Functions and Regulations of Fibroblast Growth Factor Signaling During

Embed Size (px)

Citation preview

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    1/13

    Functions and regulations of fibroblast growth factor signaling during

    embryonic development

    Bernard Thisse, Christine Thisse *

    Institut de Genetique et de Biologie Moleculaire et Cellulaire, UMR 7104, CNRS/INSERM/ULP, 1 rue Laurent Fries, BP 10142, CU de Strasbourg,

    67404 ILLKIRCH cedex, France

    Received for publication 16 June 2005, revised 29 July 2005, accepted 5 September 2005

    Available online 10 October 2005

    Abstract

    Fibroblast growth factors (FGF) are secreted molecules which function through the activation of specific tyrosine kinases receptors, the FGF

    receptors that transduce the signal by activating different pathways including the Ras/MAP kinase and the phospholipase-C gamma pathways.

    FGFs are involved in the regulation of many developmental processes including patterning, morphogenesis, differentiation, cell proliferation or

    migration. Such a diverse set of activities requires a tight control of the transduction signal which is achieved through the induction of different

    feedback inhibitors such as the Sproutys, Sef and MAP kinase phosphatase 3 which are responsible for the attenuation of FGF signals, limiting

    FGF activities in time and space.

    D 2005 Elsevier Inc. All rights reserved.

    Keywords: FGF; FGF receptors; Ras/MAP kinase; HSPG; Sprouty; Sef; MKP3

    Introduction

    Fibroblast growth factors represent a large family of secreted

    molecules. Upon binding to their cognate receptors, FGFs

    activate signal transduction pathways which are required for

    multiple developmental processes. The wide-ranging biological

    roles of FGFs, the variety of signaling pathways activated and

    the complex and dynamic expressions of FGF ligands and

    receptors imply that FGF signaling must be tightly regulated.

    The scope of the present review is to provide readers with

    updated information about FGF signaling. After a short

    introduction on FGFs evolutionary history and structural

    characteristics and on FGF receptors (FGFRs) particularities,various functions of the FGF signaling will be presented,

    focusing on their implication in development. The last part of

    the review will describe the attenuation of FGF signaling mainly

    dependent on the Ras/MAP kinase signaling pathway, through

    the action of different factors, such as the Sproutys, Sef or MAP

    kinase phosphatase 3. Recent publication of various studies has

    led to controversial results and we will present the raising

    debate on how FGF can be regulated and on how multiple

    feedback regulators of the FGF pathway act at different levels ofthe signal transduction cascades to attenuate FGF signaling.

    Key structural properties of FGFs and FGF Receptors

    FGFs represent an extended family of polypeptide growth

    factors found through evolution, ranging from nematode to

    human whereas they have not been identified in unicellular

    organisms (for review,Itoh and Ornitz, 2004). Ininvertebrates,

    only three Drosophila genes (branchless,pyramus and thisbe)

    have been found and two (egl-17 and let-756) in Caenorhab-

    ditis elegans. In contrast, in vertebrates, a large number ofFGF

    genes has been identified: 10 FGFs in zebrafish (FGF2 4, 6,8,10 ,17a,17b,18,24), 6 inXenopus(FGF2 4,810), 13 in

    chicken (FGF1 4, 810, 12, 13, 16, 1820), 22 genes in

    mouse (FGF1 18, 2023) and human (FGF1 14, 1623).

    FGFgenes are scattered throughout the genome, indicating

    that the FGF gene family was generated both by gene and

    chromosomal duplication and translocation during evolution

    (for reviewOrnitz and Itoh, 2001).

    All FGFs share an internal core of similarity as well as a

    high affinity for heparin. FGFs act through binding and

    activation of FGFRs which are tyrosine kinase receptors that

    contain three immunoglobulin-like domains and a heparin-

    0012-1606/$ - see front matterD

    2005 Elsevier Inc. All rights reserved.doi:10.1016/j.ydbio.2005.09.011

    * Corresponding author.

    E-mail address: [email protected] (C. Thisse).

    Developmental Biology 287 (2005) 390 402

    www.elsevier.com/locate/ydbio

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    2/13

    binding sequence (Lee et al., 1989; Johnson et al., 1990,Fig.

    1). FGFR forms are expressed in two possible ways: by the

    expression of splice variants of a given FGFR gene or by the

    expression of different FGFR genes (Johnson and Williams,

    1993). Alternative splicing specifies the sequence of the

    carboxy-terminal half of the Ig domain III, resulting in either

    IIIb or IIIc isoforms (Miki et al., 1992; Chellaiah et al., 1994).

    This alternative splicing event is regulated in a tissue-specific

    manner and dramatically affects ligand receptor binding

    specificity (seePowers et al., 2000).

    Fig. 1. FGF receptors and FGF signal transduction. FGFRs are modular proteins comprising 3 immunoglobulin domains (IgI, IgII and IgIII). IgI and IgII are

    separated by an acidic box (AD). IgII contains a heparin binding domain (HBD). The IgIII domain is followed by a unique transmembrane (TM), a juxtamembrane

    (JM) and a kinase domain (KD) interrupted by an interkinase domain (IKD). FGF ligands linked to heparin sulfate proteoglycan (HSPG) bind to IgII and IgIII of

    FGFR. This results in the dimerization and the subsequent transactivation by phosphorylation of specific tyrosine residues. The main two transduction pathways

    involve the phospholipase C-g(PLCg) and the Ras/MAP kinase. The SH2 domain of the PLCginteracts with the phosphorylated Y766 of the activated receptor. The

    activated PLCg hydrolyzes the phosphatidyl-inositol-4,5-diphosphate (PIP2) to inositol-1,4,5-triphophate (IP3) and the diacylglycerol (DAG). IP3 releases Ca 2+

    while DAG activates the protein kinase C-y (PKCy). Activated PKCy activates Raf by phosphorylating its S338 and stimulates the downstream pathway in a Ras

    independent manner. The main pathway involves the interaction of the docking protein FRS2awith the amino-acid residues 407 433 (Xu et al., 1998). This protein

    is activated by phosphorylation on multiple tyrosine residues and subsequently interacts and activates Grb2 linked to Sos, a nucleotide exchange factor involved in

    the activation of Ras. Activated Ras then activates Raf which stimulates MEK which in turn phosphorylates the MAP kinase ERK. This last activated component

    translocates to the nucleus and phosphorylates specific transcription factors of the Ets family which in turn activate expression of specific FGF target genes. P:phosphorylation.

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402 391

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    3/13

    Role of Heparin Sulfate Proteoglycans in FGF ligand

    binding

    In addition to high affinity for their receptors, members of

    the FGF family have strong affinity for heparin (Burgess and

    Maciag, 1989). FGF FGFR interaction and signaling are

    further regulated by the spatial and temporal expression ofendogenous heparan sulfate proteoglycan (HSPG). HSPGs are

    cell-surface and extracellular matrix macromolecules that

    comprise a core protein to which heparan sulfate (HS)

    glycosaminoglycan (GAG) chains are attached. HSPGs play

    crucial roles in regulating key developmental pathways such as

    the FGF signaling (for review, Lin, 2004). The genetic

    evidences of such a role for HSPGs come from the isolation

    of mutants, both in invertebrates and vertebrates that show

    defective FGF signaling (Lin et al., 1999; Nybakken and

    Perrimon, 2002). Biochemical (Nakato and Kimata, 2002) and

    cristallographic studies (Schlessinger et al., 2000; Pellegrini et

    al., 2000) have revealed the importance of 60 sulfation of HSfor FGF signaling which appears to be required for FGF1

    FGFR2 and FGF2FGFR1 interactions. HSPGs were found to

    directly interact with FGFs and their receptors in a ternary

    complex on the cell surface (Ornitz, 2000; Pellegrini, 2001).

    One possibility could be that Heparin/HSPGs within the FGF

    signaling complex facilitate the FGFFGFR interaction. The

    second possibility could be that Heparin/HSPGs enhance the

    stability of a FGF FGFR complex. Examination of the

    spatiotemporal changes in HS to promote complexes formation

    with FGFRs and FGFs led to the finding that these changes

    depend on glycosyltransferases and sugar sulfotransferases.

    Furthermore, for each FGF FGFR pair examined, unique

    developmental HS binding patterns were identified that

    correlate with different HS binding requirements, as well as

    variations in FGF signaling. These results suggested that each

    FGFFGFR combination seeks distinct HS domains that are

    spatially and temporally regulated during development. These

    domains are unique to HS and distinguish it from heparin,

    which is highly sulfated and lacks any domain structure,

    explaining the ability of these FGFFGFR pairs to interact

    equally with heparin in vitro. Importantly, the HS activity

    necessary for ternary complex assembly does not represent the

    additive binding requirements of individual FGFs or FGFRs;

    rather, it represents requirements dictated by the synergistic

    interaction of the FGFs, HS and FGFRs. Finally, given that HSspecifically mediates each FGFFGFR interaction examined,

    these results suggest that developmental changes in HS may

    also specifically modulate the signaling of other families of

    morphogens and growth factors (Allen and Rapraeger, 2003).

    In addition, these interactions limit FGF diffusion and release

    into interstitial spaces, regulating FGF activity in a given tissue

    (Moscatelli, 1987; Flauennhaft et al., 1990).

