Download Fibroblast growth factors as regulators of central nervous system

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Neuropharmacology wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Transcript
Am J Physiol Regul Integr Comp Physiol 284: R867–R881, 2003;
10.1152/ajpregu.00533.2002.
invited review
Fibroblast growth factors as regulators of central
nervous system development and function
Rosanna Dono
Faculty of Biology, Department of Developmental Biology, Utrecht University,
NL-3584CH Utrecht, The Netherlands
neural stem cells; anterior neural ridge; isthimic organizer; neocortex;
neural development
FIBROBLAST GROWTH FACTORS (FGFs) constitute a large
family of structurally related polypeptide growth factors found in organisms ranging from nematodes to
humans (17, 53, 71, 74, 148, 189, 200). To date, the
mammalian FGF proteins are encoded by twenty-two
distinct genes known as Fgf1 to Fgf18 and Fgf20 to
Fgf23 in the mouse (25, 71, 133, 189, 236). FGF proteins are small peptides of 155 to 268 amino acid
residues (25, 71, 148, 189, 235). The degree of sequence
identity between different family members is 30–60%
in a “central domain” of ⬃120 amino acids. This domain confers to FGFs a common tertiary structure and
the ability to bind to heparin (42, 245). Most FGFs are
constitutively secreted using the endoplasmic reticulum-Golgi secretory pathway (77, 89, 203, 235). A small
subgroup of FGF proteins, such as FGF1, -2, -9, -16,
and -20, lacks the NH2-terminal signal sequence but is
still transported in the extracellular space (36, 103,
135, 137). A third subgroup of FGF proteins, FGF11 to
FGF14, lacks the signal peptide and remains intracellular (182).
Secreted FGFs signal to target cells by binding and
activating cell-surface tyrosine kinase FGF receptors
(FGFRs; 34, 82, 105, 146). FGFRs are transcribed from
four different genes and consist of an extracellular
domain, a single transmembrane domain, and a cytoplasmic tyrosine kinase domain (34, 82, 105). The extracellular domain contains three immunoglobulin-like
Address for reprint requests and other correspondence: R. Dono,
Dept. of Developmental Biology, Faculty of Biology, Utrecht Univ.,
Padualaan 8, NL-3584CH Utrecht, The Netherlands (E-mail:
[email protected]).
http://www.ajpregu.org
domains, called loops I, II, and III (81, 105). Although
loops II and III contact the bound ligand, the region
that dictates the specificity of binding is the COOHterminal portion of loop III (20, 150). Alternative
mRNA splicing of the COOH-terminal portion of loop
III creates several forms of FGFRs with unique ligandbinding properties (81, 150, 212). Once an FGF ligand
is bound, the receptor dimerizes and phosphorylates
intermolecular tyrosine residues, triggering initiation
of FGFR signal transduction (97, 138, 159). FGFR
signaling activates a number of signal transduction
molecules, including those of the Ras and phospholipase C-␥ pathways (92, 97, 139, 221). Interestingly,
the intracellular FGF12 and FGF14 do not bind to
FGFRs; nevertheless, they interact with the mitogenactivated protein (MAP) kinase scaffold protein IsletBrain-2 in neurons (145, 182, 222). The regulation of
FGFR-ligand interaction is complex. Receptor isoforms
can form heterodimers and share redundant ligand
binding specificity (150). Moreover, ligand binding is
affected by the distribution of heparan sulfate proteoglycans (HSPGs) on the cell surface and in the extracellular matrix (109, 151, 190). Extracellular FGFs,
indeed, bind tightly to HSPGs, which may restrict FGF
diffusion and favor interaction with receptors on
nearby cells (11, 142). Finally, HSPGs promote and
stabilize assembly of the FGF ligand-receptor complex
(181).
The function of FGFs and FGFRs during embryonic
development and adult physiology has been addressed
by gain- and loss-of-function experiments in several
animal model organisms. These studies have shown
that FGFs act as key regulators of developmental
0363-6119/03 $5.00 Copyright © 2003 the American Physiological Society
R867
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
Dono, Rosanna. Fibroblast growth factors as regulators of central
nervous system development and function. Am J Physiol Regul Integr
Comp Physiol 284: R867–R881, 2003; 10.1152/ajpregu.00533.2002.—
Fibroblast growth factors (FGFs) are multifunctional signaling proteins
that regulate developmental processes and adult physiology. Over the
last few years, important progress has been made in understanding the
function of FGFs in the embryonic and adult central nervous system. In
this review, I will first discuss studies showing that FGF signaling is
already required during formation of the neural plate. Next, I will
describe how FGF signaling centers control growth and patterning of
specific brain structures. Finally, I will focus on the function of FGF
signaling in the adult brain and in regulating maintenance and repair of
damaged neural tissues.
R868
INVITED REVIEW
neural induction occurs via inhibiting BMP signaling
in prospective neural cells, since BMP signaling promotes epidermal cell fate (231). Fgf ligands and receptors are expressed by the prospective neural cells and
by the adjacent inducing tissues [Fgf2, -3, -4, and -8
(175, 194, 195, 232); FgfR1, FgfR2, and FGF3 (219,
232)]. The involvement of FGF signaling in this process
has been the subject of intense investigation since the
discovery that FGFs promote expression of neuronal
markers in Xenopus laevis ectodermal cells from an
early gastrula stage (87, 88, 101). It has been proposed
that FGFs are required for neural induction, since
inhibition of FGF signaling in Xenopus embryos interferes with development of neural tissue (73, 102).
Moreover, FGF2- and FGF4-soaked beads can induce
ectopic neural structures when applied to chick primitive streak stage embryos (5).
Dorsal ectodermal cells of an early Xenopus gastrula
acquire neural fates in response to signaling by Spemann organizer cells (230). These signals antagonize
BMP activity and thereby prevent specification of epidermal cell fates (100, 180). Known organizer signals
include the two BMP antagonists noggin and chordin
(163, 179, 247). Strikingly, inhibition of FGF signaling
in ectodermal explants from a Xenopus early gastrula
precludes the induction of neural tissues by the Spemann’s organizer (73, 102). Moreover, noggin- and
chordin-mediated neural induction is abolished in the
absence of FGF signaling (102, 179). Taken together,
these studies indicate that FGFs cooperate with BMP
antagonists to induce neural cell fates.
In apparent controversy, other studies in mouse and
chicken embryos have shown that in these vertebrates
BMP antagonists are not required for neural induction
FGFS DURING INDUCTION AND EARLY PATTERNING
OF THE NEURAL PLATE
Neural induction is the first and fundamental step in
the formation of the vertebrate CNS. During this process, pluripotent dorsal ectodermal cells undergo a “cell
fate switch” and become neural stem cells instead of
epidermal cell types (230). It is generally accepted that
Table 1. Fgfs and FgfRs in the developing CNS
Expressing Structure
Genes
Te
Di
Ms
Mt
My
Sc
Fgf1
Fgf2
●
●
●
●
●
●
●
●
●
●
Fgf3
Fgf7
Fgf8
●
●
●
Fgf9
Fgf10
Fgf11
Fgf12
Fgf13
Fgf14
Fgf15
Fgf17
Fgf18
FgfR1
FgfR2
FgfR3
FgfR4
ANR
●
AM
●
Is
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
Phenotype
None
Neocortex, spinal
cord
Inner ear
None
Forebrain midbrain,
cerebellum
ND
ND
ND
ND
ND
None
ND
Cerebellum
ND
Early lethal, spina
bifida
Early lethal
None
ND
Ref. Nos.
2, 136, 227
36, 37, 153, 192
122, 229
126, 236
25, 134
23
Unpublished observation
189
68, 189
68, 189
222, 223
49, 133
235, 236
75, 236
152, 237, 239, 240
7, 37, 160
32, 158, 234
155, 224
Summary of expression in rat and mouse embryos. Phenotypes data derived from gene targeting experiments in mice. Fgf, fibroblast
growth factor gene; FgfR, fibroblast growth factor receptor gene; ANR, anterior neural ridge; AM, anterior midline; Is, isthmus; Te,
telencephalon; Di, diencephalons; Ms, mesencephalon; Mt, metencephalon; My, myelencephalon; Sc, spinal cord, ND, not determined.
AJP-Regul Integr Comp Physiol • VOL
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
events. For example, FGFs control growth and survival
of the postimplantation mouse embryos (7, 43), cell
migration during gastrulation (21, 22, 199), and establishment of the anterior-posterior (A/P) body axis (22,
134). At later developmental stages, FGFs function in
those organs and tissues in which reciprocal interactions between epithelial and mesenchymal cells are
important for morphogenesis and differentiation (31,
183, 198). The discovery that certain human skeletal
disorders are caused by point mutations in FgfR1,
FgfR2, and FgfR3 (143, 149, 176, 228), and the genetic
analysis of FGFR and FGF ligand functions in the
mouse (23, 64, 112), have revealed essential roles for
FGF signaling in chondrogenesis and osteogenesis.
Disruption of FGF signaling may also underlie other
pathologies, such as hypotension (37, 244), diabetes
(67), and the hypophosphatemic rickets disorder (226).
In this review, I will focus on the function of FGF
family members in the central nervous system (CNS).
Several Fgfs and FgfRs are expressed in the embryonic
and adult CNS (Table 1 and Refs. 51, 69, 70, 189, 193,
233, 241). I will summarize some of the findings showing that FGFs act as key regulators of CNS development and function. I will also discuss studies that
address FGF signaling in the adult brain and neural
stem cells.
R869
INVITED REVIEW
Fig. 1. Schematic representation of the developing central nervous
system (CNS) in mouse embryos. A: sagittal view of a mouse CNS at
embryonic day 13. At this developmental stage, the embryonic neural
tube is already patterned along the anterior/posterior (A/P) and
dorso/ventral (D/V) axis. B: sagittal view of a postnatal day 2 mouse
brain. The progenitor cells of the neocortex are located in the ventricular (VZ) and subventricular (VZ) zone. Te, telencephalon; Di,
diencephalon; Ms, mesencephalon; Is, isthmus; R1 to R8, rhombomere 1 to rhombomere 8; SC, spinal cord; Nx, neocortex; Hp,
hippocampus; St, striatum; Th, thalamus; Cb, cerebellum.
FGFS AS MEDIATORS OF NEUROEPITHELIAL
ORGANIZER FUNCTIONS
Before neural tube closure, the developing CNS is
subdivided along its A/P axis, also known as the rostrocaudal axis, into the following four distinct domains:
the forebrain, the midbrain, the hindbrain, and the
spinal cord (246). At later developmental stages, the
forebrain gives rise to the anterior telencephalon and
the more caudal diencephalon (Fig. 1A and Ref. 246).
The midbrain will develop as one mesencephalic vesicle (Fig. 1A), whereas the hindbrain is divided in rhombomeres, with the most anterior rhombomere 1 and
rhombomere 2, known as the metencephalon (Fig. 1A).
As a result of these initial patterning events, the neural tube becomes regionalized, and neural progenitors
acquire positional identity. For example, spinal cord
progenitor cells will from now on generate spinal cord
neurons, whereas telencephalic progenitors will only
generate telencephalic neurons. It has been shown that
A/P patterning of the early neuroectoderm is controlled
by local signaling centers. These signaling centers act
within the neural plate to induce and maintain regional identity in the surrounding neuroepithelium.
For example, the anterior neural ridge (ANR), which
lies at the junction between the anterior ectoderm and
the anterior neural plate, is necessary for growth and
maintenance of the forebrain identity (188). The isAJP-Regul Integr Comp Physiol • VOL
thimic organizer (IsO) lies at the midhindbrain junction and regulates proper development of the mesencephalic and metencephalic derivatives [e.g., optic
tectum and cerebellum (124)]. FGF family members,
such as Fgf8, Fgf17, and Fgf18, are expressed at the
ANR and IsO (Fig. 2A and Refs. 25, 125, 235, 236).
