-1 (Nec-1) on chromosome 13 (Copeland et al., 1992); and insulin-like growth factor 1 receptor (Igflr), feline sarcoma oncogene (Fes) and tyrosinase (Tyr) on chromosome 7 (Lunsford et al., 1990; Copeland et al., 1992). One locus, interleukin-9 (IL-9) on chromosome 13 has not been previously reported for our panel. The IL-9 probe (pP40.2B4) was a 550 bp mouse cDNA (Van Snick et al., 1989). The IL-9 probe hybridized to an 18.0 kb fragment in SphI-digested B6 DNA and a 5.4 kb fragment in S DNA. The 5.4 kb M. spretus-specific fragment was followed in backcross mice.

Recombination distances were calculated as described (Green, 1981) using the computer program SPRETUS MADNESS. Gene order was determined by minimizing the number of recombination events required to explain the allele distribution patterns.

To determine the temporal expression pattern for trkC during murine development, a trkC-specific probe was used in RNAse protection assays employing a β-actin internal control as described in Materials and methods. 10 μg of total mouse embryo RNA were assayed and, although a faint protected band is first observed on day 7.5, trkC gene expression is clearly detected at embryonic day 8.5 coinciding with the timing of neural tube formation (Fig. 1). Relative to the internal β-actin control, a substantial increase in trkC mRNA expression occurs around embryonic day 10.5 (25–35 somites).

Fig. 1.

RNAse protection analysis of trkC expression in mouse embryos from 7.5 to 17.5 days of gestation. Approximately 10 μg of total RNA from embryonic trunks (T) and heads (H) were analyzed using a 32P-labeled trkCspecific probe (see Materials and methods). A β-actin probe was used in the same hybridization mixture with trkCas an internal control for RNA quantitation. Co-electrophoresed DNA size markers are indicated at the left.

Fig. 1.

RNAse protection analysis of trkC expression in mouse embryos from 7.5 to 17.5 days of gestation. Approximately 10 μg of total RNA from embryonic trunks (T) and heads (H) were analyzed using a 32P-labeled trkCspecific probe (see Materials and methods). A β-actin probe was used in the same hybridization mixture with trkCas an internal control for RNA quantitation. Co-electrophoresed DNA size markers are indicated at the left.

The trk and trkB genes are expressed predominantly in the nervous system (Martin-Zanca et al., 1990; Klein et al., 1990a). Having determined that trkC transcripts are present in the embryo from the earliest stages of neural induction, we next performed RNA in situ hybridization of postimplantation embryo sections from day 7.5 through day 17.5 and in neo-nate and adult central nervous system (CNS; see Materials and methods). For comparison, in the course of this study, we hybridized adjacent embryo sections with trk, trkB and trkC probes.

The results obtained from our in situ experiments on trkC expression are summarized in Table 1. The expression of this gene is highly complex. We first observed trkC hybridization as a weak signal in 7.5 day egg cylinders with a pattern indicating expression in the early neuroectoderm (not shown). Whole-mount in situ experiments are in progress to extend these observations. trkC gene expression is evident throughout the neuroepithelium of 9.5 (12–30 somites) day embryos (Fig. 2A-D). At these early stages, we also observe trkC transcripts in regions where neural crest cells are known to localize. These include forming dorsal root ganglia (DRG) and regions adjacent to the neural tube and the dorsal aorta (Fig. 2C,D; arrows), locations identified as migratory routes used by cells of neural crest origin (see below).

Table 1.

Summary of trkC expression

Summary of trkC expression
Summary of trkC expression
Fig. 2.

trkC expression in 9.5, 11.5 and 13.5 day embryos. (A,C) Dark-field and (B,D) light-field views of a sagittal (A,B) and a frontal (C,D) section of a 9.5 day embryo hybridized with a trkC-specific probe. (E) Dark-field and (F) light-field optics of an 11.5 day mouse embryo sagittal section. (G,H,I) Dark-field views of serial adjacent sections from a 13.5 day embryo. Arrows in A-D indicate the forming DRG and migratory neural crest cells whereas in E-I they indicate the DRG (single arrow) and the sympathetic (double arrows). n, neural tube; a, dorsal aorta; cerebellum (Cb), trigeminal (V) and vestibulo-cochlear (VIII) ganglia, stomach (g).

Fig. 2.

trkC expression in 9.5, 11.5 and 13.5 day embryos. (A,C) Dark-field and (B,D) light-field views of a sagittal (A,B) and a frontal (C,D) section of a 9.5 day embryo hybridized with a trkC-specific probe. (E) Dark-field and (F) light-field optics of an 11.5 day mouse embryo sagittal section. (G,H,I) Dark-field views of serial adjacent sections from a 13.5 day embryo. Arrows in A-D indicate the forming DRG and migratory neural crest cells whereas in E-I they indicate the DRG (single arrow) and the sympathetic (double arrows). n, neural tube; a, dorsal aorta; cerebellum (Cb), trigeminal (V) and vestibulo-cochlear (VIII) ganglia, stomach (g).

During organogenesis, additional sites of active trkC are found. Fig. 2E-F shows a sagittal section from an 11.5 day embryo (35 somites). trkC expression is exhibited throughout the CNS (telencephalon, diencephalon, mesencephalon, rhombencephalon and neural tube) and comparison with an adjacent trkB hybridized section (not shown) indicates overlapping expression in the mantle (postmitotic) layer of the spinal cord, whereas the ependymal (mitotic) layer expresses only trkB. The trkB and trkC genes are coexpressed in the branchial arch arteries; however, more caudal trkB transcripts are present in the mesonephric region while trkC is expressed in the midgut and genital ridge associated mesenchyme (see Fig. 2E-F).

At 13.5 days of development, the trkC gene is expressed throughout the neural network (Fig. 2G-I). At this stage, trkC mRNA is present in neural tissues where trk and trkB gene activity has not been previously detected (MartinZanca et al., 1990; Klein et al., 1989, 1990a). One striking example is the expression of trkC in neural-crest-derived cells of the developing enteric nervous system (Figs 2H, 7). In contrast, trkC and trkB transcripts are present in the trigeminal (Vth cranial) ganglion (Fig. 2H,I) where punctate hybridization reflects expression in a small subset of cells (Fig. 2I). The three afferent branches of the trigeminal ganglion (ophthalmic, maxillary and mandibular) can be readily traced by abundant trkC hybridization, presumably reflecting expression in the neural crest-derived Schwann cells that migrate and differentiate along the trigeminal processes (Fig. 2G).