    FGF transduction pathway

    FGFR transmits extracellular signals to various cytoplasmic

    signal transduction pathways through tyrosine phosphorylation

    (Fig. 1). FGFRs exist as inactivated monomers, which are

    activated when two FGF molecules connected by a heparan

    sulfate proteoglycan bind to the extracellular IgII and IgIII

    domains of the receptor leading to its homodimerization

    (Schlessinger et al., 2000). This dimerization brings together

    the intracellular domain of the receptor leading to transauto-

    phosphorylation of several critical tyrosine residues. Activation

    of the receptor allows proteins containing Src homology (SH2)or phosphotyrosine binding (PTB) domains to bind to sequence

    recognition motifs in FGFR1, resulting in phosphorylation and

    activation of these proteins (Pawson, 1995; Forman-Kay and

    Pawson, 1999; Dhalluin et al., 2000). For example, the

    Phospholipase C-g (PLCg) was identified to be associated

    with FGFR following ligand-dependent activation. This asso-

    ciation is due to binding between the SH2 domain of PLCgand

    Tyr 766 of FGFR1. Activated PLCg can hydrolyze phospha-

    tidylinositol-4,5-diphosphate (PIP2) to inositol-1, 4, 5-triphos-

    phate (IP3) and diacylglycerol (DAG). IP3 induces Ca2+

    release from intracellular stores, whereas DAG is a protein

    kinase C activator. The activated protein kinase C-y (PKCy) isable to phosphorylate and to stimulate Raf therefore leading to

    activate the MAP kinase pathway in a Ras-independent manner

    (Ueda et al., 1996).

    The main signaling activated through the stimulation of

    FGFRs is the Ras/MAP kinase pathway. Links between

    elements of this pathway and the activated FGFR involve the

    interaction of the juxtamembrane domain of FGFR with the

    FGFR substrate 2a (FRS2a) which is a membrane-anchored

    docking protein. FRS2 lacks an SH2 domain, but contains a

    phosphotyrosine binding domain (Kouhara et al., 1997), which

    was shown to mediate phosphotyrosine-independent interac-

    tion with amino-acid residues 407433 in FGFR1. Activation

    of FGFR1 leads to tyrosine phosphorylation of FRS2 at several

    sites. This allows the binding of Grb2, a small adaptor

    molecule, which exists in complex with the nucleotide

    exchange factor Sos and is involved in the activation of the

    GTP-binding protein Ras. In this complex, Sos catalyzes the

    exchange of GDP for GTP on Ras, promoting Ras activation as

    well as downstream elements of the pathway, comprising Raf,

    MEK and MAP kinases (ERK1, 2), the last member of which

    finally enters the nucleus and phosphorylates target transcrip-

    tion factors (Sternberg and Alberola-Ila, 1998). Among these

    are the Ets domain containing factors ERM and PEA3

    (Wasylyk et al., 1998; Sharrocks, 2001). They belong to a

    family of transcription factors, which share a winged-helixloophelix domain with which they bind to DNA as

    monomers. They are able to form heterocomplexes with

    transcription factors and are activated most prominently by

    the MAP kinase(Wasylyk et al., 1998).

    A study performed during the course of somite development

    revealed the contribution of these Ets-domain containing

    proteins activated by FGFs. The FGF signal secreted from

    the myotome induces formation of a scleraxis (Scx)-expressing

    tendon progenitor population in the sclerotome. Brent and

    Tabin (2004) showed that transcriptional activation by Pea3

    and Erm in response to FGF signaling is both necessary and

    sufficient for Scx expression in the somite. They proposed that

    the domain of the somitic tendon progenitors is regulated both

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402392

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    4/13

    by the restricted expression of Pea3 and Erm, and by the

    precise spatial relationship between these Ets transcription

    factors and the FGF signal originating in the myotome.

    Implication of FGF signaling during development

    FGFs were described to be implicated in multiple roles suchas patterning, morphogenesis, differentiation, regulation of cell

    proliferation or migration (for review see Goldfarb, 1996;

    Powers et al., 2000; Coumoul and Deng, 2003). In the next

    paragraphs, we will illustrate major functions of FGF signaling

    during development.

    Early patterning and dorso-ventral axis formation

    Induction and patterning of mesoderm is one of the earliest

    events in which FGF signaling was revealed to be essential.

    The first mesodermal inducer identified was basic FGF (Slack

    et al., 1987), for which its role is evolutionarily conserved.The effect of FGF on mesoderm formation was initially

    studied in Xenopus in which not only FGF was described to

    have a mesoderm inducing capacity (Slack et al., 1987;

    Kimelman and Kirschner, 1987; Kimelman et al., 1988) but

    components of the FGF signaling pathway such as FGFRs,

    HSPGs, FRS2, Grb2 and the Ras downstream cascade, when

    inhibited, were shown to block mesoderm formation and to

    induce both gastrulation and posterior defects (Whitman and

    Melton, 1992; MacNicol et al., 1993; Umbhauer et al., 1995;

    LaBonne, 1995; Gotoh, 1995).

    Involvement of FGFs in the establishment of the dorso-

    ventral (D/V) axis has been extensively studied using the

    zebrafish model. Genetic analysis revealed that BMP2b/BMP7

    heterodimers (Schmid et al., 2000)act as morphogens secreted

    ventrally and are responsible for the specification of ventral cell

    fates. The expression of BMPgenes, initially activated in the

    whole blastula becomes progressively restricted ventrally. This

    ventral restriction coincides with the spreading of FGF activity

    from the dorsal side of the embryo, suggesting an implication

    of FGF signaling in the dorsal downregulation of BMP gene

    expression. Consistently, general activation of the FGF/Ras/

    MAP kinase signaling pathway inhibits BMP gene expression

    in the whole blastula. Conversely, inhibition of FGF signaling

    causes BMPgene expression to enlarge dorsally, leading to an

    expansion of ventral cell fates. Altogether, this shows that FGFsignaling is essential to delimit the expression domain ofBMPs

    in blastula stage embryos. Therefore, FGFs act upstream of the

    ventral morphogens and function as initial signals for the

    establishment of the D/V patterning (Furthauer et al., 2004).

    FGF signaling was also identified to be implicated in

    ventro-posterior mesodermal tissue maintenance. Consistently,

    blocking FGF signaling by dominant-negative FGFR results in

    deletions of tail and trunk structures (Amaya et al., 1991;

    Griffin et al., 1995). As well, targeted disruption of mouse

    FGFR1 leads to the disorganization of axial mesoderm and to a

    defective development of the posterior structures. FGFs control

    the specification and maintenance of mesoderm by regulation

    of T box genes (Griffin et al., 1995; Smith et al., 1991; Strong

    et al., 2000; Sun et al., 1999; Ciruna and Rossant, 2001; Griffin

    et al., 1998; Zhao et al., 2003). One FGF mesodermal target is

    brachyury, required for posteriormesoderm and axis formation

    in mouse, zebrafish andXenopus(Smith et al., 1991; Herrmann

    et al., 1990; Halpern et al., 1993; Conlon et al., 1996). In

    Xenopus, Brachyury induces eFGF, which is required for

    myogenic celllineage (Stanley et al., 2001; Fisher et al., 2002)and vice versa (Isaacs et al., 1994; Schulte-Merker and Smith,

    1995).

    Role of FGFs on cell movements

    Evidence has been provided that FGF signaling plays a

    direct role in cell movements during convergence extension

    (Nutt et al., 2001). In the mouse, FGF signaling may affect

    gastrulation movements via the transcription factor snail

    (Ciruna and Rossant, 2001). FGFR1 mutant embryos die at

    late gastrulation, showing defects in cell migration, cell fate

    specification and patterning. FGFR1 mutant cells may fail tomigrate properly because they maintain high levels of E-

    cadherin resulting from a snail downregulation. An additional

    study performed in the mouse revealed that, in the absence of

    both FGF8 and FGF4, epiblast cells move into the streak and

    undergo an epithelial-to-mesenchymal transition; however,

    most cells then fail to move away from the streak. As a

    consequence, no embryonic mesoderm- or endoderm-derived

    tissues develop demonstrating that signaling via FGF8 and/or

    FGF4 is required for cell migration away from the primitive

    streak(Sun et al., 1999).

    Moreover, there is an increasing evidence for a direct

    chemotactic response of migrating cells in response to FGF

    signals. As such, FGF2 and FGF8 are potent chemoattractants

    in migration of mesencephalic neural crest cells (Kubota and

    Itoh, 2000) and FGF2 and FGF4 attract mesenchymal cells

    during limb bud development in the mouse(Webb et al., 1997;

    Li and Muneoka, 1999). Also, FGF10 exerts a potent

    chemoattractive effect to direct lung distal epithelial buds to

    their destination(Park et al., 1998). Furthermore, FGFs can act

    as chemotactic signals to coordinate cell movements during

    gastrulation. In the chick, cells can be guided by environmental

    signals and beads coated with FGFs alter their trajectories. The

    observed movement patterns of anterior streak cells can be

    explained by an FGF8-mediated chemorepulsion of cells away

    from the streak followed by chemoattraction towards an FGF4signal produced by the forming notochord(Yang et al., 2002).