Genetic and embryological studies have shown that
they play important roles in executing neuroepithelium organizer functions (188). For example, removal
of the ANR in neural plate explants leads to downregulation of Bf1 expression, a transcription factor essential for growth and patterning of the telencephalic
vesicles (188, 238). The fact that beads soaked with
FGF8 can induce Bf1 expression suggests that FGF8
can substitute for ANR functions (188). In agreement
with these studies, Fgf8 is expressed by the ANR cells
from day 8.5 onward (25), and embryos carrying a
hypomorphic Fgf8 allele have small telencephalic vesicles (134). Moreover, zebrafish embryos lacking a functional FGF8 protein (“acerebellar” mutants) show disruption of the commissural axon pathway and
patterning defects in the basal telencephalon (185).
The phenotypes observed in acerebellar mutants are
less severe than those resulting from ANR ablation
(185, 188) suggesting that FGF8 is not the only medi284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
(10, 195, 196). Induction of the neural tissue occurs at
blastula stage and precedes establishment of a functional organizer [the mouse node and the chick Hensen’s node (91, 195, 232)]. In chicken embryos, FGF
signaling downregulates BMP expression in the prospective neural cells during blastula stages (232).
Streit et al. (195) have also shown that FGFs, produced
by organizer precursor cells at the posterior margin,
can initiate and promote neural tissue development
through a BMP-independent mechanism. Therefore,
FGF-mediated neural induction occurs via two distinct
mechanisms in chick epiblastic cells: the first mediated
by repression of BMP expression, the second being a
BMP-independent pathway. It will be interesting to
examine to which extent the proposed mechanisms are
conserved in mammals.
In addition to these early functions, FGFs promote
posterior neural fate during A/P patterning of the neural plate (35, 72, 96, 131, 165, 194). However, the
underlying molecular mechanisms are still largely unknown. Studies in Xenopus embryos have shown that
induction of posterior neural tissue mediated by the
transcription factors XBF2 and Xmeis3 requires the
activation of the FGF signaling-dependent Ras-MAP
kinase pathway (174). Studies in chicken embryos
have also shown that FGFs promote development of
posterior neural tissue by maintaining the proliferating neural progenitors contributing to posterior CNS
development (127). At later developmental stages, attenuation of FGF signaling is instead required to promote neuronal differentiation in the developing spinal
cord (33).
R870
INVITED REVIEW
ator of ANR function and acts in combination with
other ANR signals.
In contrast to the situation in the ANR, FGF8 is a
key mediator of the IsO functions. It is expressed by
the IsO cells when the IsO is active (embryonic days
8–12.5; Fig. 3A and Ref. 25). When FGF8 protein beads
are applied to different neural locations, for example
diencephalon and mesencephalon, FGF8 can induce
ectopic expression of genes normally present at the
mes-metencephalic junction and additional cerebellar
structures (26, 106, 123). Moreover, application of
FGF8 beads to the hindbrain rhombomere 1 shifts the
anterior boundary of the rhombomere 1 more posteri-
orly, as evidenced by changes in Hox gene expression
patterns (79). In agreement with these gain-of-function
studies, FGF8 loss-of-function mutations in zebrafish
and mice lead to IsO tissue loss (134, 171). In addition,
FGF8 zebrafish mutants lack cerebellar structure and
show patterning defects in the developing tectum and
its retinotectal projections (164, 171). Thus FGF8 secreted by the IsO cells influences cell specification,
leading to the induction and patterning of specific CNS
structures. Consistent with a role for FGF8 in cell fate
specification, FGF8 in combination with SHH and
FGF4 can induce dopaminergic and serotonergic neurons in neural plate explants (242). These and other
Fig. 3. Spatial distribution of Fgf8 and FGF2 during
patterning of the vertebrate CNS. A: distribution of
Fgf8 transcripts detected by whole mount in situ hybridization on a mouse embryo at embryonic day 9.25
(E9.25). Note the expression of Fgf8 by the anterior
medial cells of the telencephalon (arrowhead) and by
the isthimic organizer (IsO) cells (arrow). B: distribution of FGF2 proteins detected by whole mount antibody staining on a chicken embryo at stage 17 [staging
according to Hamburger and Hamilton (65)]. FGF2 proteins are present throughout the embryonic brain, and
protein levels are higher in the developing telencephalon and metencephalon (arrowheads). FGF2 proteins
are also found in the developing spinal cord (arrow). Ba,
branchial arch; FL, forelimb; HL, hindlimb; S, somite.
AJP-Regul Integr Comp Physiol • VOL
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
Fig. 2. Fibroblast growth factor (FGF) 2 functions during development of the neuronal layers in the embryonic
neocortex. Left: schematic representation of a coronal section from an embryotic day 14 mouse brain at the level of
the lateral ventricles (LV). The rectangle indicates the approximate position of the enlargement shown on right.
Right: enlargement of the ventricular zone (VZ) and developing cortical plate (CP). During development of the
neuronal layers (NL) of the neocortex, progenitor cells in the VZ divide asymmetrically (M) to generate one
proliferating cell that remains in the VZ (arrowhead) and one postmitotic immature neuron. Postmitotic immature
neurons leave the ventricular zone and migrate toward the margin of the cerebral wall, where they form the
neuronal layers. Proliferating progenitors generate postmitotic neurons at precise time points. Neurons that
become postmitotic during early development will form the deep layers, whereas those leaving the cell cycle at later
stages will migrate through the existing cell layers and form more superficial ones. In FGF2-deficient mice, a
fraction of postmitotic neurons fail to reach their target neocortical layers. FGF2 is highly expressed by the cells
of the ventricular zone, indicating that FGF2 is part of the signaling network that determines the laminar fate of
postmitotic migrating neurons. GE, ganglionic eminence. Modified from Ref. 177.
R871
INVITED REVIEW
FGF PROTEINS AS REGULATORS OF NEOCORTEX
DEVELOPMENT
During mammalian CNS development, the neocortex arises from the dorsal telencephalon (Fig. 1B). This
structure will undergo rapid expansion by midembryogenesis so that it will become the predominant brain
structure (140). As development proceeds, the neocortex is partitioned into anatomically distinct areas
along the A/P and mediolateral axis. For example, the
motor and sensory cortices develop in the anterior and
the visual cortex develops more posterior (140). Neurons of the neocortex will also be organized into six
distinct layers running from the lumen of the neural
tube to the margin of the cerebral wall (130). These cell
arrangements have functional consequences, since
AJP-Regul Integr Comp Physiol • VOL
neurons will develop synaptic connections according to
the position they occupy within the neocortical areas
and layers (130). It has been shown that growth and
patterning of the neocortex is strictly dependent on
localized production of instructive signals acting on the
progenitor cells, which lie near the lateral ventricle in
a layer known as the ventricular zone (VZ; Fig. 1B and
Ref. 130). Several studies have shown that FGFs are
among the regulatory signaling molecules. The expression of Fgfs and FgfRs in the developing mouse and rat
neocortices is spatially and temporally regulated. In
particular, the cells of the VZ express high levels of
FgfR1, FgfR2, and FgfR3 during the expansion of the
neocortical progenitor pool and throughout neurogenesis (37, 158, 160, 168). In agreement with receptor
distribution, FGF ligands are also found in the VZ. In
particular, FGF2 proteins are abundant at early developmental stages and nearly absent by the end of neurogenesis (37, 147, 167). Other FGFs, such as Fgf7 and
Fgf18, are also transiently expressed in the developing
neocortex [embryonic days 14.5–15.5 (75, 126)], and
their transcripts are found in the VZ and developing
cortical plate, respectively. In contrast to these, Fgfs,
Fgf8, and Fgf17 are predominantly expressed by the
anterior medial cells of the neocortical primordium
(Fig. 3A and Refs. 25 and 235), a signaling center for
neocortex A/P patterning (168). This suggests that
these FGFs might act as paracrine factors to regulate
development of the anterior neocortex. The biological
effects of FGFs on neocortical progenitor cells have
been first studied on cultured neocortical cells. FGF2
turned out to be among the most potent mitogenic and
survival factors for many CNS cell types, including
embryonic neocortical VZ cells (18, 169, 213, 214).
Other studies have shown that FGF2 acts either alone
or in combination with neurotrophins to promote differentiation of neocortical precursor cells (48, 144,
157). Finally, multipotential embryonic day 10 mouse
cortical cells will generate either neurons or astrocytes
in response to different concentrations of FGF2 in
culture media (166). Interestingly, other FGFs have
additional or distinct functions on neural progenitors.
For example, FGF8 can induce dopaminergic neurons
in explants of rostral fore- and midbrain, whereas
FGF4 and FGF2, but not FGF8, can ectopically induce
serotonergic neurons in midbrain explants (242). Moreover, FGF4 and FGF8b (an FGF8 isoform) promote
proliferation and survival of neuronal precursors,
whereas only FGF8b promotes differentiation along
the astrocyte lineage (63). The in vivo function of some
of these FGFs has been investigated by combining
embryonic manipulation with mouse molecular genetics. Genetic analysis of FGF2 functions in mice has
shown that FGF2 regulates neuronal density and cytoarchitecture of the developing neocortex (37, 153,
210). Neocortices of FGF2-deficient mice contain fewer
neurons at maturity (37, 153, 210) because of possible
defects in proliferation of progenitors (167). In addition, a fraction of postmitotic neurons fail to reach their
target layer in the developing neocortex of FGF2-deficient mice. These cells either remain in deeper layers
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
findings have led to the proposal that FGF8 produced
by the ANR and the IsO, in combination with SHH and
FGF4, creates a grid of positional information in the
neural tube that specifies forebrain and midbrain dopaminergic neurons and hindbrain serotonergic neurons (242).
Fgf17 is also expressed by IsO cells after the onset of
Fgf8 expression (170, 236). Loss-of-function studies in
mice have shown that these FGFs cooperate in regulating cerebellar growth and shape by maintaining the
precursor cell pool in an undifferentiated proliferating
state (236). Interesting, FGF2 proteins are also present
in the metencephalon (Fig. 3B and Ref. 37), and a
single peripheral injection of FGF2 stimulates granule
cell production and enhances cerebellar growth in newborn rats (19). The IsO and its adjacent cell layers also
express Fgf18 and Fgf15, shortly after the onset of Fgf8
transcription (49, 125). Taken together, these observations suggest that at least four different FGFs may act
sequentially to determine the final size and shape of
the cerebellum.
In addition to control cerebellar development, FGFs
appear to have more general roles during patterning of
the vertebrate hindbrain. For example, it has been
proposed that FGFs participate in the establishment of
rhombomere identity during regionalization of the
hindbrain (Fig. 1 and Ref. 117). In zebrafish embryos,
signaling by rhombomere 4 cells influences segmental
identity and promotes neuronal differentiation of adjacent rhombomeres (129). The presumptive rhombomere 4 cells specifically express Fgf3 and Fgf8, and
development of rhombomere 5 and 6 is impaired by
blocking FGF3 and FGF8 functions (129, 218). These
studies indicate that FGF3 and FGF8 at the rhombomere 4 mediate the action of this signaling center in
promoting development of more caudal rhombomeres.
It is important to note that expression of Fgf3 in
rhombomere 4 is conserved among vertebrates (119,
120), whereas Fgf8 expression is not. This raises the
question whether this FGF-mediated signaling center
also functions in other vertebrates. The fact that Fgf3
is coexpressed with Fgf4 in chicken embryos (184)
suggests that other Fgfs may substitute for Fgf8 in
other species.
R872
INVITED REVIEW
FGFS IN THE DEVELOPING SPINAL CORD
Development of the spinal cord leads to the establishment of different neuronal cell types along its dorsoventral axis (14). For example, motoneurons will
differentiate ventrally, commissural neurons will form
dorsally, and neurons of the autonomic nervous system
will develop within the intermediate spinal cord layer
(14, 192). Spinal motoneurons innervate the muscles,
and their survival is dependent on trophic factors that
are produced by the targeted muscle cells and by the
AJP-Regul Integr Comp Physiol • VOL
neuron itself (39). Several Fgfs (e.g., Fgf1 and -2 and -4
and -5) are expressed in the developing skeletal muscles (38, 66, 76) and by spinal motoneurons (e.g., Fgf1
and Refs. 9, 40, 83). Motoneurons also express FGFRs
(162, 220). In vitro, FGF2, FGF5, and FGF9 promote
survival of cultured chick and rat spinal motoneurons
(62, 76, 83) and FGF2 and FGF9 upregulate the choline
acetyltransferase (ChAT) activity in a dose-dependent
manner (62, 83). The biological effects of FGFs on
motoneurons have also been studied on experimentally
induced motoneuron damages and in animal models of
motoneuron diseases. As in vitro, also in vivo local
infusion of FGFs can rescue motoneuron death induced
by nerve fiber lesions (axotomy) or spinal cord injury
(28, 108, 206). However, levels of ChAT activity remain
low even in the presence of FGF2 (61). FGF2 also
mediates motoneuron survival in the wobbler mouse
affected by a motoneuron disease (78). FGF2-deficient
mice show neuronal deficiencies in the cervical spinal
cord region that also affect motoneuron density (37).