Several insights can be drawn from a direct comparison of trk, trkB and trkC expression at later stages in development. Fig. 3 shows in situ hybridization of day 17.5 sections, from a single fetus, which were hybridized with probes to the three related genes. trk expression is distinctive in its tight regulation primarily limited to trigeminal (V), superior cervical (C), sympathetic (S) and dorsal root ganglia (arrows; Fig. 3A; Martin-Zanca et al., 1990). In comparison, trkB is expressed in these ganglia and at high levels in the CNS and additional structures including the whisker pad (W), tongue (T) and tooth buds (white arrowhead; Fig. 3B). This complex expression pattern is consistent throughout mid-gestation (days 14-18; L. F. P., unpublished observations). Expression of the trkC gene mirrors that of trkB in many regions of the embryo (Fig. 3B,C). Like the trkB gene (Klein et al 1990a; Parada et al., 1992), trkC is expressed throughout the nervous system including many of the same cranial and spinal ganglia (V, VII, VIII, X, DRG). The trkC and trkB genes are also apparently coexpressed in several non-neural structures including the tongue, the whisker pad (vibrissae) mesenchyme and forming teeth (Fig. 3B,C). In the tooth, trkB is expressed in the outer dermal papilla while trkC transcripts appear confined to the core of the dermal papilla (compare Fig. 3B and C). trkC transcripts are also found in several additional neural crest-derived structures including enteric ganglia (see Fig. 7) and in other non-neural cells (see below).

Fig. 3.

Comparative expression of Trk-family genes. Dark-field views of serial adjacent sagittal sections of a 17.5 day mouse embryo hybridized with a trk (A), trkB (B) or trkC (C)-specific probe. White triangles indicate the tooth papilla; arrows, DRG; Sr, striatum; t, thalamus; T, tongue; V, fifth cranial ganglion; C, superior cervical ganglion; S, sympathetic ganglia; W, whisker pad; i, intestine. The black triangles in panel A point to an artifact.

Fig. 3.

Comparative expression of Trk-family genes. Dark-field views of serial adjacent sagittal sections of a 17.5 day mouse embryo hybridized with a trk (A), trkB (B) or trkC (C)-specific probe. White triangles indicate the tooth papilla; arrows, DRG; Sr, striatum; t, thalamus; T, tongue; V, fifth cranial ganglion; C, superior cervical ganglion; S, sympathetic ganglia; W, whisker pad; i, intestine. The black triangles in panel A point to an artifact.

Embryonic CNS

The pattern of trkC gene expression is summarized in Table 1 and examples of in situ hybridization in the embryonic CNS are shown in Figs 2, 3C and 4. Interestingly, this gene shows both overlapping and exclusive expression profiles in the cephalic CNS when compared to the trkB gene. In the telencephalic cortical plate, both trkC and trkB transcripts are highly represented whereas in the intermediate layer only trkC mRNA was observed. In the ventricular layer, trkB expression is high and trkC expression is low (Figs 3B,C, 4A-D). In the striatum (Sr), trkC mRNA is most abundant centrally while trkB transcripts predominate along the ependymal layer. Finally, the thalamic nuclei (t) appear to express either trkB or trkC (Fig. 3).

A cross section through the caudal spinal cord of a 15.5 day embryo reveals trkC transcripts throughout the dorsalventral axis of the mantle layer including the ventral horn where motor neurons are located (Fig. 4E,F). No evidence of expression was seen in the floor plate at any developmental stages examined.

Fig. 4.

trkC expression in embryonic CNS. (A) Dark-field and (B) bright-field of a coronal section from 17.5 day embryonic brain.I Dark-field and (D) bright-field magnification of 17.5 day embryonic cortex. (E) Dark-field and (F) bright-field view of a section through the mid-lumbar region of a 15.5 day embryo spinal cord. cx, cortex; c, cortical layer; i, intermediate layer; v, ventricular layer; cp, caudate putamen; h, hippocampus; t, thalamus; m, ventral horn (motor neuron region of the spinal cord); f, floor plate.

Fig. 4.

trkC expression in embryonic CNS. (A) Dark-field and (B) bright-field of a coronal section from 17.5 day embryonic brain.I Dark-field and (D) bright-field magnification of 17.5 day embryonic cortex. (E) Dark-field and (F) bright-field view of a section through the mid-lumbar region of a 15.5 day embryo spinal cord. cx, cortex; c, cortical layer; i, intermediate layer; v, ventricular layer; cp, caudate putamen; h, hippocampus; t, thalamus; m, ventral horn (motor neuron region of the spinal cord); f, floor plate.

trkC transcripts are also abundant in the embryonic metencephalon (cerebellum). Fig. 5 provides examples of trkC expression in cerebellum at embryonic days 13.5, 17.5 and postnatal day 10. Transcripts are seen throughout embryonic stages with high levels in the immature but postmitotic Purkinje cells (Fig. 5A-D; Altman and Bayer, 1985). No expression was observed in the external germinal layer (egl). In all postnatal stages examined, including adult, trkC mRNA levels are very high in the cerebellar granule layers and in Purkinje neurons (Fig. 5E,F and not shown).

Fig. 5.

trkC expression in the developing mouse cerebellum. Dark-field (A,C) and bright-field (B,D) of a sagittal section through a 13.5 (A,B); 17.5 (C,D) mouse embryonic cerebellum; dark-field view (E) and bright-field photomicrograph (F) of 10 day postnatal mouse cerebellum sections hybridized with a trkC probe. Cb, cerebellum; g, granular layer; P, Purkinje cells; egl, external germinal layer.

Fig. 5.

trkC expression in the developing mouse cerebellum. Dark-field (A,C) and bright-field (B,D) of a sagittal section through a 13.5 (A,B); 17.5 (C,D) mouse embryonic cerebellum; dark-field view (E) and bright-field photomicrograph (F) of 10 day postnatal mouse cerebellum sections hybridized with a trkC probe. Cb, cerebellum; g, granular layer; P, Purkinje cells; egl, external germinal layer.

Peripheral nervous system

The neural crest gives rise to a large component of the peripheral nervous system (PNS). trkC transcripts are found in many PNS structures and also in cells that appear to be migratory, neural crest cells (Table 1, Fig. 2A,C,E, and unpublished results). As previously noted (Figs 2, 3), all Trk genes are expressed to a greater or lesser extent in neural crest-derived embryonic cranial, spinal and sympathetic ganglia. Fig. 6 shows in situ hybridization of three adjacent DRG sections from a 17.5 day embryo. These sections (5 μm) provide direct comparison of Trk gene expression and, in agreement with a report by Carroll and co-workers (1992), show that a majority of neurons express trk (Fig. 6A), while trkB (Fig. 6B) and trkC (Figs 6C, 2I) transcripts are present in a reduced subset of neurons. Very similar distribution of the three genes is seen in trigeminal ganglion (Fig. 3; see also Martin-Zanca et al., 1990; and Klein et al., 1990a).

Fig. 6.

Expression of the Trk gene family members in 17.5 embryonic DRG. Dark-field views of serial adjacent sections through a E17.5 DRG hybridized with a trk (A), trkB (B) and trkC (C)-specific probe.