    Neural induction and FGF signaling

    Several studies on neural induction have led to the ?default

    modelX which proposes that ectodermal cells are fated to

    become neural by default. In this model, cells are normally

    inhibited from a neural fate by BMPs that are expressed in the

    ectoderm, from which they must be released for neural

    induction to occur (for reviews, Hemmati-Brivanlou and

    Melton, 1997; Weinstein and Hemmati-Brivanlou, 1999).

    Additional data have suggested that FGF signaling plays a

    crucial role for an early step of neural induction in the chick. In

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402 393

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    5/13

    this specie, induction of neural tissue precedes the formation of

    the Hensens node. FGFs are sufficient to induce preneural

    markers and inhibition of FGF signaling blocks neural

    induction by Hensens node (Streit et al., 2000; Wilson et al.,

    2000; Alvarez et al., 1998; Storey et al., 1998). FGF signals

    repress BMP expression which leads to induction of neural

    tissue. On the contrary, when FGFs are inhibited, BMPexpression is maintained and neural fate is blocked (Wilson

    et al., 2000). FGFs may block BMP signaling during neural

    induction of the chick by inducing Churchill which is required

    for the expression of Smad-interacting protein 1 (Sip1) in the

    neural plate. Sip1 binds and represses Smad1/5 blocking

    therefore BMP signaling (Sheng et al., 2003).

    In Xenopus, the role of FGF signaling in neural induction

    has also been intensively studied. However, its implication in

    neural induction has led to conflicting results. Several

    investigators suggested that FGF signaling is required directly

    for neural induction (Launay et al., 1996; Hongo et al., 1999;

    Streit et al., 2000; Bertrand et al., 2003). In most cases,however, experiments were performed in a background where

    the BMP signaling was partially attenuated (Kengaku and

    Okamoto, 1995; Lamb and Harland, 1995; Ribisi et al., 2000).

    In addition, several studies using dominant negative FGFR1

    have led to different results showing that neural tissue forms

    even when FGF signaling is inhibited (Kroll and Amaya, 1996;

    Ribisi et al., 2000; Barnett et al., 1998; McGrew et al., 1997;

    Holowacz and Sokol, 1999). It appears that FGF signaling

    required for anterior neural development may be mediated

    through FGFRs other than FGFR1, such as through FGFR4a

    (Hongo et al., 1999).

    In contrast, other studies report a requirement for FGFs in

    the antero-posterior patterning rather than the initial induction

    of the neural plate(Holowacz and Sokol, 1999). FGFs are able

    to change anterior neural fate to more posterior neural cell

    types (Umbhauer et al., 2000; Lamb and Harland, 1995; Ribisi

    et al., 2000; Holowacz and Sokol, 1999; Cox and Hemmati-

    Brivanlou, 1995). This may thus involve interactions with the

    Wnt signaling pathway (McGrew et al., 1997; Kazanskaya et

    al., 2000; Domingos et al., 2001; Kudoh et al., 2002).

    Finally, a controversial study provides evidence that BMP

    inhibition is required in the chick, but only as a relatively late

    step. BMP inhibition would not be sufficient to cause

    competent cells to acquire neural fates either in chick or in

    Xenopus. In the frog, FGF synergizes with BMP inhibition toinduce neural markers. In chick, inhibition of BMP signaling

    with Wnt antagonists and/or FGF is not sufficient for neural

    induction. Neural induction would not occur by default but

    would rather involve a succession of signaling events with

    some players which would remain to be identified(Linkerand

    Stern, 2004).

    Midbrain hindbrain patterning

    Early patterning of the vertebrate presumptive midbrain and

    rhombomere 1 (rh1) is regulated by a local organizer located at

    the mid/hindbrain junction(Liu and Joyner, 2001), a territory

    expressing FGF8. This factor has organizer properties and is

    required for midbrain and cerebellum development(Meyers et

    al., 1998, Reifers et al., 1998). FGF17 and FGF18 are also

    expressed in the mid/hindbrain junction in broader domains

    than FGF8, including posterior midbrain(Maruoka et al., 1998;

    Xu et al., 1998). Loss-of-function of FGF17 in the mouse

    results in truncation of the posterior midbrain and reduced

    proliferation of the anterior cerebellum. Removal of one copyof fgf8 in a fgf17 mutant background leads to exaggerated

    cerebellum phenotype(Xu et al., 2000). To complicate matters,

    fgf8 is differentially spliced to generate FGF8a and FGF8b

    isoforms. FGF8a misexpression causes expansion of the

    midbrain and FGF8b transforms the midbrain into cerebellum.

    The initial effect of FGF8b is to reduce growth of the midbrain.

    Therefore, FGF8a and FGF8b have distinct activities, both on

    growth and regulation of gene expression (reviewed in Liu and

    Joyner, 2001; Sato et al., 2001).

    Interactions of FGF8a and FGF8b with FGF17 and FGF18

    have been studied. It appears that FGF8b after being induced in

    the presumptive rh1 territory induces FGF18 in surroundingtissue, further extending the gradient of FGF RNA expression.

    FGF8b maintains two negative feedback loops by inducing the

    expression of the negative feedback FGF inhibitors Spry1 and

    Spry2 and repressing FGFR2 and FGFR3 (Liu et al., 2003).

    FGF17, FGF18 proteins and possibly FGF8a as well as a low

    level of FGF8b then regulate proliferation of the midbrain and

    cerebellum(Liu et al., 2003; Chi et al., 2003).

    Limb induction and morphogenesis

    The role of FGFs has been established in induction,

    initiation and maintenance of limb development (reviewed in

    Martin, 1998). Limb bud initiation is triggered by FGF8

    inducing the expression of FGF10. FGF10 then in turn induces

    FGF8 in ectodermal cells resulting in the formation of the

    apical ectodermal ridge (AER). FGFs from the AER maintain

    cell proliferation in the progress zone (a population of

    undifferentiated mesenchymal cells located near the distal tip

    of the limb bud, adjacent to the AER). FGF2 (Fallon et al.,

    1994), FGF4 (Laufer et al., 1994)and FGF8(Crossley etal.,

    1996) induce sonic hedgehog (shh) expression in the zone of

    polarizing activity (ZPA). Outgrowth and patterning of the limb

    result from the combined effects of FGF and Shh and

    regulation of various genes, including members of the Hox

    family (reviewed inMartin, 1998).A first model proposed for explaining the patterning of the

    limb development along the proximaldistal (PD) axis is the

    progress zone model(Summerbell et al., 1973) which explains

    that as the limb grows, cells are pushed out of the progress zone

    and become fixed in the positional value that they acquired.

    Cells that leave the proximal zone early will adopt a proximal

    fate and cells that leave the progress zone late will form distal

    elements (Sauders, reviewed in Niswander, 2003). The

    conclusion from this model is that the PD fate is specified

    progressively. Recent studies have provided an alternative

    model on how the AER may influence the PD pattern. This

    second model proposes that all PD fates are already specified

    within the early limb bud rather than progressively (Dudley et

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402394

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    6/13

    al., 2002). In the absence of FGF8 in mice, the proximal

    element of the limb is reduced or even lost. However, the distal

    elements are formed quite normally because other FGFs are

    activated (Lewandoski et al., 2000; Moon and Capecchi, 2000).

    These results of loss of proximal but not of distal elements

    cannot be interpreted within the progress zone model. In

    addition, new experimental data have proposed that FGFsfunction is first to influence the initial size of the limb (Sun et

    al., 2002). In FGF8 mutant mice, early limb bud is smaller.

    This phenotype is detected right after FGF8 expression is

    detectable, excluding the possibility that FGF8 alters cell

    proliferation or cell death. Instead, FGF8 would (by affecting

    morphogenetic movements or cell adhesion) influence the

    number of cells in the limb bud by preventing their death. In

    mutant limbs, death of proximal but not distal mesenchyme can

    be observed. The PD precartilage condensations are formed but

    smaller. Sun et al. (2002)proposed that the AER generates an

    adequate number of mesenchymal cells for correctly sized

    condensations to form. When the AER is disrupted, the numberof skeletal elements is decreased resulting in smaller or in

    absence of condensations arguing that the FGFs function is to

    control mesenchymal cell number but not the PD patterning.

    Bone formation

    The importance of FGF signaling in skeletal development

    was first revealed with the discovery that a point mutation in

    the transmembrane domain of FGFR3 is the etiology of

    achondroplasia, the most common genetic form of dwarfism

    in human (Rousseau et al., 1994; Shiang et al., 1994). A major

    role of FGFs and FGFRs has then been scored by finding

    missense mutations resulting in more than 15 human disorders,

    ranging from skeletal dysplasias, craniosysnostosis syndromes

    or short stature (for reviewCoumoul and Deng, 2003). In the

    bone, FGF2 and FGF9 transcripts are found in mesenchymal

    cells and osteoblats (Gonzalez et al., 1996; Kim et al., 1998).

    FGF18 is expressed in mesenchymal cells, in differentiating

    osteoblasts during calvarial bone development and in the

    perichondrium of long bones (Ohbayashi et al., 2002). The

    action of FGFs is dependent on the spatiotemporal pattern of

    expression of FGFRs like FGFR1 and FGFR2 which are

    located in mesenchymal cells during condensation of mesen-

    chyme prior to deposition of bone matrix at early stages of long

    bone development and in the cranial sutures (Orr-Urtreger etal., 1991; Delezoide et al., 1998). Later, they are found in

    preosteoblasts and osteoblasts together with FGFR3 (Molteni

    et al., 1999). Overexpression of FGF2 in mouse induces

    abnormal bone formation (Coffin et al., 1995)whereas its loss-

    of-function leads to inhibition of bone formation (Monteroet

    al., 2000). Mice lacking FGF18 display ossification and cranial

    sutures closure delay (Liu et al., 2002; Ohbayashi et al., 2002).