However, mice do not show apparent defects resulting
from the lack of a fraction of these motoneurons (37,
153, 244). Homozygous null Fgf9 mice die shortly after
birth, and motoneuron development has not been analyzed in these mice (24). Additional studies need to be
performed on these mutant strains to understand better if and how these FGFs contribute to motoneuron
development and function. It is likely that this process
requires a combination of different trophic factors,
among which are FGFs (14).
Physiological and pharmacological studies on FGF2deficient mice have shown that FGF2 is essential for
other spinal cord functions. In particular, lack of FGF2
causes an impaired baroreceptor reflex response to
hypotensive stimuli in adult mice (37). As a result of
the neuronal regulation defect, FGF2-deficient mice
show a reduced resting arterial blood pressure (37,
244). During CNS development, Fgf2 is expressed by
progenitor cells of neuronal circuits involved in the
central regulation of blood pressure, for example, in the
myelencephalon (Fig. 3B and Ref. 36) and the intermediolateral neurons of the spinal cord (192). Reexpression of FGF2 in the developing nervous system of
FGF2-deficient embryos leads to a rescue of the baroreceptor reflex and of the hypotensive phenotype (36).
These genetic studies indicate that FGF2 signaling is
essential for development of the neural circuitry regulating central regulation of blood pressure (36). The
distribution of certain classes of neurons of the intermediate cell layer is affected in FGF2-deficient embryos (E. ten Hove and R. Dono, unpublished observations), suggesting that positional identity of spinal cord
neurons may be altered.
FGFS IN BRAIN PHYSIOLOGY AND PATHOLOGY
Most of the Fgf family members and Fgf receptors
remain expressed in different cell types of the adult
brain (Fig. 4 and Refs. 50, 69, 70, 189, 193, 233, 241).
A number of recent studies are beginning to address
the role of FGFs in brain physiology and pathology.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
or accumulate in the corpus callosum (37). These genetic studies show that lack of FGF2 affects cell positioning in the developing neocortex. Development of
the neuronal layers of the neocortex begins once newly
generated postmitotic neurons leave the VZ and migrate to the cortical plate. It has been shown that the
migratory paths of postmitotic neurons are defined by
instructive signals acting on the progenitors in the VZ
as these cells undergo their last mitotic division (130).
Signals also act in the cortical plate and direct laminar
organization of migrating neurons. Neuronal cells are
exposed to FGF2 before or during onset of neuronal
migration (Fig. 2 and Ref. 37). Indeed, FGF2 is expressed at high levels by the cells of the VZ, whereas it
is not expressed by migrating neurons or other cortical
plate cells (37). Thus it is likely that FGF2 is part of the
signaling network that acts on the progenitor cells and
defines the cell fate and migratory path of postmitotic
neurons (Fig. 2).
The neuronal defects observed in FGF2-deficient cortices predominantly affect the frontal motor sensory
areas (37, 153, 211). This observation raises the possibility that FGFs are part of the signaling network
regulating differential growth and patterning of neocortical areas. As discussed above, Fgf8 is expressed by
the anterior medial cells of the telencephalon (25).
Fukuchi-Shimogori and Grove (44) have shown that
anterior expansion of the FGF8 source shifts the
boundaries of the cortical areas more posterior,
whereas reducing the endogenous FGF8 signal shifts
these boundaries more anterior. Moreover, an ectopic
posterior source of FGF8 instructs surrounding cells to
acquire anterior identity. It is important to note that
no changes in cortical size were observed in these
experiments. Thus Fgf8 appears to specify positional
identity in the neocortical primordium without affecting cell proliferation (44). Regional specification of the
neocortical neural stem cells is mediated by gradients
of transcriptional regulators such as Emx2 and Pax6.
In particular, Emx2 is expressed at high levels by
neocortical progenitors of the posterior VZ and at low
levels by those of the anterior VZ. In contrast, Pax6 is
high in the anterior VZ and low in the posterior (12,
121). FGF8 beads applied to the dorsal telencephalon
of developing chicken embryos inhibit Emx2 expression (27). It will be important to understand whether
FGF8 regulates regionalization of the neocortex by
acting on these transcription factors or on other unknown regulators.
R873
INVITED REVIEW
Most studies have focused on the possibility that
FGF2 acts as a neurotrophic factor on mature brain
neurons. For example, several laboratories have
shown that FGF2 can promote survival of neocortical, hippocampal, cerebellar, dopaminergic, spinal
cord, and sensory neurons isolated from adult CNS
(1, 13, 60, 95, 116, 128). The neuronal survival promoted by FGF2 is independent of its mitogenic activity for glial cells (217). However, studies on FGF2deficient mice have failed to reveal an increased cell
death in the brain of embryos and adults (37, 167).
Interestingly, FGF2 levels increase after CNS damage (50, 59), ischemia (47, 243), or seizure (243).
Changes in FGF2 levels are also observed in patients
with neurodegenerative disorders, such as Alzheimer’s and Parkinsons’ diseases (52, 207). These observations raise the possibility that FGF2 acts to
protect the brain from pathological events, where it
promotes survival and/or has additional effects on
neural cells. The biological effects of FGF2 in response to brain damage are currently being tested
using a variety of experimental systems. As discussed above, neuronal cell death can be induced in
vivo by performing axotomy. Peterson et al. (161)
have shown that axotomy-induced death of glutamatergic neurons is prevented by grafting fibroblasts
expressing FGF2 before axotomy. Brain damage can
also be induced by injecting kainic acid, which causes
seizures and neuronal cell death (113). When FGF2
is infused in the rat brain before seizure, it can
prevent cell loss in the hippocampal region (113).
FGF2 seems to rescue the injured neurons and promote brain regeneration through multiple strategies.
Infusion of antibodies against FGF2 in the lateral
ventricles leads to a significant reduction in sprouting of cholinergic neurons within a denervated hippocampus (41). These results are consistent with
AJP-Regul Integr Comp Physiol • VOL
neurite-promoting effects of FGF-2 on cultured cholinergic neurons (6, 154). Progenitor cells of the
hippocampus proliferate and differentiate in response to cerebral ischemia or seizures (110, 111,
201). Yoshimura et al. (243) have demonstrated that
this process is affected in FGF2-deficient mice. Neurogenesis is restored upon delivering exogenous
FGF2 to the hippocampus before cerebral ischemia
and/or seizures (243). The molecular mechanisms
underlying the potential FGF2-mediated repair of
damaged brain cells are still unknown. Lenhard et
al. (107) have shown that the neuroprotective effects
of FGF2 on glutamate-induced hippocampus lesions
are in part mediated by glial cell-derived neurotrophic factors. Instead, the neuroprotective action of
FGF2 in stroke-induced cell death is dependent on
the induction of activin A (208).
The potential use of FGF2 for treatment of brain
disorders is very attractive. Clinical trials of intravenous administration of FGF2 for treatment of
acute stroke are in progress (8, 9). However, further
studies are required to determine if FGF2 is an
efficient therapeutic reagent for treatment of disorders affecting the adult brain. Direct insight into a
more general role for FGFs in brain physiology and
pathology will come from detailed analysis of mouse
mutant strains carrying loss-of-function mutation of
FGFs or FGFRs expressed in the adult brain. Indeed,
genetic analysis of FGF14 functions has shown that
FGF14-deficient mice develop ataxia and hyperkinetic movement disorders similar to those found in
patients affected by Huntington’s disease, Parkinson’s disease, and dystonia (222). These motor abnormalities are associated with dysfunction of the
basal ganglia system and result from defects in axonal trafficking and synapsis.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
Fig. 4. Distribution of FGF2 proteins in the adult
mouse brain. A: FGF2 proteins were detected by whole
mount antibody staining on a coronal cryosections of an
adult mouse brain, using FGF2-specific antibodies (38).
The highest levels of FGF2 proteins were found in the
hippocampal pyramidal neurons of the CA1 region
(CA1) and in astrocytes (arrows). The staining in the
corpus callosum (CC) is nonspecific, as determined by
using control cryosection of adult FGF2-deficient mouse
brains. The rectangles indicate the approximate position of the enlargements shown in B and C. B: enlarged
view showing FGF2-positive hippocampal pyramidal
neurons of the CA1 region. C: enlarged view showing
FGF2-positive astrocytes (arrowheads) at the level of
the dentate gyrus (DG). Nx, neocortex.
R874
INVITED REVIEW
FGFS AND THE NEURAL STEM CELLS IN THE
ADULT CNS
AJP-Regul Integr Comp Physiol • VOL
CONCLUSIONS AND FUTURE DIRECTIONS
Research over the past years has advanced our understanding of the role of FGF signaling in the embryonic and adult CNS. These studies have shed light on
the role of FGFs during neural induction, patterning of
specific CNS regions, and in the establishment of functional neuronal circuits. However, several major issues
remain unanswered. The gene targeting approach in
mice and analysis of zebrafish mutants have clarified
the functions of some of the Fgfs expressed by neural
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
For many years, the adult brain was wrongly considered an entirely postmitotic structure. More recently,
research has corroborated previous findings by Altman
and coworkers (4, 45, 84, 86, 98, 205) that established
that specific areas of the adult brain retain the capacity
for neurogenesis. Neurogenesis occurs in at least two
sites of the adult brain: the subgranular zone of the
hippocampus (3, 85) and the telencephalic subventricular zone (SVZ; 94, 114, 118). The subgranular zone
generates the granule cells of the hippocampus (85, 86,
191), whereas the SVZ is the source of new olfactory
bulb neurons (93, 114, 115, 118). Genesis of neurons
has also been reported in the primate prefrontal cortex,
temporal cortex, and parietal cortex (4, 54, 84), and it
has been proposed that the SVZ might also be the
source of these neurons (4, 54, 84). Neurogenesis in the
adult brain relies on neural stem cells (29, 90, 172,
225). Neural stem cells are defined as undifferentiated
cell types that undergo self-renewal and thereby retain
their multilineage potential (132). Recently, neural
stem cells have been isolated from many CNS regions,
including the previously mentioned neurogenic zones
and nonneurogenic regions, such as the spinal cord (30,
90, 141, 225). In vitro, proliferating neural stem cells
form aggregates, so-called neurospheres, that maintain both the capability of self-renewal and the ability
to differentiate into neurons, astrocytes, and oligodendrocytes (46, 55, 173, 209, 213, 225). It has been shown
that the in vitro expansion and differentiation of neural stem cells can be regulated by adding extracellular
factors to the culture medium (46, 55, 58, 80, 156, 209,
213, 225). FGF2 and EGF are the most potent mitogens
and survival factors for cultured neural stem cells (46,
141, 172, 173, 209, 213, 225). For example, FGF2 and
EGF have been used both alone and in combination to
isolate and maintain stem cells of the adult SVZ and
the spinal cord in culture (55, 56, 141, 173, 187, 225).
FGF2 alone is sufficient to maintain neural stem cells
from either the adult striatum (57) or hippocampus
(46) of rodents. Taupin et al. (204) have recently shown
that cystatin C, a cystein proteinase inhibitor, is able
to potentiate the mitogenic activity of FGF2 and enables expansion of rat hippocampal neural stem cells
from single cells. It will be interesting to examine if
cystatin C also cooperates with FGF2 to promote proliferation of neural stem cells isolated from other CNS
regions.