Fig. 6.

Expression of the Trk gene family members in 17.5 embryonic DRG. Dark-field views of serial adjacent sections through a E17.5 DRG hybridized with a trk (A), trkB (B) and trkC (C)-specific probe.

The ganglia of the enteric nervous system are formed from migratory neural crest cells originating in the vagal crest region (Le Douarin and Teillet, 1973). These cells hybridize with trkC probes (Fig. 7A) as exemplified by hybridization of embryonic day 13.5 stomach where trkCexpressing migratory neural crest cells can be observed intercalated within the mesenchyme. Several days later, trkC transcripts can be localized to the forming ganglia that appear throughout the muscle and submucosal layers of the intestine (Fig. 7B). We do not observe trkC expression in the epithelium (Fig. 7). Thus, trkC expression is found in diverse PNS structures some of which do not appear to express other Trk-family receptors (Martin-Zanca et al., 1990; Klein et al., 1990a).

Fig. 7.

trkC expression in the developing gut. Bright-dark-field view of sections through E13.5 embryonic stomach (A) and E17.5 intestine (B). lu indicates the stomach (A) and intestine (B) lumen; arrows in panel A point at migratory neural crest cells; arrowheads (B) indicate the forming myenteric plexus.

Fig. 7.

trkC expression in the developing gut. Bright-dark-field view of sections through E13.5 embryonic stomach (A) and E17.5 intestine (B). lu indicates the stomach (A) and intestine (B) lumen; arrows in panel A point at migratory neural crest cells; arrowheads (B) indicate the forming myenteric plexus.

Adult CNS expression

trkC gene expression remains abundant in the adult mouse CNS. The cerebral cortex exhibits high levels of transcripts particularly in the superficial layers and in the deep layers with reduced expression in the intermediate layers (Fig. 8). TrkC transcripts are also abundant in the caudate putamen from embryonic stages (Fig. 4A,B), a pattern that remains unchanged in the adult CNS (Fig. 8B). Similarly, thalamic nuclei maintain trkC expression throughout development and in the adult (Fig. 8A,B). Thalamic nuclei do not appear to express trkB transcripts in adult brain (Klein et al., 1990a).

Fig. 8.

trkC expression in early postnatal and adult mouse brain. (a) Dark-field view of an horizontal section throughout a 10 day postnatal mouse brain. (B,C) Dark-field and bright-field views of coronal sections of 6 weeks old mouse brain hybridized with a trkC-specific probe. C, cortex; CP, caudatus putamen; F, fornix; Th, thalamic nuclei; CA1, CA2, CA3: stratum pyramidali; dg, dentate gyrus; me, medulla; Cb, cerebellum; hi, hippocampus; Hy, hypothalamus; O, olfactory nucleus; f, fimbria hippocampi.

Fig. 8.

trkC expression in early postnatal and adult mouse brain. (a) Dark-field view of an horizontal section throughout a 10 day postnatal mouse brain. (B,C) Dark-field and bright-field views of coronal sections of 6 weeks old mouse brain hybridized with a trkC-specific probe. C, cortex; CP, caudatus putamen; F, fornix; Th, thalamic nuclei; CA1, CA2, CA3: stratum pyramidali; dg, dentate gyrus; me, medulla; Cb, cerebellum; hi, hippocampus; Hy, hypothalamus; O, olfactory nucleus; f, fimbria hippocampi.

In the hippocampus of adult brain, the pyramidal cell layer as well as the neurons of the dentate gyrus exhibit high levels of trkC mRNA (Fig. 8A-C). In the cerebellum, both the Purkinje and the granule cell layer express high levels of trkC transcripts. Thus, expression of the trkC gene observed in the mature brain is consistent with that exhibited during embryonic development.

TrkC expression outside the nervous system

As can be appreciated from Figs 2 and 3, trkC-encoding transcripts are abundant throughout the embryo including sites that are not necessarily consistent with expression in neural cells (see Table 1). Note the presence of trkC transcripts in vibrissae and dental papillae (Fig. 3). We also find trkC transcripts in cells of the submandibular gland (Fig. 9A,B) and in the mesenchyme surrounding mesonephric and urogenital ducts (Fig. 9C,D). High levels of trkC transcripts can also be localized to the brown adipose tissue dorsal to the cervical spinal cord (Fig. 9D,E); the cortex of the metanephros; and adrenal gland (Fig. 9F,G). Finally, abundant trkC expression is found along the subendothelial mesenchyme of arteries throughout the developing embryo (Figs 9H,I and 2, 3).

Fig. 9.

trkC expression in non-neuronal E15.5 embryonic tissues. (A) Dark-field and (B) light-field of the submandibular sali vary gland. (C) Dark-field and (D) bright-field of a section through the genital region; s, sympathetic ganglia. (E) Dark-field and (F) light-field view of cervical brown adipose tissue. (G) Dark-field and (H) light-field of a section throughout the kidney and the adrenal gland (a). (I, J) Dark-field and light-field magnification of a sagittal section through the dorsal aorta.

Fig. 9.

trkC expression in non-neuronal E15.5 embryonic tissues. (A) Dark-field and (B) light-field of the submandibular sali vary gland. (C) Dark-field and (D) bright-field of a section through the genital region; s, sympathetic ganglia. (E) Dark-field and (F) light-field view of cervical brown adipose tissue. (G) Dark-field and (H) light-field of a section throughout the kidney and the adrenal gland (a). (I, J) Dark-field and light-field magnification of a sagittal section through the dorsal aorta.

Chromosomal mapping

We next wished to identify the chromosomal location of the three Trk genes to determine whether they are linked or map to separate chromosomal locations and to assess the possibility that these three genes might be associated with previously reported neurological mutations in the mouse. The chromosomal location of trk, trkB and trkC was determined by interspecific backcross analysis using progeny derived from matings of [(C57BL/6J × Mus spretus)F1 × C57BL/6J] mice that have been typed for over 1100 loci distributed among all mouse chromosomes (Copeland and Jenkins, 1991; see Materials and methods).

Each of the Trk genes mapped to a different mouse autosome (Fig. 10), indicating that they have become well dispersed during chromosome evolution. To determine whether any of the Trk genes mapped near a known mouse mutation with a phenotype that might be consistent with a defect in a Trk gene, we aligned our interspecific maps of chromosomes 3, 13 and 7 with composite linkage maps that report the map location of many uncloned mouse mutations (compiled from GBASE, a computerized database maintained at The Jackson Laboratory, Bar Harbor, ME). Interestingly, all three genes mapped near loci that exhibit neurological phenotypes.

Fig. 10.