    Altogether, FGF signaling appears to regulate cell proliferation

    and differentiation positively in membranous and long bone

    osteogenesis. FGF2 was shown also to act in vivo on osteoblast

    precursor cells replication to increase osteoblast number. FGF

    signaling also controls genes involved in osteoblast apoptosis.

    High FGF signaling may reduce apoptosis of immature

    osteoblasts and thereby enhance the osteoblast population,

    whereas continuous signaling may promote apoptosis in more

    mature osteoblasts, therefore limit the early increase in the

    osteoblasts pool. FGFs interact with other growth factors

    signaling pathways to regulate osteoblast function. For

    example, FGF2 and FGFR2 inhibit the expression of the

    BMP antagonist Noggin in the cranial sutures, resulting inincreased BMP4 activity and suture fusion (Warren et al.,

    2003). Therefore, FGF signaling can control cranial suture

    fusion indirectly through BMP signaling. FGFs regulate also

    genes involved in matrix degradation(Kawaguchi et al., 1995)

    and induce the expression of tissue inhibitors of metallopro-

    teinases (Delany and Canalis, 1998; Varghese et al., 1995)

    indicating that FGFs (mainly FGF2) may modulate bone matrix

    proteolysis by regulating collagenase activity. Overall, activa-

    tion of FGF signaling isable to regulate genes involved at all

    steps of osteogenesis (Fig. 2). This array of genes may

    participate in the regulation of cell progression from osteopro-

    genitor cells to the end of the osteoblast life (for review, Marie,2003). Further studies have led to the identification of signaling

    pathways that are involved in FGF signaling in osteoblasts,

    mainly MAP kinase signal transduction cascade(Mansukhani

    et al., 2000; Debiais et al., 2001).

    Attenuation of FGF signaling

    The multiple range of biological effects of FGFs and the

    variety of signaling pathways activated by this family of

    ligands imply that FGF signaling must be tightly regulated

    regarding timing, duration and spread of the signal. This can be

    achieved by both positive and negative feedback loops. FGF

    signaling can be regulated at many levels in the extracelular

    space and within the cell. A growing number of proteins have

    been identified that specifically regulate the activity of FGF

    signaling pathways through FGFRs. Many of these regulatory

    factors belong to the FGF synexpression group (Furthauer et

    al., 2002; Niehrs and Meinhardt, 2002; Tsang et al., 2002). Not

    only are these factors coexpressed with FGFs but they are

    themselves regulated by FGF signaling and inhibit FGF

    signaling by establishing a negative feedback loop.

    The Sprouty proteins

    The first feedback regulator of the FGF pathway isolated isSprouty (Spry). Spry was originally identified as an antagonist

    of FGF signaling that patterns apical branching in Drosophila

    tracheae (Hacohen et al., 1998). In spry mutants, ectopic

    airway branches are induced due to excess of FGF signaling.

    Spry was then described to be a general inhibitor of receptor

    tyrosine kinase (RTK) signaling (for review,Kim and Bar-Sagi,

    2004). Spry proteins are evolutionarily conserved with four

    members (Spry1, 2, 3, 4) identified in mammals (Dikic and

    Giordano, 2003). These proteins share a conserved carboxy-

    terminal cystein-rich domain, necessary for their specific

    localization and function. Their amino-terminal part is diver-

    gent except that they possess an invariant tyrosine phospho-

    rylation site. It is thought that this sequence divergence dictates

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402 395

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    7/13

    various functions by mediating different protein protein

    interactions. Gain- and/or loss-of-functions studies in the

    mouse, Xenopus and zebrafish (Minowada et al., 1999;

    Mailleux et al., 2001; Nutt et al., 2001; Furthauer et al.,

    2001, 2004) have demonstrated that Sprys antagonize FGF

    signaling.Spryexpression is induced by the FGF signaling and

    has various biological consequences, including regulation of

    cell proliferation and differentiation, endocytic sorting of

    activated RTKs, control of cytoplasmic calcium concentration,

    migratory behavior of cells, retro-control of D/V patterning and

    lung and bone development. Spry genes contribute also to

    reciprocal epithelial mesenchymal and stromal signaling

    controlling ureteric branching, which involves the coordination

    of Fgf/Wnt11/Gdnf pathways during mammalian kidney

    development(Chi et al., 2004).

    Study of Sprys function has led to intense investigation and

    controversy(Fig. 3). Works from many groups have identified

    Sprys as negative regulators of the MAP kinase ERK.

    However, contradictory reports have identified Spry inhibitionupstream of Ras(Casci et al., 1999; Hanafusa et al., 2002) or at

    the level of Raf(Reich et al., 1999; Sasaki et al., 2003; Yusoff

    et al., 2002). Moreover, Spry family members have varying

    interacting partners, such as c-Cbl, Grb2, Raf1 and FRS2 (for

    review, Dikic and Giordano, 2003).

    The mechanisms by which Sprys interfere with ERK

    activation have been extensively studied on EGF signaling

    pathway and have come from the finding that Spry proteins are

    regulated by tyrosine phosphorylation (Tyr55 in spry2, Tyr53

    in Spry1/4) in response to stimuli. Phosphorylated Tyr55 serves

    as a docking site for the SH2 domain of c-Cbl (Egan et al.,

    2002; Wong et al., 2002; Rubin et al., 2003; Hall et al., 2003 ).

    Usually, c-Cbl binds to phosphorylated tyrosines on the

    activated EGFR, which in turn results in its ubiquination,

    endocytosis and degradation through theproteasomal/lysosom-

    al pathways (Dikic and Giordano, 2003). As Spry proteins

    phosphorylated on Tyr 55 can also interact with c-Cbl, they are

    in direct competition with activated EGFRs for binding to this

    protein. As a result, the EGFR is not ubiquinated, internalized

    or degraded in the presence of Spry, leading to prolonged cell-

    surface exposure of the receptor and sustained signaling

    activity. Therefore, in this context, Sprys function as positive

    regulators of the EGF signaling resulting in an upregulation of

    the MAP kinase ERK activity. A similar degradation of Spry2

    by c-Cbl may also be true for FGFR signaling (Hall et al.,

    2003).

    However, most of the studies performed on FGF signaling

    pathway show that Sprys associate with Grb2. For example,

    mouse Spry2 and Xspry1 can interact with the SH2 domain of

    Grb2 after FGFR-induced phosphorylation of Tyr55(Hanafusa

    et al., 2002). As a result, the interaction of Grb2 with FRS2 is

    blocked and FGF-induced signal transduction is repressed. Inaddition, Spry2 and 4 have been shown to bind directly to Raf

    through a Raf-binding motif in their C-terminal domain and

    suppress phosphorylation of Raf on Ser338 by the Raf kinase

    PKCy (Sasaki et al., 2003).

    Altogether, Sprys function through different pathways,

    which determine the overall effect on RTK signaling. First,

    binding of c-Cbl to tyrosine-phosphorylated Spry results in the

    stabilization of EGFR and in sustained ERK activity. Second,

    binding of Grb2 to tyrosine-phosphorylated Spry can interrupt

    the FGFR/Ras/Raf/MAPK pathway upstream of Ras. Finally,

    binding of the C-terminal domain of Spry by its Raf-binding

    motif inhibits Raf phosphorylation by PLCg-activated PKCy.

    However, the criteria to decide between Spry different

    Fig. 2. FGF signaling regulates osteogenesis. During intramembranous bone development, bone formation is characterized by the proliferation of mesenchymal cells

    followed by their commitment to become osteoprogenitor cells. They differentiate into preosteoblasts then into bone matrix-forming mature osteoblasts. FGFs are

    required for each step of bone formation through FGFRs to control various genes involved in osteoblast commitment, differentiation and apoptosis. ALP: alkaline

    phosphatase, BSP: bone sialoprotein, COL I: collagen type I, OC: osteocalcin, OP: osteospondin, ON: osteonectin (adapted from P. J.Marie, 2003).

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402396

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    8/13

    Fig. 3. Models for the attenuation of FGF signaling by Sprys. (I) Spry acts downstream of the FGFR and upstream of Ras by binding to the adaptor protein Grb2 at the le

    the activation of Ras and the subsequent stimulation of the downstream elements of the pathway. (II) Spry also acts downstream of Ras by directly binding Raf. This is ach

    located in its C-terminal part and this prevents both activation by Ras and phosphorylation of Ser338 by the PKC y. This explains how spry can inhibit the transduct

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    9/13

    functions are not yet known and will have to be defined in the

    near future.