One major question in stem cell biology concerns the
type of progeny that neural stem cells can generate.
Differentiation of cultured neural stem cells can be
induced by mitogen withdrawal or by addition of extracellular factors that will drive differentiation along
the neuronal or glia cell lineage (58, 80, 202). However,
it is not clear whether differentiating neural stem cells
are able to generate functional neuronal subtype, such
as GABAergic interneurons or cortical pyramidal neurons, in addition to differentiation of a generic neuronal phenotype. Recent studies using FGF2-responsive
hippocampal neural stem cells show that these cells
can differentiate into cells with phenotypes of GABAergic, dopaminergic, and cholinergic neurons when exposed to both retinoic acid and neurotrophins (202). In
addition, forced expression of the orphan nuclear receptor Nurr1 (104) in these cells induces predominant
differentiation of tyrosine hydroxylase (TH)-positive
dopaminergic neurons (178, 215).
Another way of testing the developmental potential
of neural stem cells in vivo is their reimplantation in
the adult CNS (15, 16). Suhonen et al. (197) have
demonstrated that cultures of adult rat hippocampal
neural stem cells undergo neuronal differentiation
only when implanted in neurogenic sites. When transplanted in the hippocampus, cells that migrate to the
neuronal layer of the dentate gyrus give rise to hippocampal-like neurons. In contrast, hippocampal progenitors differentiate into TH-positive neurons, a
marker for dopaminergic neurons, when implanted in
the rostral migratory stream leading to the olfactory
bulb (197). As discussed previously, neural stem-like
cells can be cultured from the adult spinal cord in the
presence of FGF2 (187). Shihabuddin et al. (186)
showed that these cells, although isolated from a nonneurogenic region, exhibit a broad developmental potential upon exposure to different environmental stimuli. Spinal cord neural stem cells give rise to glial cells
if transplanted back into the adult spinal cord. When
transplanted in the hippocampus, cells that integrate
in the neuronal layer of the dentate gyrus differentiate
into hippocampal-like neurons of the granular cell
layer (186). Alternatively, they acquire an astroglial
and oligodendroglial phenotype when integrated in
nonneurogenic regions of the hippocampus (186).
These studies indicate that neural stem cells expanded
in the presence of FGF2 retain pluripotency and integrate into the host tissue where they respond to local
differentiation signals.
The possibility that FGF2 may also promote proliferation of neural stem cells in vivo is currently being
investigated. For example, Wagner et al. (216) have
shown that subcutaneous injection of FGF2 increases
the number of proliferating cells in the SVZ and olfactory tract. Injection of FGF2 in the lateral ventricle of
an adult rat brain expands the SVZ progenitor pool and
increases the number of neurons in the olfactory bulb
(99). Finally, combined injection of FGF2 and cystatin
C to the adult dentate gyrus stimulates proliferation
and neurogenesis in the adult rat hippocampus (204).
R875
INVITED REVIEW
I am grateful to Liliana Minichiello and Klaus Unsicker for
discussion on Fig. 4. I also thank Jacqueline Deschamps, Rolf Zeller,
and members of the laboratory for critical comments that have
improved this manuscript. I apologize to many researchers in this
fast-moving field whose work I have not cited because of space
limitations.
REFERENCES
1. Acosta CG, Fabrega AR, Masco DH, and Lopez HS. A
sensory neuron subpopulation with unique sequential survival
dependence on nerve growth factor and basic fibroblast growth
factor during development. J Neurosci 21: 8873–8885, 2001.
2. Alam KY, Frostholm A, Hackshaw KV, Evans JE, Rotter
A, and Chiu IM. Characterization of the 1B promoter of fibroblast growth factor 1 and its expression in the adult and developing mouse brain. J Biol Chem 271: 30263–30271, 1996.
3. Altman J and Das GD. Autoradiographic and histological
evidence of postnatal hippocampal neurogenesis in rats.
J Comp Neurol 124: 319–335, 1965.
4. Altman J and Das GD. Post-natal origin of microneurones in
the rat brain. Nature 207: 953–956, 1965.
5. Alvarez IS, Araujo M, and Nieto MA. Neural induction in
whole chick embryo cultures by FGF. Dev Biol 199: 42–54,
1998.
6. Anderson KJ, Dam D, Lee S, and Cotman CW. Basic fibroblast growth factor prevents death of lesioned cholinergic neurons in vivo. Nature 332: 360–361, 1988.
7. Arman E, Haffner-Krausz R, Chen Y, Heath JK, and
Lonai P. 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, 1998.
8. Ay H, Ay I, Koroshetz WJ, and Finklestein SP. Potential
usefulness of basic fibroblast growth factor as a treatment for
stroke. Cerebrovasc Dis 9: 131–135, 1999.
9. Ay I, Sugimori H, and Finklestein SP. Intravenous basic
fibroblast growth factor (bFGF) decreases DNA fragmentation
and prevents downregulation of Bcl-2 expression in the ischemic brain following middle cerebral artery occlusion in rats.
Brain Res Mol Brain Res 87: 71–80, 2001.
10. Bachiller D, Klingensmith J, Kemp C, Belo JA, Anderson
RM, May SR, McMahon JA, McMahon AP, Harland RM,
Rossant J, and De Robertis EM. The organizer factors Chordin and Noggin are required for mouse forebrain development.
Nature 403: 658–661, 2000.
11. Baeg GH, Lin X, Khare N, Baumgartner S, and Perrimon
N. Heparan sulfate proteoglycans are critical for the organization of the extracellular distribution of Wingless. Development
128: 87–94, 2001.
AJP-Regul Integr Comp Physiol • VOL
12. Bishop KM, Goudreau G, and O’Leary DD. Regulation of
area identity in the mammalian neocortex by Emx2 and Pax6.
Science 288: 344–349, 2000.
13. Brewer GJ. Isolation and culture of adult rat hippocampal
neurons. J Neurosci Methods 71: 143–155, 1997.
14. Briscoe J and Ericson J. Specification of neuronal fates in
the ventral neural tube. Curr Opin Neurobiol 11: 43–49, 2001.
15. Brustle O, Choudhary K, Karram K, Huttner A, Murray
K, Dubois-Dalcq M, and McKay RD. Chimeric brains generated by intraventricular transplantation of fetal human brain
cells into embryonic rats. Nat Biotechnol 16: 1040–1044, 1998.
16. Brustle O, Spiro AC, Karram K, Choudhary K, Okabe S,
and McKay RD. In vitro-generated neural precursors participate in mammalian brain development. Proc Natl Acad Sci
USA 94: 14809–14814, 1997.
17. Burdine RD, Chen EB, Kwok SF, and Stern MJ. egl-17
encodes an invertebrate fibroblast growth factor family member
required specifically for sex myoblast migration in Caenorhabditis elegans. Proc Natl Acad Sci USA 94: 2433–2437, 1997.
18. Cavanagh JF, Mione MC, Pappas IS, and Parnavelas JG.
Basic fibroblast growth factor prolongs the proliferation of rat
cortical progenitor cells in vitro without altering their cell cycle
parameters. Cereb Cortex 7: 293–302, 1997.
19. Cheng Y, Tao Y, Black IB, and DiCicco-Bloom E. A single
peripheral injection of basic fibroblast growth factor (bFGF)
stimulates granule cell production and increases cerebellar
growth in newborn rats. J Neurobiol 46: 220–229, 2001.
20. Cheon HG, LaRochelle WJ, Bottaro DP, Burgess WH, and
Aaronson SA. High-affinity binding sites for related fibroblast
growth factor ligands reside within different receptor immunoglobulin-like domains. Proc Natl Acad Sci USA 91: 989–993,
1994.
21. Ciruna B and Rossant J. FGF signaling regulates mesoderm
cell fate specification and morphogenetic movement at the
primitive streak. Dev Cell 1: 37–49, 2001.
22. Ciruna BG, Schwartz L, Harpal K, Yamaguchi TP, and
Rossant J. 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, 1997.
23. Colvin JS, Bohne BA, Harding GW, McEwen DG, and
Ornitz DM. Skeletal overgrowth and deafness in mice lacking
fibroblast growth factor receptor 3. Nat Genet 12: 390–397,
1996.
24. Colvin JS, White AC, Pratt SJ, and Ornitz DM. Lung
hypoplasia and neonatal death in Fgf9-null mice identify this
gene as an essential regulator of lung mesenchyme. Development 128: 2095–2106, 2001.
25. Crossley PH and Martin GR. The mouse Fgf8 gene encodes
a family of polypeptides and is expressed in regions that direct
outgrowth and patterning in the developing embryo. Development 121: 439–451, 1995.
26. Crossley PH, Martinez S, and Martin GR. Midbrain development induced by FGF8 in the chick embryo. Nature 380:
66–68, 1996.
27. Crossley PH, Martinez S, Ohkubo Y, and Rubenstein JL.
Coordinate expression of Fgf8, Otx2, Bmp4, and Shh in the
rostral prosencephalon during development of the telencephalic
and optic vesicles. Neuroscience 108: 183–206, 2001.
28. Cuevas P, Carceller F, and Gimenez-Gallego G. Acidic
fibroblast growth factor prevents death of spinal cord motoneurons in newborn rats after nerve section. Neurol Res 17: 396–
399, 1995.
29. Davis AA and Temple S. A self-renewing multipotential stem
cell in embryonic rat cerebral cortex. Nature 372: 263–266,
1994.
30. Davis AP and Capecchi MR. Axial development and appendicular skeleton defects in mice with a targeted disruption of
hoxd-11. Development 120: 2187–2196, 1994.
31. De Moerlooze L, Spencer-Dene B, Revest J, Hajihosseini
M, Rosewell I, and Dickson C. An important role for the IIIb
isoform of fibroblast growth factor receptor 2 (FGFR2) in mesenchymal-epithelial signalling during mouse organogenesis.
Development 127: 483–492, 2000.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
cells (Table 1). Further studies will require tissue- and
stage-specific loss-of-function mutations in combination with the analysis of FGF compound mutant embryos and adults. FGFs are key regulators of CNS
development; therefore, it can be expected that mutations in Fgf genes or Fgf receptors underlie human
congenital malformation affecting CNS development
and function. Genetic analysis in vertebrates should
provide animal models to study the pathology underlying brain and spinal cord dysfunctions. A second
major area of research will concern the identification of
pathways that are activated in response to FGF signaling and their involvement in triggering cell typespecific responses. Finally, a better understanding of
the genetic hierarchies and interactions among FGFs
and other regulatory signals will lead to a more comprehensive view of molecular networks governing CNS
development and function.
R876
INVITED REVIEW
AJP-Regul Integr Comp Physiol • VOL
53. Gospodarowicz D. Localization of a fibroblast growth factor
and its effect alone and with hydrocortisone on 3T3 cell growth.
Nature 249: 123–127, 1974.
54. Gould E, Reeves AJ, Graziano MS, and Gross CG. Neurogenesis in the neocortex of adult primates. Science 286: 548–
552, 1999.
55. Gritti A, Cova L, Parati EA, Galli R, and Vescovi AL. Basic
fibroblast growth factor supports the proliferation of epidermal
growth factor-generated neuronal precursor cells of the adult
mouse CNS. Neurosci Lett 185: 151–154, 1995.
56. Gritti A, Frolichsthal-Schoeller P, Galli R, Parati EA,
Cova L, Pagano SF, Bjornson CR, and Vescovi AL. Epidermal and fibroblast growth factors behave as mitogenic regulators for a single multipotent stem cell-like population from
the subventricular region of the adult mouse forebrain. J Neurosci 19: 3287–3297, 1999.
57. Gritti A, Parati EA, Cova L, Frolichsthal P, Galli R,
Wanke E, Faravelli L, Morassutti DJ, Roisen F, Nickel
DD, and Vescovi AL. Multipotential stem cells from the adult
mouse brain proliferate and self-renew in response to basic
fibroblast growth factor. J Neurosci 16: 1091–1100, 1996.
58. Gross RE, Mehler MF, Mabie PC, Zang Z, Santschi L, and
Kessler JA. Bone morphogenetic proteins promote astroglial
lineage commitment by mammalian subventricular zone progenitor cells. Neuron 17: 595–606, 1996.