Linkage maps showing the chromosomal locations of Ntrk loci in mouse. The Ntrk loci were mapped by interspecific backcross analysis. The number of recombinant N2 animals over the total number of animals typed plus the recombination frequencies, expressed as genetic distance in centimorgans (± 1 s.e.), is shown for each pair of loci on the left of the chromosome maps. Where no recombinants were found between loci, the upper 95% confidence limit of the recombination distance is given in parentheses. The position of loci in human chromosomes, determined in previously reported studies, are shown to the right of the chromosome maps. References for the human map positions of loci mapped in this study can be obtained from GDB (Genome Data Base), a computerized database of human linkage information maintained by The William H. Welch Medical Library of The Johns Hopkins University (Baltimore, MD).

Fig. 10.

Linkage maps showing the chromosomal locations of Ntrk loci in mouse. The Ntrk loci were mapped by interspecific backcross analysis. The number of recombinant N2 animals over the total number of animals typed plus the recombination frequencies, expressed as genetic distance in centimorgans (± 1 s.e.), is shown for each pair of loci on the left of the chromosome maps. Where no recombinants were found between loci, the upper 95% confidence limit of the recombination distance is given in parentheses. The position of loci in human chromosomes, determined in previously reported studies, are shown to the right of the chromosome maps. References for the human map positions of loci mapped in this study can be obtained from GDB (Genome Data Base), a computerized database of human linkage information maintained by The William H. Welch Medical Library of The Johns Hopkins University (Baltimore, MD).

trk mapped in the vicinity of the spontaneous neurological mutation spastic (spa) located on mouse chromosome 3 (Chai, 1961; Van Heyningen et al., 1975; Wilson et al., 1986; reviewed in Green, 1989). trkB mapped in the vicinity of the spontaneous neurological mutation Purkinje cell degeneration (pcd) located on chromosome 13 (Mullen et al., 1976; O’Gorman, 1985; reviewed in Green, 1989). Finally, trkC mapped in the vicinity of a gene on mouse chromosome 7 that has been shown to affect susceptibility of inbred mice to audiogenic seizures. (Neumann and Collins, 1991).

Recent studies from a number of laboratories, showing that trk and trkB are expressed mainly in the nervous system and encode functional receptors for the neurotrophins, have been important for advancing our understanding of the mechanisms that regulate neural development.

In the present study, we show that trkC, a third member of the Trk gene family (Lamballe et al., 1991; Tsoulfas et al., 1993), is expressed in a broad component of the nervous system and in non-neural tissues, including cell types where other Trk genes are not active.

trkC is expressed early in development

trkC expression was first observed in the gastrulating embryo, coordinate with the temporal expression pattern of its ligand NT-3, the more abundant neurotrophin detected early in development (Maisonpierre et al., 1990b). This is the earliest example of an active Trk gene reported to date, raising the possibility that the trkC RTK may mediate important functions in early neuroepithelium formation.

Neural crest

Kalcheim and co-workers (1992) have reported a proliferative activity of NT-3 in migratory neural crest cells. In the present study, we provide evidence for NT-3 receptor expression in migratory neural crest cells, thus supporting the notion that Trk receptors and their ligands may mediate alternative functions to trophic signals during development (for review, see Barbacid et al., 1991).

The trkC gene is expressed in many neural crest derivatives including cranial, dorsal root and sympathetic ganglia. In trigeminal and dorsal root ganglia, all Trk genes are active. trk mRNA is perhaps most abundant and confined to neurons in these ganglia. We note that in addition to expression in sympathetic ganglia late in development, trk expression can also be detected in isolated neurons of the basal forebrain, hindbrain and in other locations of the late embryonic and adult brain (Holtzman et al., 1992; L. F. P., unpublished observations). Like trkB, trkC is expressed in fewer neurons in sensory ganglia. However, in the trigeminal ganglion, the afferent projections are outlined by expression of these genes suggesting that trkB and trkC are expressed in neural crest-derived Schwann cells that migrate along the axonal projections to their sites of differentiation.

Another notable site of trkC-specific expression is the enteric neural crest and ganglia. Through trkC hybridization, it is possible to trace the migration of vagal neural crest cells into the gut region, suggesting that neurotrophins may have important functions in the migratory and trophic potential of enteric neurons.

trkC is also expressed in a group of non-neural tissues that, in quail-chick studies (Le Douarin, 1982) or in radioisotopic labeling experiments in the amphibian embryo (Chibon, 1964), have suggested neural crest origin. Among these are the artery walls derived from aortic arches, tooth papillae (Chibon, 1970) and salivary glands. Using RNAse protection and northern analyses, we also observe trkC transcripts in the thymus (Tsoulfas et al., 1993), another tissue whose connective tissue component is derived from the neural crest.

Brown adipose tissue, particularly in the cervical region dorsal to the spinal cord, also expresses high levels of trkC transcripts (Fig. 9E,F). Because of the high silver grain density caused by strong expression, we are unable to identify precisely the expressing cell types within this tissue. Electron microscopy studies indicate that brown adipose is particularly well innervated and Schwann cells have been observed sheathing unmyelinated nerve fibers in the lobules, close to the vessels, and between adipocytes (Bargmann et al., 1968; Linck et al., 1973). Therefore, trkC expression in this tissue may reflect the presence of transcripts in neural cells and not in adipocytes. However, in chick, the subcutaneous adipose of the face and the ventral neck region is derived from the neural crest (Le Douarin, 1982), while the precise origin of the brown adipose tissue has not been well established (Néchad, 1986). Thus, it is also possible that this adipose tissue may be of neural crest origin and retain trkC expression.

Trk genes are expressed in a broad spectrum of neural crest derivatives implying that neurotrophins may function in cells where activity has not previously been associated.

The trkB and trkC expression profiles exhibited in migratory neural crest cells along the descending aorta suggest that trkB and trkC products may represent good markers for subsets of neural-crest-derived cells.

CNS

In the CNS, the trkB and trkC genes are coexpressed and differentially expressed in intriguing patterns. We observed coexpression of these genes in the hippocampus (pyramidal and granule neurons), in the cerebellum (Purkinje cells), in the ventral spinal cord and in regions of the cerebral cortex (L. T. and L. F. P., unpublished data). In contrast, thalamic nuclei appear to express trkC transcripts preferentially in the adult. Similarly in the cerebellum, trkC expression is abundant in the granular layer where little or no trkB transcripts could be detected (see Klein et al., 1990b).

The neurotrophin theory has been well substantiated in the sensory nervous system through the study of NGFrelated neurotrophins and through the availability of welldefined primary culture systems (Barde, 1989). The search for trophic molecules acting on motor neurons has proven more elusive. Culturing of motor neurons is technically difficult and, although tissue extracts have been reported to promote motor neuron survival, only one molecule in particular, CNTF (Sendtner et al., 1992a) has been identified to promote survival of these cells in vivo. Until recently, efforts to identify a motor neuron response to the NGF family of neurotrophins have not been successful. Very recently BDNF has been shown to rescue motor neurons in vivo in rat and chick (Yan et al., 1992; Oppenheim et al., 1992; Sendtner et al., 1992b). The present results indicate the existence of trkC and trkB (Klein et al., 1990a) transcripts during embryogenesis in the ventral horn of the spinal cord and in other regions where motor neurons are localized. These results therefore suggest the need for a renewed effort at understanding neurotrophin function in motor neurons.