    Other FGF synexpression group members

    In addition to sprys, several other genes involved in the

    regulation of the FGF pathway have been isolated through thecourse of large-scale in situ hybridization screens (Gawantka et

    al., 1998; Thisse et al., 2001; Kudoh et al., 2001) designed to

    identify genes showing expression pattern similar or over-

    lapping with FGFs. Various genes were isolated such as

    XFLRT3 (Bottcher et al., 2004), Sef (Furthauer et al., 2002;

    Tsang et al., 2002), the MAP kinase phosphatase 3 (MKP3,

    also called Pyst1, Eblaghie et al., 2003; Tsang et al., 2004),

    PEA3 (for polyoma virus enhancer activator 3) and ERM

    (Munchberg et al., 1999). These last two factors are the

    downstream players of the FGF signaling pathway while the

    others, also induced by FGF stimulation are involved in the

    attenuation of the signal (for review, Tsang and Dawid, 2004).

    XFLRT3

    This gene encodes a transmembrane protein characterized

    by a cluster of leucine-rich repeats and one FNIII domain

    within the extracellular region. Its expression is induced after

    activation of FGF signaling and downregulated after inhibition

    of this pathway. XFLRT3 signaling results in phosphorylation

    of the MAP kinase ERK and is blocked by the MAP kinase

    phosphatase 1. In gain- and loss-of-function experiments

    performed in Xenopus, FLRT3 was shown to phenocopy the

    Fig. 4. FGF signaling and its regulation. The stimulation of FGFR by FGF ligands results in the activation of specific target genes. Among them are several feedback

    inhibitors which are involved in the attenuation of FGF signaling. Spry acts at the level of Grb2 and/or at the level of Raf. Sef and XFLRT3 are both located at the

    membrane and interact with FGFR. Sef functions as a negative regulator while XFLRT3 enhances the FGF activity resulting in the phosphorylation of the MAP

    kinase ERK. Sef was shown also to affect the phosphorylation of the MAP kinase, either directly at the level of ERK or by preventing the phosphorylation of ERK byMEK. Finally, MKP3 negatively regulates the FGF signaling pathway by dephosphorylation of activated MAP kinase.

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402398

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    10/13

    FGF signaling (Bottcher et al., 2004). Surprisingly, in zebra-

    fish, although FLRT3 is expressed in well known centers of

    FGF signaling, neither gain nor loss of FGF signaling appears

    to directly affect FLRT3 expression. Inactivating FLRT3

    through injection of antisense morpholinos impairs conver-

    gence-extension movements that direct axial elongation but

    fails to mimick phenotypes observed after either loss or gainfunction of FGF activity. These observations suggest that, in

    zebrafish, FLRT3 may be involved in directing cell movements

    but does not appear to be related to the FGF pathway

    (Furthauer M., Thisse, B. and Thisse, C., unpublished data).

    Sef

    The Sef protein is conserved across zebrafish, mouse and

    human but not in invertebrates. Homology searches with Sef

    revealed that it encodes a putative transmembrane protein, with

    weak similarities to the Interleukin-17 receptor and a putative

    tyrosine phosphorylation site into the intracellular region(Furthauer et al., 2002; Tsang et al., 2002). In zebrafish, loss-

    or gain-of-function studies revealed Sef to function as an

    antagonist of FGF signaling and to interfere with FGF signal

    transduction by acting dowstream or at the level of MEK and

    upstream (or at the level) of the MAP kinase (Furthauer et al.,

    2002). Further studies performed using human Sef (hSef)

    showed that it binds to activated forms of MEK, inhibits the

    dissociation of the MEK ERK complex and blocks nuclear

    translocation of activated ERK. Consequently, hSef inhibits

    phosphorylation and activation of the nuclear ERK substrate

    Elk-1, while it does not affect phosphorylation of the

    cytoplasmic ERK substrate RSK2. Downregulation of endog-

    enous hSef by hSef siRNA enhances the stimulus-induced

    ERK nuclear translocation and the activity of Elk-1. These

    results favor the hypothesis that hSef acts as a spatial regulator

    for ERK signaling by targeting ERK to the cytoplasm (Toriiet

    al., 2004). Other reports showed that the intracellular domain of

    Sef is required for its inhibitory function and its interaction

    with FGFR1 and FGFR2 (Tsang et al., 2002; Kovalenko et al.,

    2003, Xiong et al., 2003). Ectopic expression of Sef blocks the

    phosphorylation of FRS2. The conclusion of these studies was

    that Sef acts at the level of FGFRs blocking the Ras and PI3K

    pathways downstream. To complicate the further understanding

    of Sef function, alternative spliced isoforms of Sef were found

    in human (Preger et al., 2004). Sef-b which lacks the signalpeptide is localized in the cytoplasm and can suppress FGF

    signaling downstream of MEK. Because FGFs deliver multiple

    biological responses, it is possible that like Spry, Sef can inhibit

    the FGF pathway at various points depending on the context, in

    order to allow fine-tuning of signal regulation.

    MAP kinase phosphatase 3 (MKP3)

    Two distinct domains are characteristic of MKP3 proteins,

    the N-terminal region, which contains a high-affinity ERK/

    MAPK binding domain and the C-terminal domain, which

    constitute a dual specificity phosphatase (Stewart et al., 1999;

    Zhang et al., 2003; Zhao and Zhang, 2001). MKP3 negatively

    regulates the MAP kinase cascade through dephosphorylation

    of activated MAPK proteins. Upon binding of phosphorylated

    MAPKs to the MAPK-binding domain in MKP3, a conforma-

    tional activation of the C-terminal phosphatase domain is

    achieved, leading to the inactivation of MAPKs (Camps et al.,

    2000; Fjeld et al., 2000; Zhao and Zhang, 2001). In chicken,

    ectopic expression of MKP3 results in the disruption of limboutgrowth, a phenotype characteristic of FGF signaling

    blockage (Kawakami et al., 2003; Eblaghie et al., 2003). In

    zebrafish, MKP3 limits the extent of FGF activity of the

    gastrula embryo(Tsang et al., 2004).

    Concluding remarks

    A substantial amount of important work has been done to

    establish the role of FGF signaling during development showing

    that FGFs regulate cell fate and specification of germ layers.

    FGF downstream elements have been characterized and it

    becomes evident that a growing number of these factors arethemselves tightly regulated by FGFs. It is clear from various

    studies that multiple feedback inhibitors exist and that they act at

    different levels of the signal transduction cascades to attenuate

    FGF signal(Fig.4). Regulation can occur at different levels of

    the signal transduction pathways including the membrane down

    to the level of phosphorylation of the MAP kinase. Some of the

    feedback inhibitors are involved in more than one process (Sef,

    Spry) and negatively regulate other RTK signaling pathways.

    The next challenge will be to integrate the intricate

    interactions between these different regulators within the

    FGF pathway and between different RTK pathways. In the

    future, it will be important to assess the pathways employed by

    FGFs in different cells and define the determinants for

    signaling specificity. This can include feedback regulation,

    threshold-type responses and posttranslational modifications.

    This will be complicated by the fact that these modulators

    combine a wide range of activity. The complete understanding

    of these mechanisms will have diverse implications in biology

    as well as in therapeutic approaches.

    Acknowledgment

    We thank the ?FGF communityXfor help and apologize for

    not having being able to cite as many papers as we would have

    wished because of limited space.

    References

    Allen, B.L., Rapraeger, A.C., 2003. Spatial and temporal expression of heparan

    sulfate in mouse development regulated FGF and FGF receptor assembly.

    J. Cell Biol. 163, 637648.

    Alvarez, I.S., Araujo, M., Nieto, M.A., 1998. Neural induction in whole chick

    embryo cultures by FGF. Dev. Biol. 199 (1), 4254.

    Amaya, E., Musci, T.J., Kirschner, M.W., 1991. Expression of a dominant

    negative mutant of the FGF receptor disrupts mesoderm formation in

    Xenopus embryos. Cell 66 (2), 257270.

    Barnett, M.W., Old, R.W., Jones, E.A., 1998. Neural induction and patterning

    by fibroblast growth factor, notochord and somite tissue in Xenopus. Dev.Growth Differ. 40 (1), 4757.

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402 399

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    11/13

    Bertrand, V., et al., 2003. Neural tissue in ascidian embryos is induced by

    FGF9/16/20, acting via a combination of maternal GATA and Ets

    transcription factors. Cell 115 (5), p. 615.

    Bottcher, R.T., et al., 2004. The transmembrane protein XFLRT3 forms a

    complex with FGF receptors and promotes FGF signalling. Nat. Cell Biol.

    1, 3844.

    Brent, A.E., Tabin, C.J., 2004. FGF acts directly on the somitic tendon

    progenitors through the Ets transcription factors Pea3 and Erm to regulatescleraxis expression. Development 131 (16), 3885 3896.

    Burgess, W.H., Maciag, T., 1989. The heparin-binding (fibroblast) growth

    factor family of proteins. Annu. Rev. Biochem. 58, 575606.

    Camps, M., Nichols, A., Arkinstall, S., 2000. Dual specificity phosphatases: a

    gene family for control of MAP kinase function. FASEB J. 14, 6 16.

    Casci, T., Vinos, J., Freeman, M., 1999. Sprouty, an intracellular inhibitor of

    Ras signaling. Cell 96 (5), 655665.

    Chellaiah, A.T., et al., 1994. Fibroblast growth factor receptor (FGFR): 3.

    Alternative splicing in immunoglobulin-like domain III creates a

    receptor highly specific for acidic FGF/FGF-1. J. Biol. Chem. 269 (15),

    11620 11627.