59. Grothe C, Meisinger C, and Claus P. In vivo expression and
localization of the fibroblast growth factor system in the intact
and lesioned rat peripheral nerve and spinal ganglia. J Comp
Neurol 434: 342–357, 2001.
60. Grothe C and Nikkhah G. The role of basic fibroblast growth
factor in peripheral nerve regeneration. Anat Embryol (Berl)
204: 171–177, 2001.
61. Grothe C and Unsicker K. Basic fibroblast growth factor in
the hypoglossal system: specific retrograde transport, trophic
and lesion-related responses. J Neurosci Res 32: 318–328, 1992.
62. Grothe C, Wewetzer K, Lagrange A, and Unsicker K.
Effects of basic fibroblast growth factor on survival and choline
acetyltransferase development of spinal cord neurons. Brain
Res Dev Brain Res 62: 257–261, 1991.
63. Hajihosseini MK and Dickson C. A subset of fibroblast
growth factors (Fgfs) promote survival, but Fgf-8b specifically
promotes astroglial differentiation of rat cortical precursor
cells. Mol Cell Neurosci 14: 468–485, 1999.
64. Hajihosseini MK, Wilson S, De Moerlooze L, and Dickson
C. A splicing switch and gain-of-function mutation in FgfR2IIIc hemizygotes causes Apert/Pfeiffer-syndrome-like phenotypes. Proc Natl Acad Sci USA 98: 3855–3860, 2001.
65. Hamburger V and Hamilton HL. A series of normal stages in
the development of chick embryo. J Morphol 88: 49–92, 1951.
66. Hannon K, Kudla AJ, McAvoy MJ, Clase KL, and Olwin
BB. Differentially expressed fibroblast growth factors regulate
skeletal muscle development through autocrine and paracrine
mechanisms. J Cell Biol 132: 1151–1159, 1996.
67. Hart AW, Baeza N, Apelqvist A, and Edlund H. Attenuation
of FGF signalling in mouse beta-cells leads to diabetes. Nature
408: 864–868, 2000.
68. Hartung H, Feldman B, Lovec H, Coulier F, Birnbaum D,
and Goldfarb M. Murine FGF-12 and FGF-13: expression in
embryonic nervous system connective tissue and heart. Mech
Dev 64: 31–39, 1997.
69. Hattori Y, Miyake A, Mikami T, Ohta M, and Itoh N.
Transient expression of FGF-5 mRNA in the rat cerebellar
cortex during post-natal development. Brain Res Mol Brain Res
47: 262–266, 1997.
70. Hattori Y, Yamasaki M, Konishi M, and Itoh N. Spatially
restricted expression of fibroblast growth factor-10 mRNA in
the rat brain. Brain Res Mol Brain Res 47: 139–146, 1997.
71. Hebert JM, Basilico C, Goldfarb M, Haub O, and Martin
GR. Isolation of cDNAs encoding four mouse FGF family members and characterization of their expression pattern during
embrogenesis. Dev Biol 138: 454–463, 1990.
72. Holowacz T and Sokol S. FGF is required for posterior neural
patterning but not for neural induction. Dev Biol 205: 296–308,
1999.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
32. Deng C, Wynshaw Boris A, Zhou F, Kuo A, and Leder P.
Fibroblast growth factor receptor 3 is a negative regulator of
bone growth. Cell 84: 911–921, 1996.
33. Diez del Corral R, Breitkreuz DN, and Storey KG. Onset of
neuronal differentiation is regulated by paraxial mesoderm and
requires attenuation of FGF signalling. Development 129:
1681–1691, 2002.
34. Dionne CA, Crumley G, Bellot F, Kaplow JM, Searfoss G,
Ruta M, Burgess WH, Jaye M, and Schlessinger J. Cloning
and expression of two distinct high-affinity receptors crossreacting with acidic and basic fibroblast growth factors. EMBO
J 9: 2685–2692, 1990.
35. Domingos PM, Itasaki N, Jones CM, Mercurio S, Sargent
MG, Smith JC, and Krumlauf R. The Wnt/beta-catenin pathway posteriorizes neural tissue in Xenopus by an indirect mechanism requiring FGF signalling. Dev Biol 239: 148–160, 2001.
36. Dono R, Faulhaber J, Galli A, Zuniga A, Volk T, Texido G,
Zeller R, and Ehmke H. FGF2 signaling is required for the
development of neuronal circuits regulating blood pressure.
Circ Res 90: E5–E10, 2002.
37. Dono R, Texido G, Dussel R, Ehmke H, and Zeller R.
Impaired cerebral cortex development and blood pressure regulation in FGF-2-deficient mice. EMBO J 17: 4213–4225, 1998.
38. Dono R and Zeller R. Cell-type-specific nuclear translocation
of fibroblast growth factor-2 isoforms during chicken kidney
and limb morphogenesis. Dev Biol 163: 316–330, 1994.
39. Eisen JS. Patterning motoneurons in the vertebrate nervous
system. Trends Neurosci 22: 321–326, 1999.
40. Elde R, Cao YH, Cintra A, Brelje TC, Pelto-Huikko M,
Junttila T, Fuxe K, Pettersson RF, and Hokfelt T. Prominent expression of acidic fibroblast growth factor in motor and
sensory neurons. Neuron 7: 349–364, 1991.
41. Fagan AM, Suhr ST, Lucidi-Phillipi CA, Peterson DA,
Holtzman DM, and Gage FH. Endogenous FGF-2 is important for cholinergic sprouting in the denervated hippocampus.
J Neurosci 17: 2499–2511, 1997.
42. Faham S, Hileman RE, Fromm JR, Linhardt RJ, and
Rees DC. Heparin structure and interactions with basic fibroblast growth factor. Science 271: 1116–1120, 1996.
43. Feldman B, Poueymirou W, Papaioannou VE, DeChiara
TM, and Goldfarb M. Requirement of FGF-4 for postimplantation mouse development. Science 267: 246–249, 1995.
44. Fukuchi-Shimogori T and Grove EA. Neocortex patterning
by the secreted signaling molecule FGF8. Science 294: 1071–
1074, 2001.
45. Gage FH. Mammalian neural stem cells. Science 287: 1433–
1438, 2000.
46. Gage FH, Coates PW, Palmer TD, Kuhn HG, Fisher LJ,
Suhonen JO, Peterson DA, Suhr ST, and Ray J. Survival
and differentiation of adult neuronal progenitor cells transplanted to the adult brain. Proc Natl Acad Sci USA 92: 11879–
11883, 1995.
47. Ganat Y, Soni S, Chacon M, Schwartz ML, and Vaccarino
FM. Chronic hypoxia up-regulates fibroblast growth factor ligands in the perinatal brain and induces fibroblast growth
factor-responsive radial glial cells in the sub-ependymal zone.
Neuroscience 112: 977–991, 2002.
48. Ghosh A and Greenberg E. Distinct roles for bFGF and NT-3
in the regulation of cortical neurogenesis. Neuron 15: 89–103,
1995.
49. Gimeno L, Hashemi R, Brulet P, and Martinez S. Analysis
of Fgf15 expression pattern in the mouse neural tube. Brain
Res Bull 57: 297–299, 2002.
50. Gomez-Pinilla F and Cotman CW. Transient lesion-induced
increase of basic fibroblast growth factor and its receptor in
layer VIb (subplate cells) of the adult rat cerebral cortex. Neuroscience 49: 771–780, 1992.
51. Gomez-Pinilla F and Cotman CW. Distribution of fibroblast
growth factor 5 mRNA in the rat brain: an in situ hybridization
study. Brain Res 606: 79–86, 1993.
52. Gomez-Pinilla F, Cummings BJ, and Cotman CW. Induction of basic fibroblast growth factor in Alzheimer’s disease
pathology. Neuroreport 1: 211–214, 1990.
R877
INVITED REVIEW
AJP-Regul Integr Comp Physiol • VOL
93. Kornack DR and Rakic P. The generation, migration, and
differentiation of olfactory neurons in the adult primate brain.
Proc Natl Acad Sci USA 98: 4752–4757, 2001.
94. Kornack DR and Rakic P. Cell proliferation without neurogenesis in adult primate neocortex. Science 294: 2127–2130,
2001.
95. Kornblum HI, Raymon HK, Morrison RS, Cavanaugh KP,
Bradshaw RA, and Leslie FM. Epidermal growth factor and
basic fibroblast growth factor: effects on an overlapping population of neocortical neurons in vitro. Brain Res 535: 255–263,
1990.
96. Koshida S, Shinya M, Nikaido M, Ueno N, SchulteMerker S, Kuroiwa A, and Takeda H. Inhibition of BMP
activity by the FGF signal promotes posterior neural development in zebrafish. Dev Biol 244: 9–20, 2002.
97. Kouhara H, Hadari YR, Spivak-Kroizman T, Schilling J,
Bar-Sagi D, Lax I, and Schlessinger J. A lipid-anchored
Grb2-binding protein that links FGF-receptor activation to the
Ras/MAPK signaling pathway. Cell 89: 693–702, 1997.
98. Kuhn HG, Dickinson-Anson H, and Gage FH. Neurogenesis
in the dentate gyrus of the adult rat: age-related decrease of
neuronal progenitor proliferation. J Neurosci 16: 2027–2033,
1996.
99. Kuhn HG, Winkler J, Kempermann G, Thal LJ, and Gage
FH. Epidermal growth factor and fibroblast growth factor-2
have different effects on neural progenitors in the adult rat
brain. J Neurosci 17: 5820–5829, 1997.
100. Lamb T, Knecht A, Smith W, Stachel S, Economides A,
Stahl N, Yancopolous G, and Harland R. Neural induction
by the secreted polypeptide noggin. Science 262: 713–718, 1993.
101. Lamb TM and Harland RM. Fibroblast growth factor is a
direct neural inducer, which combined with noggin generates
anterior-posterior neural pattern. Development 121: 3627–3636,
1995.
102. Launay C, Fromentoux V, Shi DL, and Boucaut JC. A
truncated FGF receptor blocks neural induction by endogenous
Xenopus inducers. Development 122: 869–880, 1996.
103. LaVallee TM, Tarantini F, Gamble S, Mouta Carreira C,
Jackson A, and Maciag T. Synaptotagmin-1 is required for
fibroblast growth factor-1 release. J Biol Chem 273: 22217–
22223, 1998.
104. Law SW, Conneely OM, DeMayo FJ, and O’Malley BW.
Identification of a new brain-specific transcription factor,
NURR1. Mol Endocrinol 6: 2129–2135, 1992.
105. Lee PL, Johnson DE, Cousens LS, Fried VA, and Williams
LT. Purification and complementary DNA cloning of a receptor
for basic fibroblast growth factor. Science 245: 57–60, 1989.
106. Lee SMK, Danielian PS, Fritzsch B, and McMahon AP.
Evidence that FGF8 signalling from the midbrain-hindbrain
junction regulates growth and polarity in the developing midbrain. Development 124: 959–969, 1997.
107. Lenhard T, Schober A, Suter-Crazzolara C, and Unsicker
K. Fibroblast growth Factor-2 requires glial-cell-derived
neurtrophic factor exerting its neuroprotective actions on glutamate-lesioned hippocampal neurons. Mol Cell Neurosci 20:
181–197, 2002.
108. Li L, Oppenheim RW, Lei M, and Houenou LJ. Neurotrophic agents prevent motoneuron death following sciatic nerve
section in the neonatal mouse. J Neurobiol 25: 759–766, 1994.
109. Lin X, Buff EM, Perrimon N, and Michelson AM. Heparan
sulfate proteoglycans are essential for FGF receptor signaling
during Drosophila embryonic development. Development 126:
3715–3723, 1999.
110. Liu J, Solway K, Messing RO, and Sharp FR. Increased
neurogenesis in the dentate gyrus after transient global ischemia in gerbils. J Neurosci 18: 7768–7778, 1998.
111. Liu L, Bradley WD, and Kardami E. Perinatal phenotype
and hypothyroidism are associated with elevated levels of 21.5to 22-kDa basic fibroblast growth factor in cardiac ventricle.