Chromosomal location

In mouse, the three Trk family members map to different mouse autosomes and are located in regions containing known neurological mutations. For example, the mouse trk gene maps near the spa mutation on chromosome 13. spa homozygotes show spastic symptoms although no anatomical abnormalities have been described. Glycine receptor deficiencies have been reported in these mice (White and Heller, 1982; White, 1985), but it remains to be determined whether these deficiencies are the cause or a secondary effect of the spa mutation.

Likewise, trkB maps on chromosome 3, in a region where the pcd mutation has previously been mapped. pcd homozygotes show a moderate ataxia beginning at 3 to 4 weeks of age. They are smaller than their normal littermates but live a fairly normal life span although adult males are infertile. In these mice, Purkinje cells begin to degenerate by 15 to 18 days of age followed by a slower degeneration of the photoreceptor cells of the retina and mitral cells of the olfactory bulb. Later (50-60 days of age) discrete populations of thalamic neurons also degenerate. Studies of fusion chimeras between pcd/pcd and +/+ embryos suggest that pcd acts directly within the Purkinje cells themselves (Mullen, 1977). The trkB gene is expressed in Purkinje cells, in retinal ganglion cells, in motor neurons and in a spectrum of other neural tissues. Thus the pcd phenotype could be accounted for by mutations in the trkB locus.

The trkC locus maps in a region of chromosome 7 that has been associated with increased tendency to exhibit auditory defects in mice (Neumann and Collins, 1991). These auditory defect susceptibilities are ill-defined at this time but it is possible that the trkC locus may contribute to such abnormalities since the NT-3 gene and its receptor (data not shown) is expressed in the sensory components of the developing ear during embryogenesis (Pirvola et al., 1992). As germline mutations are induced in the Trk genes via homologous recombination in embryonic stem cells, it should be possible to perform allelism studies to determine whether any of these mutations do in fact result from defects in Trk genes.

The present mouse mapping studies can also be used to predict where the Trk genes will map in humans. For example, trkB maps between Il-9 and Nec-1 which have both been mapped to the long arm of human chromosome 5 (Fig. 10), suggesting that trkB will reside on the long arm of human chromosome 5 as well. Likewise, trkC does not recombine with Fes, which has been mapped to 15q25-qter. Finally, trk has been mapped by two groups to the long arm of human chromosome 1 (Miozzo et al., 1990; Morris et al., 1991). This placement is consistent with the mouse mapping studies shown in Fig. 10.

We have studied the expression of the trkC gene during development and in the adult CNS, and compared these data with that of the other identified Trk gene family members (Martin-Zanca et al., 1990; Klein et al., 1990a). These genes map to unlinked chromosomal locations and are primarily, though not exclusively, expressed in the nervous system. Understanding the expression patterns of Trk genes provides important information regarding potential sites of neurotrophin action and identification of cells that preferentially coexpress combinations of Trk receptors or which uniquely express one specific receptor. These data suggest that experimental strategies must be defined to explore the physiological significance of Trk receptor coexpression and of their possible function in previously unidentified sites of activity including motor neurons and non-neural cells. Finally knowledge of the sites of Trk-gene expression will provide important insights for the analysis of mice that lack Trk-gene function due to mutations generated by homologous recombination in embryonic stem cells.

We thank Dan Soppet for the mouse trkB sequence and for many helpful discussions. We are grateful to the members of the Parada lab for their support and, in particular, to James Pickel and Dan Soppet for their advise and critical reading of the manuscript. We also thank B. Cho and M. Barnstead for technical assistance and Richard Fredrickson for his skillful assistance with artwork and Cindy Fitzpatrick for manuscript preparation. This research was supported by the National Cancer Institute, DHHS, under contract NO1-CO-74101 with ABL.