    Chi, C.L., et al., 2003. The isthmic organizer signal FGF8 is required for cell

    survival in the prospective midbrain and cerebellum. Development 130

    (12), 2633 2644.

    Chi, L., et al., 2004. Sprouty proteins regulate ureteric branching bycoordinating reciprocal epithelial Wnt11, mesenchymal Gdnf and stromal

    Fgf7 signalling during kidney development. Development 131 (14),

    33453356.

    Ciruna, B., Rossant, J., 2001. FGF signaling regulates mesoderm cell fate

    specification and morphogenetic movement at the primitive streak. Dev.

    Cell 1 (1), 3749.

    Coffin, J.D., et al., 1995. Abnormal bone growth and selective translational

    regulation in basic fibroblast growth factor (FGF-2) transgenic mice. Mol.

    Biol. Cell 12, 18611873.

    Conlon, F.L., et al., 1996. Inhibition of Xbra transcription activation causes

    defects in mesodermal patterning and reveals autoregulation of Xbra in

    dorsal mesoderm. Development 122 (8), 2427 2435.

    Coumoul, X., Deng, C.X., 2003. Roles of FGF receptors in mammalian

    development and congenital diseases. Birth Defects Res. Part C Embryo

    Today 69 (4), 286304.Cox, W.G., Hemmati-Brivanlou, A., 1995. Caudalization of neural fate by

    tissue recombination and bFGF. Development 121 (12), 4349 4358.

    Crossley, P.H., Martinez, S., Martin, G.R., 1996. Midbrain development

    induced by FGF8 in the chick embryo. Nature 380 (6569), 6668.

    Debiais, F., et al., 2001. Fibroblast growth factor-2 (FGF-2) increases N-

    cadherin expression through protein kinase C and Src-kinase pathways in

    human calvaria osteoblasts. J. Cell. Biochem. 81 (1), 6881.

    Delezoide, A.L., et al., 1998. Spatio-temporal expression of FGFR 1, 2 and 3

    genes during human embryo-fetal ossification. Mech. Dev. 77 (1), 19 30.

    Delany, A.M., Canalis, E., 1998. Dual regulation of stromelysin-3 by

    fibroblast growth factor-2 in murine osteoblasts. J. Biol. Chem. 273 (26),

    1659516600.

    Dhalluin, C., et al., 2000. 1H, 13C and 15N resonance assignments of the SNT

    PTB domain in complex with FGFR1 peptide. J. Biomol. NMR 18 (4),

    371372.Dikic, I., Giordano, S., 2003. Negative receptor signalling. Curr. Opin. Cell

    Biol. 15 (2), 128135.

    Domingos, P.M., et al., 2001. The Wnt/beta-catenin pathway posteriorizes

    neural tissue in Xenopus by an indirect mechanism requiring FGF

    signalling. Dev. Biol. 239 (1), 148160.

    Dudley, A.T., Ros, M.A., Tabin, C.J., 2002. A re-examination of

    proximodistal patterning during vertebrate limb development. Nature

    418 (6897), 539 544.

    Eblaghie, M.C., et al., 2003. Negative feedback regulation of FGF signaling

    levels by Pyst1/MKP3 in chick embryos. Curr. Biol. 13 (12), 10091018.

    Egan, J.E., et al., 2002. The bimodal regulation of epidermal growth factor

    signaling by human Sprouty proteins. Proc. Natl. Acad. Sci. U. S. A. 99 (9),

    60416046.

    Fallon, J.F., et al., 1994. FGF-2: apical ectodermal ridge growth signal for chick

    limb development. Science 264 (5155), 104107.

    Fisher, M.E., Isaacs, H.V., Pownall, M.E., 2002. eFGF is required for activation

    of XmyoD expression in the myogenic cell lineage of Xenopus laevis.

    Development 129 (6), 13071315.

    Fjeld, C., et al., 2000. Mechanistic basis for catalytic activation of mitogen-

    activated protein kinase phosphatase 3 by extracellular signal-regulated

    kinase. J. Biol. Chem. 275, 67496757.

    Flauennhaft, R., Moscatelli, D., Rifkin, D.B., 1990. Heparin and heparan

    sulfate increase the radius of diffusion and action of basic fibroblast growthfactor. J. Cell Biol. 111, 16511659.

    Forman-Kay, J.D., Pawson, T., 1999. Diversity in protein recognition by PTB

    domains. Curr. Opin. Struct. Biol. 9, 690695.

    Furthauer, M., et al., 2001. Sprouty4 acts in vivo as a feedback-in-

    duced antagonist of FGF signaling in zebrafish. Development 128

    (12), 21752186.

    Furthauer, M., et al., 2002. Sef is a feedback-induced antagonist of Ras/MAPK-

    mediated FGF signalling. Nat. Cell Biol. 2, 170 174.

    Furthauer, M., et al., 2004. FGF signalling controls the dorsoventral patterning

    of the zebrafish embryo. Development 131 (12), 2853 2864.

    Gawantka, V., et al., 1998. Gene expression screening in Xenopus identifies

    molecular pathways, predicts gene function and provides a global view of

    embryonic patterning. Mech. Dev. 77 (2), 95141.

    Goldfarb, M., 1996. Functions of fibroblast growth factors in vertebrate

    development. Cytokine Growth Factor Rev. 7 (4), 311 325.Gonzalez, A.M., et al., 1996. Distribution of fibroblast growth factor (FGF)-2

    and FGF receptor-1 messenger RNA expression and protein presence in the

    mid-trimester human fetus. Pediatr. Res. 39 (3), 375385.

    Gotoh, Y., 1995. Involvement of the MAP kinase cascade in Xenopus

    mesoderm induction. EMBO J. 14 (11), 24912498.

    Griffin, K., Patient, R., Holder, N., 1995. Analysis of FGF function in normal

    and no tail zebrafish embryos reveals separate mechanisms for formation of

    the trunk and the tail. Development 121 (9), 29832994.

    Griffin, K.J., et al., 1998. Molecular identification of spadetail: regulation of

    zebrafish trunk and tail mesoderm formation by Tbox genes. Development

    125 (17), 3379 3388.

    Hacohen, N., et al., 1998. Sprouty encodes a novel antagonist of FGF

    signaling that patterns apical branching of the Drosophila airways. Cell

    92 (2), 253263.

    Hall, A.B., et al., 2003. hSpry2 is targeted to the ubiquitin-dependentproteasome pathway by c-Cbl. Curr. Biol. 13 (4), 308 314.

    Halpern, M.E., et al., 1993. Induction of muscle pioneers and floor plate is

    distinguished by the zebrafish no tail mutation. Cell 75 (1), 99 111.

    Hanafusa, H., et al., 2002. Sprouty1 and Sprouty2 provide a control

    mechanism for the Ras/MAPK signalling pathway. Nat. Cell Biol. 11,

    850858.

    Hemmati-Brivanlou, A., Melton, D., 1997. Vertebrate neural induction. Annu.

    Rev. Neurosci. 20, 43 60.

    Herrmann, B.G., et al., 1990. Cloning of the T gene required in mesoderm

    formation in the mouse. Nature 343 (6259), 617622.

    Holowacz, T., Sokol, S., 1999. FGF is required for posterior neural patterning

    but not for neural induction. Dev. Biol. 205 (2), 296 308.

    Hongo, I., Kengaku, M., Okamoto, H., 1999. FGF signaling and the anterior

    neural induction in Xenopus. Dev. Biol. 216 (2), 561581.

    Isaacs, H.V., Pownall, M.E., Slack, J.M., 1994. eFGF regulates Xbra expressionduring Xenopus gastrulation. EMBO J. 13 (19), 44694481.

    Itoh, N., Ornitz, D.M., 2004. Evolution of the Fgf and Fgfr gene families.

    Trends Genet. 20 (n-11), 563569.

    Johnson, D.E., Williams, L.T., 1993. Structural and functional diversity in the

    FGF receptor multigene family. Adv. Cancer Res. 60, 141.

    Johnson, D.E., et al., 1990. Diverse forms of a receptor for acidic and basic

    fibroblast growth factors. Mol. Cell. Biol. (9), 47284736.

    Kawaguchi, H., et al., 1995. Transcriptional induction of prostaglandin

    G/H synthase-2 by basic fibroblast growth factor. J. Clin. Invest. 96 (2),

    923930.

    Kawakami, Y., et al., 2003. MKP3 mediates the cellular response to FGF8

    signalling in the vertebrate limb. Nat. Cell Biol. 5, 513519.

    Kazanskaya, O., Glinka, A., Niehrs, C., 2000. The role ofXenopus dickkopf1

    in prechordal plate specification and neural patterning. Development 127

    (22), 48924981.

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402400

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    12/13

    Kengaku, M., Okamoto, H., 1995. bFGF as a possible morphogen for the

    anteroposterior axis of the central nervous system in Xenopus. Develop-

    ment 121 (9), 31213130.

    Kim, H.J., Bar-Sagi, D., 2004. Modulation of signalling by Sprouty: a

    developing story. Nat. Rev., Mol. Cell Biol. 5 (6), 441450.