Dev Biol 157: 507–516, 1993.
112. Liu Z, Xu J, Colvin JS, and Ornitz DM. Coordination of
chondrogenesis and osteogenesis by fibroblast growth factor 18.
Genes Dev 16: 859–869, 2002.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
73. Hongo I, Kengaku M, and Okamoto H. FGF signaling and
the anterior neural induction in Xenopus. Dev Biol 216: 561–
581, 1999.
74. Hoshikawa M, Ohbayashi N, Yonamine A, Konishi M,
Ozaki K, Fukui S, and Itoh N. Structure and expression of a
novel fibroblast growth factor, FGF-17, preferentially expressed
in the embryonic brain. Biochem Biophys Res Commun 244:
187–191, 1998.
75. Hu MC, Qiu WR, Wang YP, Hill D, Ring BD, Scully S,
Bolon B, DeRose M, Luethy R, Simonet WS, Arakawa T,
and Danilenko DM. FGF-18, a novel member of the fibroblast
growth factor family, stimulates hepatic and intestinal proliferation. Mol Cell Biol 18: 6063–6074, 1998.
76. Hughes RA, Sendtner M, Goldfarb M, Lindholm D, and
Thoenen H. Evidence that fibroblast growth factor 5 is a major
muscle-derived survival factor for cultured spinal motoneurons.
Neuron 10: 369–377, 1993.
77. Iida S, Naito K, Sajamoto H, Kato O, Hirohashi S, Sato T,
Onda M, Sugimura T, and Terada M. Human hst-2 (FGF-6)
oncogene: cDNA cloning and characterization. Oncogene 7:
303–309, 1992.
78. Ikeda K, Iwasaki Y, Tagaya N, Shiojima T, Kobayashi T,
and Kinoshita M. Neuroprotective effect of basic fibroblast
growth factor on wobbler mouse motor neuron disease. Neurol
Res 17: 445–448, 1995.
79. Irving C and Mason I. Signalling by FGF8 from the isthmus
patterns anterior hindbrain and establishes the anterior limit
of Hox gene expression. Development 127: 177–186, 2000.
80. Johe KK, Hazel TG, Muller T, Dugich-Djordjevic MM,
and McKay RD. Single factors direct the differentiation of
stem cells from the fetal and adult central nervous system.
Genes Dev 10: 3129–3140, 1996.
81. Johnson DE, Lu J, Chen H, Werner S, and Williams LT.
The human fibroblast growth factor receptor genes: a common
structural arrangement underlies the mechanisms for generating receptor forms that differ in their third immunoglobulin
domain. Mol Cell Biol 11: 4627–4634, 1991.
82. Johnson DE and Williams LT. Structural and functional
diversity in the FGF receptor multigene family. Adv Cancer Res
60: 1–41, 1993.
83. Kanda T, Iwasaki T, Nakamura S, Ueki A, Kurokawa T,
Ikeda K, and Mizusawa H. FGF-9 is an autocrine/paracrine
neurotrophic substance for spinal motoneurons. Int J Dev Neurosci 17: 191–200, 1999.
84. Kaplan MS. Neurogenesis in the 3-month-old rat visual cortex.
J Comp Neurol 195: 323–338, 1981.
85. Kaplan MS and Bell DH. Neuronal proliferation in the
9-month-old rodent-radioautographic study of granule cells in
the hippocampus. Exp Brain Res 52: 1–5, 1983.
86. Kaplan MS and Bell DH. Mitotic neuroblasts in the 9-day-old
and 11-month-old rodent hippocampus. J Neurosci 4: 1429–
1441, 1984.
87. Kengaku M and Okamoto H. Basic fibroblast growth factor
induces differentiation of neural tube and neural crest lineages
of cultured ectoderm cells from Xenopus gastrula. Development
119: 1067–1078, 1993.
88. Kengaku M and Okamoto H. bFGF as a possible morphogen
for the anteroposterior axis of the central nervous system in
Xenopus. Development 121: 3121–3130, 1995.
89. Kiefer P, Acland P, Pappin D, Peters G, and Dickson C.
Competition between nuclear localisation and secretory signals
determines the subcellular fate of single CUG-initiated form of
FGF3. EMBO J 13: 4126–4136, 1994.
90. Kilpatrick TJ and Bartlett PF. Cloned multipotential precursors from the mouse cerebrum require FGF-2, whereas glial
restricted precursors are stimulated with either FGF-2 or EGF.
J Neurosci 15: 3653–3661, 1995.
91. Klingensmith J, Ang SL, Bachiller D, and Rossant J.
Neural induction and patterning in the mouse in the absence of
the node and its derivatives. Dev Biol 216: 535–549, 1999.
92. Klint P and Claesson-Welsh L. Signal transduction by fibroblast growth factor receptors. Front Biosci 4: D165–D177, 1999.
R878
INVITED REVIEW
AJP-Regul Integr Comp Physiol • VOL
134. Meyers EN, Lewandoski M, and Martin GR. An Fgf8 mutant allelic series generated by Cre- and Flp-mediated recombination. Nat Genet 18: 136–141, 1998.
135. Mignatti P, Morimoto T, and Rifkin DB. Basic fibroblast
growth factor, a protein devoid of secretory signal sequence, is
released by cells via a pathway independnat of the endoplasmatic reticulum-golgi complex. J Cell Physiol 151: 81–93, 1992.
136. Miller DL, Ortega S, Bashayan O, Basch R, and Basilico
C. Compensation by fibroblast growth factor 1 (FGF1) does not
account for the mild phenotypic defects observed in FGF2 null
mice. Mol Cell Biol 20: 2260–2268, 2000.
137. Miyakawa K, Hatsuzawa K, Kurokawa T, Asada M, Kuroiwa T, and Imamura T. A hydrophobic region locating at
the center of fibroblast growth factor-9 is crucial for its secretion. J Biol Chem 274: 29352–29357, 1999.
138. Mohammadi M, Dionne CA, Li W, Li N, Spivak T, Honegger AM, Jaye M, and Schlessinger J. Point mutation in FGF
receptor eliminates phosphatidylinositol hydrolysis without affecting mitogenesis. Nature 358: 681–684, 1992.
139. Mohammadi M, Honegger AM, Rotin D, Fischer R, Bellot
F, Li W, Dionne CA, Jaye M, Rubinstein M, and Schlessinger J. A tyrosine-phosphorylated carboxy-terminal peptide
of the fibroblast growth factor receptor (Flg) is a binding site for
the SH2 domain of phospholipase C-gamma 1. Mol Cell Biol 11:
5068–5078, 1991.
140. Monuki ES and Walsh CA. Mechanisms of cerebral cortical
patterning in mice and humans. Nat Neurosci Suppl 4: 1199–
1206, 2001.
141. Morshead CM, Reynolds BA, Craig CG, McBurney MW,
Staines WA, Morassutti D, Weiss S, and van der Kooy D.
Neural stem cells in the adult mammalian forebrain: a relatively quiescent subpopulation of subependymal cells. Neuron
13: 1071–1082, 1994.
142. Moscatelli D. High and low affinity binding sites for basic
fibroblast growth factor on cultured cells: absence of a role for
low affinity binding in the stimulation of plasminogen activator
production by bovine capillary endothelial cells. J Cell Physiol
131: 123–130, 1987.
143. Muenke M and Schell U. Fibroblast-growth-factor receptor
mutations in human skeletal disorders. Trends Genet 11: 308–
313, 1995.
144. Murphy M, Drago J, and Bartlett PF. Fibroblast growth
factor stimulates the proliferation and differentiation of neural
precursor cells in vitro. J Neurosci Res 25: 463–475, 1990.
145. Negri S, Oberson A, Steinmann M, Sauser C, Nicod P,
Waeber G, Schorderet DF, and Bonny C. cDNA cloning and
mapping of a novel islet-brain/JNK-interacting protein. Genomics 64: 324–330, 2000.
146. Neufeld G and Gospodarowicz D. Basic and acidic fibroblast
growth factors interact with the same cell surface receptors.
J Biol Chem 261: 5631–5637, 1986.
147. Nurcombe V, Ford MD, Wildchut JA, and Bartlett PF.
Developmental regulation of neural response to FGF-1 and
FGF-2 by heparan sulfate proteoglycan. Science 260: 103–106,
1993.
148. Ornitz DM and Itoh N. Fibroblast growth factors. Genome
Biol 2: 3005.3001–3005.3012, 2001.
149. Ornitz DM and Marie PJ. FGF signaling pathways in endochondral and intramembranous bone development and human
genetic disease. Genes Dev 16: 1446–1465, 2002.
150. Ornitz DM, Xu J, Colvin JS, McEwen DG, MacArthur CA,
Coulier F, Gao G, and Goldfarb M. Receptor specificity of
the fibroblast growth factor family. J Biol Chem 271: 15292–
15297, 1996.
151. Ornitz DM, Yayon A, Flanagan JG, Svahn CM, Levi E,
and Leder P. Heparin is required for cell-free binding of basic
fibroblast growth factor to a soluble receptor and for mitogenesis in whole cells. Mol Cell Biol 12: 240–247, 1992.
152. Orr-Urtreger A, Givol D, Yayon A, Yarden Y, and Lonai P.
Developmental expression of two murine fibroblast growth factor receptors flg and bek. Development 113: 1419–1434, 1991.
153. Ortega S, Ittmann M, Tsang SH, Ehrlich M, and Basilico
C. Neuronal defects and delayed wound healing in mice lacking
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
113. Liuz Z, D’ Amore PA, Mikati M, Gatt A, and Holmes GL.
Neuroprotective effect of chronic infusion of basic fibroblast
growth factor on seizure-associated hippocampal damage.
Brain Res 626: 335–338, 1993.
114. Lois C and Alvarez-Buylla A. Proliferating subventricular
zone cells in the adult mammalian forebrain can differentiate
into neurons and glia. Proc Natl Acad Sci USA 90: 2074–2077,
1993.
115. Lois C, Garcia-Verdugo JM, and Alvarez-Buylla A. Chain
migration of neuronal precursors. Science 271: 978–981, 1996.
116. Lowenstein DH and Arsenault L. The effects of growth
factors on the survival and differentiation of cultured dentate
gyrus neurons. J Neurosci 16: 1759–1769, 1996.
117. Lumsden A and Guthrie S. Alternative patterns of cell surface properties and neural crest migration during segmentation
of the chick hindbrain. Dev Suppl 2: 9–15, 1991.
118. Luskin MB. Restricted proliferation and migration of postnatally generated neurons derived from the forebrain subventricular zone. Neuron 11: 173–189, 1993.
119. Mahmood R, Kiefer P, Guthrie S, Dickson C, and Mason
I. Multiple roles for FGF-3 during cranial neural development
in the chicken. Development 121: 1399–1410, 1995.
120. Mahmood R, Mason IJ, and Morriss-Kay GM. Expression
of Fgf-3 in relation to hindbrain segmentation, otic pit position
and pharyngeal arch morphology in normal and retinoic acidexposed mouse embryos. Anat Embryol (Berl) 194: 13–22, 1996.
121. Mallamaci A, Iannone R, Briata P, Pintonello L, Mercurio S, Boncinelli E, and Corte G. EMX2 protein in the
developing mouse brain and olfactory area. Mech Dev 77: 165–
172, 1998.
122. Mansour SL, Goddard JM, and Capecchi MR. Mice homozygous for a targeted discruption of the proto-oncogene int-2
have developmental defects in the tail and inner ear. Development 117: 13–28, 1993.
123. Martinez S, Crossley PH, Cobos I, Rubenstein JL, and
Martin GR. FGF8 induces formation of an ectopic isthmic
organizer and isthmocerebellar development via a repressive
effect on Otx2 expression. Development 126: 1189–1200, 1999.
124. Martinez S, Wassef M, and Alvarado-Mallart RM. Induction of a mesencephalic phenotype in the 2-day-old chick
prosencepgalon is preceded by the early expression of the homeobox gene En. Neuron 6: 971–981, 1991.