Alonso
,
S.
,
Minty
,
A.
,
Bourlet
,
Y.
and
Buckingham
,
M.
(
1986
).
Comparison of three actin-coding sequences in the mouse; evolutionary relationships between the actin genes of warm-blooded vertebrates
.
J. Molec. Evolution
23
,
11
22
.
Altman
,
J.
and
Bayer
,
S. A.
(
1985
).
Embryonic development of the rat cerebellum. I. Delineation of the cerebellar primordium and early cell movements
.
J. Comp. Neurol
.
231
,
1
6
.
Aroian
,
R. V.
and
Sternberg
,
P. W.
(
1991
).
Multiple functions of let-23, a Caenorhabditis elegans receptor tyrosine kinase gene required for vulval induction
.
Genetics
128
,
251
267
.
Barbacid
,
M.
,
Lamballe
,
F.
,
Pulido
,
D.
and
Klein
,
R.
(
1991
).
The trk family of tyrosine protein kinase receptors
.
BBA Rev. Cancer
1072
,
115127
.
Barde
,
Y.-A.
(
1989
).
Trophic factors and neuronal survival
.
Neuron
2
,
1525
1534
.
Bargmann
,
W.
,
von Hehn
,
G.
and
Lindner
,
E.
(
1968
).
Über die zellen des braunen fettgeund ihre innervation
.
Z. Zellforsch. Mikroski. Anat
.
85
,
601
.
Berkemeier
,
L. R.
,
Winslow
,
J. W.
,
Kaplan
,
D. R.
,
Nikolics
,
K.
,
Goeddel
,
D. V.
and
Rosenthal
,
A.
(
1991
).
Neurotrophin-5: A novel neurotrophic factor that activates trk and trkB
.
Neuron
7
,
857
.
Carroll
,
S. L.
,
Silos-Santiago
,
I.
,
Frese
S. E.
,
Ruit
,
K. G.
,
Milbrandt
,
J.
and
Snider
,
W. D.
(
1992
).
Dorsal root ganglion neurons expressing trk are selectively sensitive to NGF deprivation in utero
.
Neuron
9
,
779
788
.
Chai
,
C. K.
(
1961
).
Hereditary spasticity in mice
.
J. Hered
.
52
,
241
243
.
Chao
,
M. V.
(
1992
).
Neurotrophin receptors: a window into neuronal differentiation
.
Neuron
9
,
583
593
.
Chibon
,
P.
(
1964
).
Analyse par la méthode de marquage nucléaire a la thymidine tritiée des dérivés de la crête neurale céphalique chez l’Urodèle Pleurodeles waltlii Michah
.
C.R. Acad. Sci. Paris Ser. III
159
,
3624
3627
.
Chibon
,
P.
(
1970
).
L’origine de l’organe adamantin des dents. Etude aumoyen du marquage nucleaire de l’ectoderme stomodeal
.
Ann. Embryol. Morphog
.
3
,
203
213
.
Copeland
,
N. G.
and
Jenkins
,
N. A.
(
1991
).
Development and applications of a molecular genetic linkage map of the mouse genome
.
Trends Genet
.
7
,
113
118
.
Copeland
,
N. G.
,
Gilbert
,
D. J.
,
Chretien
,
M.
,
Seidah
,
N. G.
and
Jenkins
,
N. A.
(
1992
).
Regional localization of three convertases, PC1 (Nec-1), PC2 (Nec-2), and Furin (Fur), on mouse chromosomes
.
Genomics
13
,
1356
1358
.
Cox
,
R. D.
,
Copeland
,
N. G.
,
Jenkins
,
N. A.
and
Lehrach
,
H.
(
1991
).
Interspersed repetitive element polymerase chain reaction product mapping using a mouse interspecific backcross
.
Genomics
10
,
375
384
.
Ernfors
,
P.
,
Ibáñez
,
C. F.
,
Ebendal
,
T.
,
Olson
,
L.
and
Persson
,
H.
(
1990
).
Molecular cloning and neurotrophic activities of a protein with structural similarities to nerve growth factor: Developmental and topographical expression in the brain
.
Proc. Natl. Acad. Sci. USA
87
,
5454
.
Geissler
,
E. N.
,
McFarland
,
E. C.
and
Russel
,
E. S.
(
1981
).
Analysis of pleiotropism at the dominant white-spotting (W) locus of the house mouse: a description of ten new W alleles
.
Genetics
97
,
337
361
.
Geissler
,
E. N.
,
Ryan
,
M. A.
and
Housman
,
D. E.
(
1988
).
The dominant white-spotting (W) locus of the mouse encodes the c-kit proto-oncogene
.
Cell
55
,
185
192
.
Glass
,
D. J.
,
Nye
,
S. H.
,
Hantzopoulos
,
P.
,
Macchi
,
M. J.
,
Squinto
,
S. P.
,
Goldfarb
,
M.
and
Yancopoulos
,
G. D.
(
1991
).
TrkB mediates BDNF/NT-3-dependent survival and proliferation in fibroblasts lacking the low affinity NGF receptor
.
Cell
66
,
405
413
.
Green
,
E. L.
(
1981
).
Linkage, recombination and mapping
. In
Genetics and Probability in Animal Breeding Experiments
, pp.
77
113
.
New York
:
Oxford University Press
.
Green
,
M. C.
(
1989
).
Catalog of mutant genes and polymorphic loci
. In
Genetic Variants and Strains of the Laboratory Mouse
(ed.
M. F.
Lyon
and
A. G.
Searle
), pp.
12
403
.
New York
:
Oxford University Press
.
Haefliger
,
J.-A.
,
Bruzzone
,
R.
,
Jenkins
,
N. A.
,
Gilbert
,
D. J.
,
Copeland
,
N. G.
and
Paul
,
D. L.
(
1992
).
Four novel members of the connexin family of gap junction proteins
.
J. Biol. Chem
.
267
,
2057
2064
.
Hallböök
,
F.
,
Ibáñez
,
C. F.
and
Persson
,
H.
(
1991
).
Evolutionary studies of the nerve growth factor family reveal a novel member abundantly expressed in xenopus ovary
.
Neuron
6
,
845
.
Hempstead
,
B.
,
Kaplan
,
D. R.
,
Martin-Zanca
,
D.
,
Parada
,
L. F.
and
Chao
,
M. V.
(
1991
).
High affinity NGF binding requires co-expression of the trk proto-oncogene product and the low affinity NGF receptor
.
Nature
350
,
678
.
Hohn
,
A.
,
Leibrock
,
J.
,
Bailey
,
K.
and
Barde
,
Y.-A.
(
1990
).
Identification and characterization of a novel member of the nerve growth factor brainderived neurotrophic factor family
.
Nature
344
,
339
.
Holtzman
,
D. M.
,
Li
,
Y.
,
Parada
,
L. F.
,
Kinsman
,
S.
,
Chen
,
C.-K.
,
Valletta
,
J. S.
,
Zhou
,
J.
,
Long
,
J. B.
and
Mobley
,
W. C.
(
1992
).
p140trk mRNA marks NGF responsive forebrain neurons: evidence that trk gene expression is induced by NGF
.
Neuron
9
,
465
478
.
Ip
,
N. Y.
,
Ibáñez
,
C. F.
,
Nye
,
S. H.
,
McClain
,
J.
,
Jones
,
P. F.
,
Gies
,
D. R.
,
Belluscio
,
L.
,
LeBeau
,
M. M.
,
Espinosa III
,
R.
,