    Kim, H.J., et al., 1998. FGF-, BMP- and Shh-mediated signalling pathways in

    the regulation of cranial suture morphogenesis and calvarial bone

    development. Development 125 (7), 1241 1251.Kimelman, D., Kirschner, M., 1987. Synergistic induction of mesoderm by

    FGF and TGF-beta and the identification of an mRNA coding for FGF in

    the earlyXenopus embryo. Cell 51 (5), 869877.

    Kimelman, D., et al., 1988. The presence of fibroblast growth factor in the

    frog egg: its role as a natural mesoderm inducer. Science 242 (4881),

    10531056.

    Kovalenko, D., et al., 2003. Sef inhibits fibroblast growth factor signaling by

    inhibiting FGFR1 tyrosine phosphorylation and subsequent ERK activa-

    tion. J. Biol. Chem. 278 (16), 1408714091.

    Kouhara, H., et al., 1997. A lipid-anchored Grb2-binding protein that links

    FGF-receptor activation to the Ras/MAPK signaling pathway. Cell 89 (5),

    693702.

    Kroll, K.L., Amaya, E., 1996. Transgenic Xenopus embryos from sperm

    nuclear transplantations reveal FGF signaling requirements during gastru-

    lation. Development 122 (10), 31733183.Kubota, Y., Itoh, K., 2000. Chemotactic migration of mesencephalic neural

    crest cells in the mouse. Dev. Dyn. 217, 170179.

    Kudoh, T., et al., 2001. A gene expression screen in zebrafish embryogenesis.

    Genome Res. 12, 1979 1987.

    Kudoh, T., Wilson, S.W., Dawid, I.B., 2002. Distinct roles for Fgf, Wnt and

    retinoic acid in posteriorizing the neural ectoderm. Development 129 (18),

    43354346.

    LaBonne, C., 1995. Role of MAP kinase in mesoderm induction and

    axial patterning during Xenopus development. Development 121 (5),

    14751486.

    Lamb, T.M., Harland, R.M., 1995. Fibroblast growth factor is a direct neural

    inducer, which combined with noggin generates anteriorposterior neural

    pattern. Development 121 (11), 3627 3636.

    Laufer, E., et al., 1994. Sonic hedgehog and FGF-4 act through a signaling

    cascade and feedback loop to integrate growth and patterning of thedeveloping limb bud. Cell 79 (6), 993 1003.

    Launay, C., et al., 1996. A truncated FGF receptor blocks neural induction by

    endogenous Xenopus inducers. Development 122 (3), 869880.

    Lee, P.L., et al., 1989. Purification and complementary DNA cloning of a

    receptor for basic fibroblast growth factor. Science 245 (4913), 57 60.

    Lewandoski, M., Sun, X., Martin, G.R., 2000. Fgf8 signalling from the AER is

    essential for normal limb development. Nat. Genet. 26 (4), 460 463.

    Li, S., Muneoka, K., 1999. Cell migration and chick limb development

    chemotactic action of FGF4 and the AER. Dev. Biol. 211, 335 347.

    Lin, X., 2004. Functions of heparan sulfate proteoglycans in cell signaling

    during development. Development 131 (24), 6009 6021.

    Lin, X., et al., 1999. Heparan sulfate proteoglycans are essential for FGF

    receptor signaling during Drosophila embryonic development. Develop-

    ment 126 (17), 37153723.

    Linker, C., Stern, C., 2004. Neural induction requires BMP inhibition only as alate step, and involves signals other than FGF and Wnt antagonists.

    Development 131 (22), 5671 5681.

    Liu, A., Joyner, A.L., 2001. Early anterior/posterior patterning of the midbrain

    and cerebellum. Ann. Rev. Neurosci. 24, 869896.

    Liu, Z., et al., 2002. Coordination of chondrogenesis and osteogenesis by

    fibroblast growth factor 18. Genes Dev. 16 (7), 859869.

    Liu, A., et al., 2003. FGF17b and FGF18 have different midbrain regulatory

    properties from FGF8b or activated FGF receptors. Development 130 (25),

    61756185.

    MacNicol, A.M., et al., 1993. Raf-1 kinase is essential for early Xenopus

    development and mediates the induction of mesoderm by FGF. Cell 73 (3),

    571583.

    Mailleux, A.A., et al., 2001. Evidence that SPROUTY2 functions as an

    inhibitor of mouse embryonic lung growth and morphogenesis. Mech. Dev.

    102 (12), 8194.

    Mansukhani, A., et al., 2000. Signaling by fibroblast growth factors (FGF) and

    fibroblast growth factor receptor 2 (FGFR2)-activating mutations blocks

    mineralization and induces apoptosis in osteoblasts. J. Cell Biol. 149 (6),

    12971308.

    Marie, P.J., 2003. Fibroblast growth factor signaling controlling osteoblast

    differentiation. Gene 316, 2332.

    Martin, G.R., 1998. The roles of FGFs in the early development of vertebrate

    limbs. Genes Dev. 12 (11), 15711586.Maruoka, Y., et al., 1998. Comparison of the expression of three highly related

    genes, FGF8, FGF17 and FGF18, in the mouse embryo. Mech. Dev. 74

    (12), 175177.

    McGrew, L.L., Hoppler, S., Moon, R.T., 1997. Wnt and FGF pathways

    cooperatively pattern anteroposterior neural ectoderm in Xenopus. Mech.

    Dev. 69 (12), 105114.

    Meyers, E.N., Lewandoski, M., Martin, G.R., 1998. An FGF8 mutant allelic

    series generated by Cre- and Flp-mediated recombination. Nat. Genet. 2,

    136141.

    Miki, T., et al., 1992. Determination of ligand-binding specificity by alternative

    splicing: two distinct growth factor receptors encoded by a single gene.

    Proc. Natl. Acad. Sci. U. S. A. 89 (1), 246250.

    Minowada, G., et al., 1999. Vertebrate Sprouty genes are induced by FGF

    signaling and can cause chondrodysplasia when overexpressed. Develop-

    ment 126 (20), 4465 4475.Molteni, A., et al., 1999. Differential expression of fibroblast growth factor

    receptor-1, -2, and -3 and syndecan-1, -2, and -4 in neonatal rat mandibular

    condyle and calvaria during osteogenic differentiation in vitro. Bone 4,

    337347.

    Montero, A., et al., 2000. Disruption of the fibroblast growth factor-2 gene

    results in decreased bone mass and bone formation. J. Clin. Invest. 105 (8),

    10851093.

    Moon, A.M., Capecchi, M.R., 2000. Fgf8 is required for outgrowth and

    patterning of the limbs. Nat. Genet. 26 (4), 455 459.

    Moscatelli, D., 1987. 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 (1), 123 130.

    Munchberg, S.R., Ober, E.A., Steinbeisser, H., 1999. Expression of the Ets

    transcription factors erm and pea3 in early zebrafish development. Mech.Dev. 88 (2), 233236.

    Nakato, H., Kimata, K., 2002. Heparan sulfate fine structure and

    specificity of proteoglycan functions. Biochim. Biophys. Acta 21573

    (3), 312318.

    Niehrs, C., Meinhardt, H., 2002. Modular feedback. Nature 417 (6884), 35 36.

    Niswander, L., 2003. Pattern formation: old models out on a limb. Nat. Rev. 4,

    133143.

    Nutt, S.L., et al., 2001.Xenopus Sprouty2 inhibits FGF-mediated gastrulation

    movements but does not affect mesoderm induction and patterning. Genes

    Dev. 15 (9), 11521166.

    Nybakken, K., Perrimon, N., 2002. Heparan sulfate proteoglycan modulation of

    developmental signaling in Drosophila. Biochim. Biophys. Acta 1573 (3),

    280291.

    Ohbayashi, N., et al., 2002. FGF18 is required for normal cell proliferation and

    differentiation during osteogenesis and chondrogenesis. Genes Dev. 16 (7),870879.

    Ornitz, D.M., 2000. FGFs, heparan sulfate and FGFRs: complex interactions

    essential for development. BioEssays 2, 108112.

    Ornitz, D.M., Itoh, N., 2001. Fibroblast growth factors. Genome Biol. 2 (3),

    112.

    Orr-Urtreger, A., et al., 1991. Developmental expression of two murine

    fibroblast growth factor receptors, flg and bek. Development 113 (4),

    14191434.

    Park, W.Y., Miranda, B., Lebeche, D., Hashimoto, G., Cardoso, W.V., 1998.

    FGF10 is a chemotactic factor for distal epithelial buds during lung

    development. Dev. Biol. 201, 125134.

    Pawson, T., 1995. Protein modules and signalling networks. Nature 373 (6515),

    573580.

    Pellegrini, L., 2001. Role of heparan sulfate in fibroblast growth factor

    signalling: a structural view. Curr. Opin. Struct. Biol. 11 (5), 629634.

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402 401

  • 8/10/2019 Functions and Regulations of Fibroblast Growth Factor Signaling During

    13/13

    Pellegrini, L., et al., 2000. Crystal structure of fibroblast growth factor receptor

    ectodomain bound to ligand and heparin. Nature 407 (6807), 10291034.

    Powers, C.J., McLeskey, S.W., Wellstein, A., 2000. Fibroblast growth factors,

    their receptors and signaling. Endocr. Relat. Cancer 3, 165 197.

    Preger, E., 2004. Alternative splicing generates an isoform of the human Sef

    gene with altered subcellular localization and specificity. Proc. Natl. Acad.