125. Maruoka Y, Ohbayashi N, Hoshikawa M, Itoh N, Hogan
BM, and Furuta Y. Comparison of the expression of three
highly related genes, Fgf8, Fgf17 and Fgf18, in the mouse
embryo. Mech Dev 74: 175–177, 1998.
126. Mason IJ, Fuller-Pace F, Smith R, and Dickson C. FGF-7
(keratinocyte growth factor) expression during mouse development suggests roles in myogenesis, forebrain regionalisation
and epithelial-mesenchymal interactions. Mech Dev 45: 15–30,
1994.
127. Mathis L, Kulesa PM, and Fraser SE. FGF receptor signalling is required to maintain neural progenitors during Hensen’s
node progression. Nat Cell Biol 3: 559–566, 2001.
128. Matsuda S, Saito H, and Nishiyama N. Effect of basic
fibroblast growth factor on neurons cultured from various regions of postnatal rat brain. Brain Res 520: 310–316, 1990.
129. Maves L, Jackman W, and Kimmel CB. FGF3 and FGF8
mediate a rhombomere 4 signaling activity in the zebrafish
hindbrain. Development 129: 3825–3837, 2002.
130. McConnell SK. Constructing the cerebral cortex: neurogenesis and fate determination. Neuron 15: 761–768, 1995.
131. McGrew LL, Hoppler S, and Moon RT. Wnt and FGF
pathways cooperatively pattern anteroposterior neural ectoderm in Xenopus. Mech Dev 69: 105–114, 1997.
132. McKay R. Stem cells in the central nervous system. Science
276: 66–71, 1997.
133. McWhirter JR, Goulding M, Weiner JA, Chun J, and
Murre C. A novel fibroblast growth factor gene expressed in
the developing nervous system is a downstream target of the
chimeric homeodomain oncoprotein E2A-Pbx1. Development
124: 3221–3232, 1997.
R879
INVITED REVIEW
154.
155.
156.
157.
158.
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
AJP-Regul Integr Comp Physiol • VOL
174. Ribisi S Jr, Mariani FV, Aamar E, Lamb TM, Frank D,
and Harland RM. Ras-mediated FGF signaling is required for
the formation of posterior but not anterior neural tissue in
Xenopus laevis. Dev Biol 227: 183–196, 2000.
175. Riese J, Zeller R, and Dono R. Nucleo-cytoplasmic translocation and secretion of fibroblast growth factor-2 during avian
gastrulation. Mech Dev 49: 13–22, 1995.
176. Rousseau F, Bonaventure J, Legeai-Mallet L, Pelet A,
Rozet JM, Maroteaux P, Le Merrer M, and Munnich A.
Mutations in the gene encoding fibroblast growth factor receptor-3 in achondroplasia. Nature 371: 252–254, 1994.
177. Roztocil T, Mater-Sadzinski L, Aliliod C, Ballivet M, and
Matter JM. NeuroM, a neural helix-loop-halix transcription
factor, defines a new transition stage in neurogenesis. Development 124: 3263–3272, 1997.
178. Sakurada K, Ohshima-Sakurada M, Palmer TD, and
Gage FH. Nurr1, an orphan nuclear receptor, is a transcriptional activator of endogenous tyrosine hydroxylase in neural
progenitor cells derived from the adult brain. Development 126:
4017–4026, 1999.
179. Sasai Y, Lu B, Piccolo S, and De Robertis EM. Endoderm
induction by the organizer-secreted factors chordin and noggin
in Xenopus animal caps. EMBO J 15: 4547–4555, 1996.
180. Sasai Y, Lu B, Steinbeisser H, and De Robertis EM. Regulation of neural induction by the Chd and Bmp-4 antagonistic
patterning signals in Xenopus (Abstract). Nature 377: 757,
1995.
181. Schlessinger J, Plotnikov AN, Ibrahimi OA, Eliseenkova
AV, Yeh BK, Yayon A, Linhardt RJ, and Mohammadi M.
Crystal structure of a ternary FGF-FGFR-heparin complex
reveals a dual role for heparin in FGFR binding and dimerization. Mol Cell 6: 743–750, 2000.
182. Schoorlemmer J and Goldfarb M. Fibroblast growth factor
homologous factors are intracellular signaling proteins. Curr
Biol 11: 793–797, 2001.
183. Sekine K, Ohuchi H, Fujiwara M, Yamasaki M, Yoshizawa T, Sato T, Yagishita N, Matsui D, Koga Y, Itoh N,
and Kato S. Fgf10 is essential for limb and lung formation. Nat
Genet 21: 138–141, 1999.
184. Shamim H and Mason I. Expression of Fgf4 during early
development of the chick embryo. Mech Dev 85: 189–192, 1999.
185. Shanmugalingam S, Houart C, Picker A, Reifers F, Macdonald R, Barth A, Griffin K, Brand M, and Wilson SW.
Ace/Fgf8 is required for forebrain commissure formation and
patterning of the telencephalon. Development 127: 2549–2561,
2000.
186. Shihabuddin LS, Horner PJ, Ray J, and Gage FH. Adult
spinal cord stem cells generate neurons after transplantation in
the adult dentate gyrus. J Neurosci 20: 8727–8735, 2000.
187. Shihabuddin LS, Ray J, and Gage FH. FGF-2 is sufficient to
isolate progenitors found in the adult mammalian spinal cord.
Exp Neurol 148: 577–586, 1997.
188. Shimamura K and Rubenstein JL. Inductive interactions
direct early regionalization of the mouse forebrain. Development 124: 2709–2718, 1997.
189. Smallwood PM, Munoz-Sanjuan I, Tong P, Macke JP,
Hendry SH, Gilbert DJ, Copeland NG, Jenkins NA, and
Nathans J. Fibroblast growth factor (FGF) homologous factors: new members of the FGF family implicated in nervous
system development. Proc Natl Acad Sci USA 93: 9850–9857,
1996.
190. Spivak-Kroizman T, Lemmon MA, Dikic I, Ladbury JE,
Pinchasi D, Huang J, Jaye M, Crumley G, Schlessinger J,
and Lax I. Heparin-induced oligomerization of FGF molecules
is responsible for FGF receptor dimerization, activation, and
cell proliferation. Cell 79: 1015–1024, 1994.
191. Stanfield BB and Trice JE. Evidence that granule cells
generated in the dentate gyrus of adult rats extend axonal
projections. Exp Brain Res 72: 399–406, 1988.
192. Stapf C, Luck G, Shakibaei M, and Blottner D. Fibroblast
growth factor-2 (FGF-2) and FGF-receptor (FGFR-1) immunoreactivity in embryonic spinal autonomic neurons. Cell Tissue
Res 287: 471–480, 1997.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
159.
fibroblast growth factor 2. Proc Natl Acad Sci USA 95: 5672–
5677, 1998.
Otto D, Frotscher M, and Unsicker K. Basic fibroblast
growth factor and nerve growth factor administered in gel foam
rescue medial septal neurons after fimbria fornix transection.
J Neurosci Res 22: 83–91, 1989.
Ozawa K, Uruno T, Miyakawa K, Seo M, and Imamura T.
Expression of the fibroblast growth factor family and their
receptor family genes during mouse brain development. Brain
Res Mol Brain Res 41: 279–288, 1996.
Palmer TD, Takahashi J, and Gage FH. The adult rat
hippocampus contains primordial neural stem cells. Mol Cell
Neurosci 8: 389–404, 1997.
Pappas IS and Parnavelas JG. Basic fibroblast growth factor
promotes the generation and differentiation of calretinin neurons in the rat cerebral cortex in vitro. Eur J Neurosci 10:
1436–1445, 1998.
Peters K, Ornitz D, Werner S, and Williams L. Unique
expression pattern of the FGF receptor 3 gene during mouse
organogenesis. Dev Biol 155: 423–430, 1993.
Peters KG, Marie J, Wilson E, Ives HE, Escobedo J, Del
Rosario M, Mirda D, and Williams LT. Point mutation of an
FGF receptor abolishes phosphatidylinositol turnover and
Ca2⫹ flux but not mitogenesis. Nature 358: 678–681, 1992.
Peters KG, Werner S, Chen G, and Williams LT. Two FGF
receptors gene are differentially expressed in epithelial and
mesenchymal tissues during limb formation and organogenesis
in the mouse. Development 114: 233–243, 1992.
Peterson DA, Lucidi-Phillipi CA, Murphy DP, Ray J, and
Gage FH. Fibroblast growth factor-2 protects entorhinal layer
II glutamatergic neurons from axotomy-induced death. J Neurosci 16: 886–898, 1996.
Philippe JM, Garces A, and deLapeyiere O. Fgf-R3 is
expressed in a subset of chicken spinal motorneurons. Mech
Dev 78: 119–123, 1998.
Piccolo S, Sasai Y, Lu B, and De Robertis EM. Dorsoventral patterning in Xenopus: inhibition of ventral signals by
direct binding of chordin to BMP-4. Cell 86: 589–598, 1996.
Picker A, Brennan C, Reifers F, Clarke JD, Holder N, and
Brand M. Requirement for the zebrafish mid-hindbrain boundary in midbrain polarisation, mapping and confinement of the
retinotectal projection. Development 126: 2967–2978, 1999.
Pownall ME, Tucker AS, Slack JM, and Isaacs HV. eFGF,
Xcad3 and Hox genes form a molecular pathway that establishes the anteroposterior axis in Xenopus. Development 122:
3881–3892, 1996.
Qian X, Davis AA, Goderie SK, and Temple S. FGF2 concentration regulates the generation of neurons and glia from
multipotent cortical stem cells. Neuron 18: 81–93, 1997.
Raballo R, Rhee J, Lyn-Cook R, Leckman JF, Schwartz
ML, and Vaccarino FM. Basic fibroblast growth factor (Fgf2)
is necessary for cell proliferation and neurogenesis in the developing cerebral cortex. J Neurosci 20: 5012–5023, 2000.
Ragsdale CW and Grove EA. Patterning the mammalian
cerebral cortex. Curr Opin Neurobiol 11: 50–58, 2001.
Ray J, Peterson DA, Schinstine M, and Gage FH. Proliferation, differentiation, and long-term culture of primary hippocampal neurons. Proc Natl Acad Sci USA 90: 3602–3606,
1993.
Reifers F, Adams J, Mason IJ, Schulte-Merker S, and
Brand M. Overlapping and distinct functions provided by
fgf17, a new zebrafish member of the Fgf8/17/18 subgroup of
Fgfs. Mech Dev 99: 39–49, 2000.
Reifers F, Bohli H, Walsh EC, Crossley PH, Stainier DY,
and Brand M. Fgf8 is mutated in zebrafish acerebellar (ace)
mutants and is required for maintenance of midbrain-hindbrain boundary development and somitogenesis. Development
125: 2381–2395, 1998.
Reynolds BA, Tetzlaff W, and Weiss S. A multipotent EGFresponsive striatal embryonic progenitor cell produces neurons
and astrocytes. J Neurosci 12: 4565–4574, 1992.
Reynolds BA and Weiss S. Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science 255: 1707–1710, 1992.
R880
INVITED REVIEW
AJP-Regul Integr Comp Physiol • VOL
212. Vainikka S, Partanen J, Bellosta P, Coulier F, Birnbaum
D, Basilico C, Jaye M, and Alitalo K. Fibroblast growth
factor receptor-4 shows novel features in genomic structure,
ligand binding and signal transduction. EMBO J 11: 4273–
4280, 1992.
213. Vescovi AL, Reynolds BA, Fraser DD, and Weiss S. bFGF
regulates the proliferative fate of unipotent (neuronal) and
bipotent (neuronal/astroglial) EGF-generated CNS progenitor
cells. Neuron 11: 951–966, 1993.
214. Vicario-Abejon C, Johe KK, Hazel TG, Collazo D, and
McKay RDG. Functions of basic fibroblast growth factor and
neurotrophins in the differentiation of hippocampal neurons.
Neuron 15: 105–114, 1995.
215. Wagner J, Akerud P, Castro DS, Holm PC, Canals JM,
Snyder EY, Perlmann T, and Arenas E. Induction of a
midbrain dopaminergic phenotype in Nurr1-overexpressing
neural stem cells by type 1 astrocytes. Nat Biotechnol 17:
653–659, 1999.