Squinto
,
S. P.
,
Persson
,
H.
,
Yancopoulos
,
G. D.
(
1992
).
Mammalian neurotrophin-4: Structure, chromosomal localization, tissue distribution, and receptor specificity
.
Proc. Natl. Acad. Sci. USA
89
,
3060
.
Jenkins
,
N. A.
,
Copeland
,
N. G.
,
Taylor
,
B. A.
and
Lee
,
B. K.
(
1982
).
Organization, distribution, and stability of endogenous ecotropic murine leukemia virus DNA sequences in chromosomes of Mus musculus
.
J. Virol
.
43
,
26
36
.
Jones
,
K. R.
and
Reichardt
,
L. F.
(
1990
).
Molecular cloning of the human gene that is a member of the nerve growth factor family
.
Proc. Natl. Acad. Sci. USA
87
,
8060
.
Kaisho
,
Y.
,
Yashimura
,
K.
and
Nakahama
,
K.
(
1990
).
Cloning and expression of a cDNA encoding a novel human neurotrophic factor
.
FEBS Lett
.
266
,
187
.
Kalcheim
,
C.
,
Carmeli
,
C.
and
Rosenthal
,
A.
(
1992
).
Neurotrophin 3 is a mitogen for cultured neural crest cells
.
Proc. Natl. Acad. Sci. USA
89
,
1661
1665
.
Kaplan
,
D. R.
,
Martin-Zanca
,
D.
and
Parada
,
L. F.
(
1991a
).
Tyrosine phosphorylation and tyrosine kinase activity of the trk protooncogene product induced by NGF
.
Nature
350
,
158
.
Kaplan
,
D. R.
,
Hempstead
,
B. L.
,
Martin-Zanca
,
D.
,
Chao
,
M. V.
and
Parada
,
L. F.
(
1991b
).
The trk proto-oncogene product: A signal transducing receptor for nerve growth factor
.
Science
252
,
554
.
Klein
,
R.
,
Parada
,
L. F.
,
Coulier
,
F.
and
Barbacid
,
M.
(
1989
).
TrkB, a novel tyrosine protein kinase receptor expressed during mouse neural development
.
EMBO J
.
8
,
3701
.
Klein
,
R.
,
Martin-Zanca
,
D.
,
Barbacid
,
M.
and
Parada
,
L. F.
(
1990a
).
Expression of the tyrosine kinase receptor gene trkB is confined to the murine embryonic and adult nervous system
.
Development
109
,
845
.
Klein
,
R.
,
Conway
,
D.
,
Parada
,
L. F.
and
Barbacid
,
M.
(
1990b
).
The trkB tyrosine kinase gene codes for a second neurogenic receptor that lacks the catalytic domain
.
Cell
61
,
647
.
Klein
,
R.
,
Jing
,
S.
,
Nanduri
,
V.
,
O’Rourke
,
E.
and
Barbacid
,
M.
(
1991
).
The Trk proto-oncogene encodes a receptor for nerve growth factor
.
Cell
65
,
189
.
Krieg
,
P. A.
and
Melton
,
D. A.
(
1987
).
In vitro RNA synthesis with SP6 RNA polymerase
.
Methods Enzymol
.
155
,
397
405
.
Lamballe
,
F.
,
Klein
,
R.
and
Barbacid
,
M.
(
1991
).
trkC: a new member of the trk family of tyrosine protein kinases, is a receptor for neurotrophin-3
.
Cell
66
,
967
.
Le Douarin
,
N. M.
(
1982
).
The Neural Crest
, pp
54
107
.
Cambridge Univ. Press
.
Le Douarin
,
N. M.
and
Teillet
,
M. A.
(
1973
).
The migration of neural crest cells to the wall of the digestive tract in avian embryo
.
J. Embryol. Exp. Morph
.
30
,
31
48
.
Leibrock
,
J.
,
Lottspeich
,
F.
,
Hohn
,
A.
,
Hofer
,
M.
,
Hengerer
,
B.
,
Masiakowski
,
P.
,
Thoenen
,
H.
and
Barde
,
Y.-A.
(
1989
).
Molecular cloning and expression of brain-derived neurotrophic factor
.
Nature
341
,
149
.
Linck
,
G.
,
Stoeckel
,
M. E.
,
Porte
,
A.
and
Petrovic
A.
(
1973
).
An electron microscope study of the specialized cell contacts and innervation of adipocyte in the brown fat of the european hamster (Cricetus cricetus)
.
Cytobiologie
7
,
431
.
Loeb
,
D. M.
,
Maragos
,
J.
,
Martin-Zanca
,
D.
,
Chao
,
M. V.
,
Parada
,
L. F.
and
Greene
,
L. A.
(
1991
).
The trk proto-oncogene rescues NGF responsiveness in mutant NGF-nonresponsive PC12 cell lines
.
Cell
66
,
961
.
Lunsford
,
R. D.
,
Jenkins
,
N. A.
,
Kozak
,
C. A.
,
Liang
,
L-F.
,
Silan
,
C. M.
,
Copeland
,
N. G.
and
Dean
,
J.
(
1990
).
Genomic mapping of murine Zp-2 and Zp-3, two oocyte-specific loci encoding zona pellucida proteins
.
Genomics
6
,
184
187
.
Maisonpierre
,
P. C.
,
Belluscio
,
L.
,
Squinto
,
S.
,
Ip
,
N. Y.
,
Furth
,
M. E.
,
Lindsay
,
R. M.
and
Yancopoulos
,
G. D.
(
1990a
).
A neurotrophin-3: A neurotrophic factor related to NGF and BDNF
.
Science
247
,
1446
.
Maisonpierre
,
P. C.
,
Belluscio
,
L.
,
Friedman
,
B.
,
Alderson
,
R. F.
,
Wiegand
,
S. J.
,
Furth
,
M. E.
,
Lindsay
,
R. M.
and
Yancopoulos
,
G. D.
(
1990b
).
NT-3, BDNF, and NGF in the developing rat nervous system: parallel as well as reciprocal patterns of expressions
.
Neuron
5
,
51015109
.
Martin-Zanca
,
D.
,
Hughes
,
S. H.
and
Barbacid
,
M.
(
1986
).
A human oncogene formed by the fusion of truncated tropomyosin and protein tyrosine kinase sequences
.
Nature
319
,
743
748
.
Martin-Zanca
,
D.
,
Oskam
,
R.
,
Mitra
,
G.
,
Copeland
,
T.
and Barbacid,M
. (
1989
).
Molecular and biochemical characterization of the human trk proto-oncogene
.
Mol. Cell. Biol
.
9
,
24
.
Martin-Zanca
,
D.
,
Barbacid
,
M.
and
Parada
,
L. F.
(
1990
).
Expression of the trk proto-oncogene is restricted to the sensory cranial and spinal ganglia of neural crest origin in mouse development
.
Genes Dev
.
4
,
683
.
Miozzo
,
M.
,
Pierotti
,
M. A.
,
Sozzi
,
G.
,
Radie
,
P.
,
Bongarzone
,
I.
,
Spurr
,
N. K.
and
Della Porta
,
G.
(
1990
).
Human TRK proto-oncogene maps to chromosome 1q32-q41
.
Oncogene
5
,
1411
1414
.
Morris
,
C. M.
,
Hao
,
Q. L.
,
Heisterkamp
,
N.
,
Fitzgerald
,
P. H.
and
Groffen
,
J.
(
1991
).
Localization of the TRK proto-oncogene to human chromosome bands 1q23-1q24
.
Oncogene
6
,
1093
1095
.
Mucenski
,
M. L.
,
Taylor
,
B. A.
,
Copeland
,
N. G.
and
Jenkins
,
N. A.
(
1988
).
Chromosomal location of Evi-1, a common site of ecotropic viral integration in AKXD murine myeloid tumors
.
Oncogene Res
.
2
,
219
233
.
Mullen
,
R. J.
(
1977
).
Site of pcd gene action and Purkinje cell mosaicism in cerebella of chimeric mice
.
Nature
270
,
245
247
.
Mullen
,
R. J.
,
Eicher
,
E. M.
and
Sidman
,
R. L.
(
1976
).
Purkinje cell degeneration, a new neurological mutation in the mouse
.
Proc. Natl. Acad. Sci. USA
73
,
208
212
.
Néchad
,
M.
(
1986
).
Structure and development of the brown adipose tissue
. In
Brown adipose tissue
(ed.
P.
Trayhurn
and
D. G.
Nicholls
), pp.
1
-
30
. London: Arnold.
Neumann
,
P. E.
and
Collins
,
R. L.
(
1991
).
Genetic dissection of susceptibility to audiogenic seizures in inbred mice
.
Proc. Natl. Acad. Sci. USA
88
,
5408
5412
.
O’Gorman
,
S.
(
1985
).
Degeneration of thalamic nuclei in ‘Purkinje cell degeneration’ mutant mice. II. Cytology of neuron loss
.
J. Comp. Neurol
.
234
,
298
316
.
Oppenheim
,
R. W.
,
Qin-Wei
,
Y.
,
Prevette
,
D.
and
Yan
,
Q.
(
1992
).
Brainderived nerotrophic factor rescues developing avian motoneurons from cell death
.
Nature
360
,
755
757
.
Otten
U.
,
Ehrhard
P.
and
Peck
,
R.
(
1989
).
Nerve growth factor induces growth and differentiation of human B lymphocytes
.
Proc. Natl. Acad. Sci. USA
86
,
10059
10063
.
Parada
,
L. F.
,
Tsoulfas
,
P.
,
Tessarollo
,
L.
,
Blair
,
J.
,
Reid
,
S.
and
Soppet
,
D.
(
1992
).
The Trk family of tyrosine kinases: receptors for NGF-related neurotrophins. In CSH Series
.
Cold Spring Harbor Symp. Quant. Biol
.
57
,
43
51
.
Parvinen
,
M.
,
Pelto-Huikko
,
M.
,
Soder
,
O.
,
Schultz
,
R.
,
Kaipia
,
A.
,
Mali
,
P.
,
Toppari
,
J.
,
Hakovirta
,
H.
,
Lönnerberg
,
P.
,
Ritzén
,
E.M.
,
Ebendal
,
T.
,
Olson
,
L.
,
Hökfelt
,
T.
and
Persson
,
H.
(
1992
).
Expression of β-Nerve Growth Factor and its receptor in rat seminiferous epithelium: specific function at the onset of meiosis
.
J. Cell Biol
.
117
,
629
641
.
Pirvola
,
U.
,
Ylikoski
,
J.
,
Palgi
,
J.
,
Lehtonen
,
E.
,
Arumäe
,
U.
and
Saarma
,
M.
(
1992
).
Brain-derived neurotrophic factor and neurotrophin 3 mRNAs in the peripheral target fields of developing inner ear ganglia
.
Proc. Natl. Acad. Sci. USA
89
,
9915
9919
.
Rosenthal
,
A.
,
Goeddel
,
D. V.
,
Nyugen
,
T.
,
Lewis
,
M.
,
Shih
,
A.
,
Laramee
,
G. R.
,
Nikolics
,
K.
and
Winslow
,
J. W.
(
1990
).
Primary structure and biological activity of a novel human neurotrophic factor
.
Neuron
4
,
767
773
.
Sendtner
,
M.
,
Schmalbruch
,
H.
Stockli
,
K. A.
,
Carrol
,
P.
,
Kreutzberg
,
G. W.
and
Thoenen
,
H.
(
1992a
).
Ciliary neurotrophic factor prevents degeneration of motor neurons in mouse mutant progressive motor neuronopathy
.
Nature
358
,
502
504
.
Sendtner
,
M.
,
Holtmann
,
B.
,
Kolbeck
,
R.
,
Thoenen
,
H.
and
Barde
,
Y-A.
(
1992b
).
Brain-derived neurotrophin factor prevents the death of motoneurons in newborn rats after nerve section
.
Nature
360
,
757
759
.
Soppet
,
D.
,
Escandon
,
E.
,
Maragos
,
J.
,
Middlemas
,
D. S.
,
Reid
,
S. W.
,
Blair
,
J.
,
Burton
,
L. E.
,
Stanton
,
B. R.
,
Kaplan
,
D. R.
,
Hunter
,
T.
,
Nikolics
,
K.
and
Parada
,
L. F.
(
1991
).
The neurotrophic factor and neurotrophin-3 are ligands for the trkB tyrosine kinase receptor
.
Cell
65
,
895
.
Squinto
,
S. P.
,
Stitt
,
T. N.
,
Aldrich
,
T. H.
,
Davis
,
S.
,
Bianco
,
S. M.
,
Radziejewski
,
C.
,
Glass
,
D. J.
,
Masiakowski
,
P.
,
Furth
,
M. E.
,
Valenzuela
,
D. M.
,
DiStefano
,
P. S.
and
Yancopoulos
,
G. D.
(
1991
).
TrkB encodes a functional receptor for brain-derived neurotrophic factor and neurotrophin-3 but nerve growth factor
.
Cell
54
,
885
.
Tessarollo
,
L.
,
Nagarajan
,
L.
,
Parada
,
L. F.
(
1992
)
C-ros: the vertebrate homolog of the sevenless tyrosine kinase receptor is tightly regulated during organogenesis in mouse embryonic development
.
Development
115
,
11
.
Tsoulfas
,
P.
,
Soppet
,
D.
,
Escandon
E.
,
Tessarollo
,
L.
,
MendozaRamirez
,
J.-L.
,
Rosenthal
,
A.
,
Nikolics
K.
and
Parada
,
L. F.
(
1993
)
The rat trkC locus encodes multiple neurogenic receptors that exhibit differential response to NT-3 in PC12 cells
.
Neuron (In press)
.
Van Heyningen
,
V.
,
Bobrow
,
M.
,
Bodmer
,
W. F.
,
Gardiner
,
S. E.
,
Povey
,
S.
and
Hopkinson
,
D. A.
(
1975
).
Chromosome assignment of some human enzyme loci: mitochondrial malate dehydrogenase to 7, mannose phosphate isomerase and pyruvate kinase to 15 and probably, esterase D to 13
.
Ann. Hum. Genet
.
38
,
295
303
.
Van Snick
,
J.
,
Goethals
,
A.
,
Renauld
,
J.-C.
,
Van Roost
,
E.
,
Uyttenhove
,
C.
,
Rubira
,
M. R.
,
Moritz
,
R. L.
and
Simpson
,
R. J.
(
1989
).
Cloning and characterization of a cDNA for a new mouse T cell growth factor (P40)
.
J. Exp. Med
.
169
,
363
368
.
White
,
W. F.
(
1985
).
The glycine receptor in the mutant mouse spastic (spa): strychnine binding characteristics and pharmacology
.
Brain Res
.
329
,
1
6
.
White
,
W. F.
and
Heller
,
A. H.
(
1982
).
Glycine receptor alteration in the mutant mouse spastic
.
Nature
298
,
655
657
.
Wilson
,
D. E.
,
Woodard
,
D.
,
Sandler
,
A.
,
Erikson
,
J.
and
Gurney
,
A.
(
1986
).
The gene for uridine monophosphatase-2 is on mouse chromosome 11
.
Am. J. Hum. Genet
.
39
,
173
.
Yan
,
Q.
,
Elliott
,
J.
and
Snider
W. D.
(
1992
).
Brain-derived neurotrophic factor rescues spinal motor neurons from axotomy-induced cell death
.
Nature
360
,
753
755
.