216. Wagner JP, Black IB, and DiCicco-Bloom E. Stimulation of
neonatal and adult brain neurogenesis by subcutaneous injection of basic fibroblast growth factor. J Neurosci 19: 6006–6016,
1999.
217. Walicke PA and Baird A. Neurotrophic effects of basic and
acidic fibroblast growth factors are not mediated through glial
cells. Brain Res 468: 71–79, 1988.
218. Walshe J, Maroon H, McGonnell IM, Dickson C, and
Mason I. Establishment of hindbrain segmental identity requires signaling by FGF3 and FGF8. Curr Biol 12: 1117–1123,
2002.
219. Walshe J and Mason I. Expression of FGFR1, FGFR2 and
FGFR3 during early neural development in the chick embryo.
Mech Dev 90: 103–110, 2000.
220. Wanaka A, Johnson EM Jr, and Milbrandt J. Localization
of FGF receptor mRNA in the adult rat central nervous system
by in situ hybridization. Neuron 5: 267–281, 1990.
221. Wang JK, Xu H, Li HC, and Goldfarb M. Broadly expressed
SNT-like proteins link FGF receptor stimulation to activators
of Ras. Oncogene 13: 721–729, 1996.
222. Wang Q, Bardgett ME, Wong M, Wozniak DF, Lou J,
McNeil BD, Chen C, Nardi A, Reid DC, Yamada K, and
Ornitz DM. Ataxia and paroxysmal dyskinesia in mice lacking
axonally transported FGF14. Neuron 35: 25–38, 2002.
223. Wang Q, McEwen DG, and Ornitz DM. Subcellular and
developmental expression of alternatively spliced forms of fibroblast growth factor 14. Mech Dev 90: 283–287, 2000.
224. Weinstein M, Xu X, Ohyama K, and Deng CX. FGFR-3 and
FGFR-4 function cooperatively to direct alveogenesis in the
murine lung. Development 125: 3615–3623, 1998.
225. Weiss S, Dunne C, Hewson J, Wohl C, Wheatley M, Peterson AC, and Reynolds BA. Multipotent CNS stem cells are
present in the adult mammalian spinal cord and ventricular
neuroaxis. J Neurosci 16: 7599–7609, 1996.
226. White KE, Evans WE, O’Riordan JL, Speer MC, Econs
MJ, Lorenz-Depiereux B, Grabowski M, Meitinger T, and
Strom TM. Autosomal dominant hypophosphataemic rickets is
associated with mutations in FGF23. Nat Genet 26: 345–348,
2000.
227. Wilcox BJ and Unnerstall JR. Expression of acidic fibroblast
growth factor mRNA in the developing and adult rat brain.
Neuron 6: 397–409, 1991.
228. Wilkie AO, Slaney SF, Oldridge M, Poole MD, Ashworth
GJ, Hockley AD, Hayward RD, David DJ, Pulleyn LJ,
Rutland P, Malcolm S, Winter R, and Reardon W. Apert
syndrome results from localized mutations of FGFR2 and is
allelic with Crouzon syndrome. Nat Genet 9: 165–172, 1995.
229. Wilkinson DG, Bhatt S, and McMahon AP. Expression
pattern of the FGF-related proto-oncogene int-2 suggests multiple roles in fetal development. Development 105: 131–136,
1989.
230. Wilson PA and Hemmati-Brivanlou A. Vertebrate neural
induction: inducers, inhibitors, and a new synthesis. Neuron
18: 699–710, 1997.
231. Wilson SI and Edlund T. Neural induction: toward a unifying
mechanism. Nat Neurosci Suppl 4: 1161–1168, 2001.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
193. Stock A, Kuzis K, Woodward WR, Nishi R, and Eckenstein FP. Localization of acidic fibroblast growth factor in
specific subcortical neuronal populations. J Neurosci 12: 4688–
4700, 1992.
194. Storey KG, Goriely A, Sargent CM, Brown JM, Burns HD,
Abud HM, and Heath JK. Early posterior neural tissue is
induced by FGF in the chick embryo. Development 125: 473–
484, 1998.
195. Streit A, Berliner AJ, Papanayotou C, Sirulnik A, and
Stern CD. Initiation of neural induction by FGF signalling
before gastrulation. Nature 406: 74–78, 2000.
196. Streit A, Lee KJ, Woo I, Roberts C, Jessell TM, and Stern
CD. Chordin regulates primitive streak development and the
stability of induced neural cells, but is not sufficient for neural
induction in the chick embryo. Development 125: 507–519,
1998.
197. Suhonen JO, Peterson DA, Ray J, and Gage FH. Differentiation of adult hippocampus-derived progenitors into olfactory
neurons in vivo. Nature 383: 624–627, 1996.
198. Sun X, Mariani FV, and Martin GR. Functions of FGF
signalling from the apical ectodermal ridge in limb development. Nature 418: 501–508, 2002.
199. Sun X, Meyers EN, Lewandoski M, and Martin GR. Targeted disruption of Fgf8 causes failure of cell migration in the
gastrulating mouse embryo. Genes Dev 13: 1834–1846, 1999.
200. Sutherland D, Samakovlis C, and Krasnow MA. Branchless encodes a Drosophila FGF homolog that controls tracheal
cell migration and the pattern of branching. Cell 87: 1091–
1101, 1996.
201. Takagi Y, Nozaki K, Takahashi J, Yodoi J, Ishikawa M,
and Hashimoto N. Proliferation of neuronal precursor cells in
the dentate gyrus is accelerated after transient forebrain ischemia in mice. Brain Res 831: 283–287, 1999.
202. Takahashi J, Palmer TD, and Gage FH. Retinoic acid and
neurotrophins collaborate to regulate neurogenesis in adultderived neural stem cell cultures. J Neurobiol 38: 65–81, 1999.
203. Tanaka A, Miyamoto K, Minamino N, Takeda M, Sato B,
Matsuo H, and Matsumoto K. Cloning and characterization
of an androgen-induced growth factor essential for the androgen-dependent growth of mouse mammary carcinoma cells.
Proc Natl Acad Sci USA 89: 8928–8932, 1992.
204. Taupin P, Ray J, Fischer WH, Suhr ST, Hakansson K,
Grubb A, and Gage FH. FGF-2-responsive neural stem cell
proliferation requires CCg, a novel autocrine/paracrine cofactor. Neuron 28: 385–397, 2000.
205. Temple S. The development of neural stem cells. Nature 414:
112–117, 2001.
206. Teng YD, Mocchetti I, Taveira-DaSilva AM, Gillis RA,
and Wrathall JR. Basic fibroblast growth factor increases
long-term survival of spinal motor neurons and improves respiratory function after experimental spinal cord injury. J Neurosci 19: 7037–7047, 1999.
207. Tooyama I, Kawamata T, Walker D, Yamada T, Hanai K,
Kimura H, Iwane M, Igarashi K, McGeer EG, and McGeer
PL. Loss of basic fibroblast growth factor in substantia nigra
neurons in Parkinson’s disease. Neurology 43: 372–376, 1993.
208. Tretter YP, Hertel M, Munz B, ten Bruggencate G,
Werner S, and Alzheimer C. Induction of activin A is essential for the neuroprotecyive action of basic fibroblast growth
factor in vivo. Nat Med 6: 739–741, 2000.
209. Tropepe V, Sibilia M, Ciruna BG, Rossant J, Wagner EF,
and van der Kooy D. Distinct neural stem cells proliferate in
response to EGF and FGF in the developing mouse telencephalon. Dev Biol 208: 166–188, 1999.
210. Vaccarino FM, Schwartz ML, Raballo R, Nilsen J, Rhee J,
Zhou M, Doetschman T, Coffin JD, Wyland JJ, and Hung
YT. Changes in cerebral cortex size are governed by fibroblast
growth factor during embryogenesis. Nat Neurosci 2: 246–253,
1999.
211. Vaccarino FM, Schwartz ML, Raballo R, Rhee J, and
Lyn-Cook R. Fibroblast growth factor signaling regulates
growth and morphogenesis at multiple steps during brain development. Curr Top Dev Biol 46: 179–200, 1999.
R881
INVITED REVIEW
AJP-Regul Integr Comp Physiol • VOL
240. Yamaguchi TP, Harpal K, Henkemeyer M, and Rossant J.
fgfr-1 is required for embryonic growth and mesodermal patterning during mouse gastrulation. Genes Dev 8: 3032–3044,
1994.
241. Yamashita T, Yoshioka M, and Itoh N. Identification of a
novel fibroblast growth factor, FGF-23, preferentially expressed in the ventrolateral thalamic nucleus of the brain.
Biochem Biophys Res Commun 277: 494–498, 2000.
242. Ye W, Shimamura K, Rubenstein JL, Hynes MA, and
Rosenthal A. FGF and Shh signals control dopaminergic and
serotonergic cell fate in the anterior neural plate. Cell 93:
755–766, 1998.
243. Yoshimura S, Takagi Y, Harada J, Teramoto T, Thomas
SS, Waeber C, Bakowska JC, Breakefield XO, and Moskowitz MA. FGF-2 regulation of neurogenesis in adult hippocampus after brain injury. Proc Natl Acad Sci USA 98:
5874–5879, 2001.
244. Zhou M, Sutliff RL, Paul RJ, Lorenz JN, Hoying JB,
Haudenschild CC, Yin M, Coffin JD, Kong L, Krenias EG,
Luo W, Boivin GP, Duffy JJ, Pawlowski SA, and
Doetschman T. Fibroblast growth factor 2 control of vascular
tone. Nat Med 4: 201–207, 1998.
245. Zhu X, Komiya H, Chirino A, Faham S, Fox G, Arakawa T,
Hsu B, and Rees D. Three-dimensional structures of acidic
and basic fibroblast growth factors. Science 251: 90 –93, 1991.
246. Zigmond MJ, Bloom FE, Landis SC, Roberts JL, and
Squire LR. Fundamental Neuroscience. San Diego, CA: Academic, 1999.
247. Zimmerman L, De Jesus-Escobar J, and Harland R. The
Spemann organizer signal noggin binds and inactivates bone
morphogenetic protein 4. Cell 86: 599–606, 1996.
284 • APRIL 2003 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.6 on May 4, 2017
232. Wilson SI, Graziano E, Harland R, Jessell TM, and Edlund T. An early requirement for FGF signalling in the acquisition of neural cell fate in the chick embryo. Curr Biol 10:
421–429, 2000.
233. Woodward WR, Nishi R, Meshul CK, Williams TE, Coulombe M, and Eckenstein FP. Nuclear and cytoplasmic localization of basic fibroblast growth factor in astrocytes and
CA2 hippocampal neurons. J Neurosci 12: 142–152, 1992.
234. Wuechner C, Nordqvist AC, Winterpacht A, Zabel B, and
Schalling M. Developmental expression of splicing variants of
fibroblast growth factor receptor 3 (FGFR3) in mouse. Int J Dev
Biol 40: 1185–1188, 1996.
235. Xu J, Lawshe A, MacArthur CA, and Ornitz DM. Genomic
structure, mapping, activity and expression of fibroblast growth
factor 17. Mech Dev 83: 165–178, 1999.
236. Xu J, Liu Z, and Ornitz DM. Temporal and spatial gradients
of Fgf8 and Fgf17 regulate proliferation and differentiation of
midline cerebellar structures. Development 127: 1833–1843,
2000.
237. Xu X, Li C, Takahashi K, Slavkin HC, Shum L, and Deng
CX. Murine fibroblast growth factor receptor 1alpha isoforms
mediate node regression and are essential for posterior mesoderm development. Dev Biol 208: 293–306, 1999.
238. Xuan S, Baptista CA, Balas G, Tao W, Soares VC, and Lai
E. Winged helix transcription factor BF-1 is essential for the
development of the cerebral hemispheres. Neuron 14: 1141–
1152, 1995.
239. Yamaguchi TP, Conlon RA, and Rossant J. Expression of
the fibroblast growth factor receptor FGFR-1/flg during gastrulation and segmentation in the mouse embryo. Dev Biol 152:
75–88, 1992.