    Sci. U. S. A. 101 (5), 12291234.

    Reich, A., Sapir, A., Shilo, B., 1999. Sprouty is a general inhibitor of receptortyrosine kinase signaling. Development 126 (18), 4139 4147.

    Reifers, F., 1998. FGF8 is mutated in zebrafish acerebellar (ace) mutants and is

    required for maintenance of midbrain hindbrain boundary development

    and somitogenesis. Development 125 (13), 2381 2395.

    Ribisi Jr., S., et al., 2000. Ras-mediated FGF signaling is required for the

    formation of posterior but not anterior neural tissue in Xenopus laevis. Dev.

    Biol. 227 (1), 183196.

    Rousseau, F., et al., 1994. Mutations in the gene encoding fibroblast growth

    factor receptor-3 in achondroplasia. Nature 371 (6494), 252 254.

    Rubin, C., Litvak, V., Medvedovsky, H., Zwang, Y., Lev, S., Yarden, Y., 2003.

    Sprouty fine-tunes EGF signaling through interlinked positive and negative

    feedback loops. Curr. Biol. 13 (4), 297307.

    Sasaki, A., et al., 2003. Mammalian Sprouty4 suppresses Ras-independent

    ERK activation by binding to Raf1. Nat. Cell Biol. 5, 427432.

    Sato, T., Araki, I., Nakamura, H., 2001. Inductive signal and tissueresponsiveness defining the tectum and the cerebellum. Development 128,

    24612469.

    Sharrocks, A.D., 2001. The ETS-domain transcription factor family. Nat. Rev.,

    Mol. Cell Biol. 2 (11), 827837.

    Schlessinger, J., et al., 2000. Crystal structure of a ternary FGFFGFR

    heparin complex reveals a dual role for heparin in FGFR binding and

    dimerization. Mol. Cell 3, 743 750.

    Schmid, B., et al., 2000. Equivalent genetic roles for bmp7/snailhouse and

    bmp2b/swirl in dorsoventral pattern formation. Development 127 (5),

    957967.

    Schulte-Merker, S., Smith, J.C., 1995. Mesoderm formation in response to

    Brachyury requires FGF signalling. Curr. Biol. 5 (1), 6267.

    Sheng, G., Dos Reis, M., Stern, C.D., 2003. Churchill, a zinc finger

    transcriptional activator, regulates the transition between gastrulation and

    neurulation. Cell 115 (5), 603613.Shiang, R., et al., 1994. Mutations in the transmembrane domain of FGFR3

    cause the most common genetic form of dwarfism, achondroplasia. Cell 78

    (2), 335342.

    Slack, J.M., et al., 1987. Mesoderm induction in early Xenopus embryos by

    heparin-binding growth factors. Nature 326 (6109), 197 200.

    Smith, J.C., et al., 1991. Expression of a Xenopus homolog of Brachyury

    (T) is an immediate-early response to mesoderm induction. Cell 67 (1),

    7987.

    Stanley, H.J., Zorn, A.M., Gurdon, J.B., 2001. eFGF and its mode of action in

    the community effect during Xenopus myogenesis. Development 128 (8),

    13471357.

    Sternberg, P.W., Alberola-Ila, J., 1998. Conspiracy theory: RAS and RAF do

    not act alone. Cell 95 (4), 447450.

    Stewart, A.E., et al., 1999. Crystal structure of the MAPK phosphatase Pyst1

    catalytic domain and implications for regulated activation. Nat. Struct. Biol.6, 174181.

    Storey, K.G., et al., 1998. Early posterior neural tissue is induced by FGF in the

    chick embryo. Development 125 (3), 473484.

    Streit, A., et al., 2000. Initiation of neural induction by FGF signalling before

    gastrulation. Nature 406 (6791), 7478.

    Strong, C.F., et al., 2000. Xbra3 induces mesoderm and neural tissue in

    Xenopus laevis . Dev. Biol. 222 (2), 405419.

    Summerbell, D., Lewis, J.H., Wolpert, L., 1973. Positional information in chick

    limb morphogenesis. Nature 244 (5417), 492496.

    Sun, X., et al., 1999. Targeted disruption of Fgf8 causes failure of cell migration

    in the gastrulating mouse embryo. Genes Dev. 13 (14), 1834 1846.

    Sun, X., Mariani, F.V., Martin, G., 2002. Functions of FGF signaling

    from the apical ectodermal ridge in limb development. Nature 418,

    501508.

    Thisse, B., et al., 2001, Expression of the zebrafish genome during embryo-

    genesis, ZFIN. http://zfin.org/cgibin/webdriver?MIval=aapubview2.apg_

    and_OID=ZDB-PUB-010810-1.

    Torii, S., et al., 2004. Sef is a spatial regulator for Ras/MAP kinase signaling.

    Dev. Cell 7 (1), 3344.Tsang, M., Dawid, I.B., 2004. Promotion and attenuation of FGF signaling

    through the Ras-MAPK pathway. Sci. STKE 228, e17.

    Tsang, M., et al., 2002. Identification of Sef, a novel modulator of FGF

    signalling. Nat. Cell Biol. 4 (2), 165169.

    Tsang, M., et al., 2004. A role for MKP3 in axial patterning of the zebrafish

    embryo. Development 131 (12), 27692779.

    Ueda, Y., et al., 1996. Protein kinase C activates the MEK-ERK pathway in

    a manner independent of Ras and dependent on Raf. J. Biol. Chem. 271

    (38), 2351223519.

    Umbhauer, M., et al., 1995. Mesoderm induction in Xenopus caused by

    activation of MAP kinase. Nature 376 (6535), 5862.

    Umbhauer, M., et al., 2000. Signaling specificities of fibroblast growth factor

    receptors in early Xenopus embryo. J. Cell Sci. 113 (16), 2865 2875.

    Varghese, S., et al., 1995. Basic fibroblast growth factor stimulates expression

    of interstitial collagenase and inhibitors of metalloproteinases in rat bonecells. Endocrinology 136 (5), 2156 2162.

    Warren, S.M., et al., 2003. The BMP antagonist noggin regulates cranial suture

    fusion. Nature 422 (6932), 625629.

    Wasylyk, B., Hagman, J., Gutierrez-Hartmann, A., 1998. Ets transcription

    factors: nuclear effectors of the Ras-MAP-kinase signaling pathway. Trends

    Biochem. Sci. 6, 213216.

    Webb, S.E., et al., 1997. Fibroblast growth factors 2 and 4 stimulate

    migration of mouse embryonic limb myogenic cells. Dev. Dyn. 209,

    206216.

    Weinstein, D.C., Hemmati-Brivanlou, A., 1999. Neural induction. Annu. Rev.

    Cell Dev. Biol. 15, 411433.

    Whitman, M., Melton, D.A., 1992. Involvement of p21ras in Xenopus

    mesoderm induction. Nature 357 (6375), 252 254.

    Wilson, S.I., et al., 2000. An early requirement for FGF signalling in the

    acquisition of neural cell fate in the chick embryo. Curr. Biol. 10 (8),421429.

    Wong, E.S., et al., 2002. Sprouty2 attenuates epidermal growth factor receptor

    ubiquitylation and endocytosis, and consequently enhances Ras/ERK

    signalling. EMBO J. 21 (18), 4796 4808.

    Xiong, S., et al., 2003. hSef inhibits PC-12 cell differentiation by interfering

    with Ras-mitogen-activated protein kinase MAPK signaling. J. Biol. Chem.

    278 (50), 5027350282.

    Xu, X., et al., 1998. Fibroblast growth factor receptor 2 (FGFR2)-mediated

    reciprocal regulation loop between FGF8 and FGF10 is essential for limb

    induction. Development 125 (4), 753 765.

    Xu, J., Liu, Z., Ornitz, D.M., 2000. Temporal and spatial gradients of FGF8 and

    FGF17 regulate proliferation and differentiation of midline cerebellar

    structures. Development 9, 18331843.

    Yang, X., et al., 2002. Cell movement patterns during gastrulation in the chick

    are controlled by positive and negative chemotaxis mediated by FGF4 andFGF8. Dev. Cell 3 (3), 425437.

    Yusoff, P., et al., 2002. Sprouty2 inhibits the Ras/MAP kinase pathway by

    inhibiting the activation of Raf. J. Biol. Chem. 277 (5), 3195 3201.

    Zhang, J., et al., 2003. A bipartite mechanism for ERK2 recognition by its

    cognate regulators and substrates. J. Biol. Chem. 278, 29901 29912.

    Zhao, Y., Zhang, Z.Y., 2001. The mechanism of dephosphorylation of

    extracellular signal-regulated kinase 2 by mitogen-activated protein kinase

    phosphatase 3. J. Biol. Chem. 276, 32382 32391.

    Zhao, J., Cao, Y., Zhao, C., Postlethwait, J., Meng, A., 2003. An SP1-like

    transcription factor Spr2 acts downstream of Fgf signaling to mediate

    mesoderm induction. EMBO J. 22 (22), 60786088.

    B. Thisse, C. Thisse / Developmental Biology 287 (2005) 390 402402

    http://%20http//www.zfin.org/cgibin/webdriver?MIval=aapubview2.apg_and_OID=ZDB-PUB-010810-1http://%20http//www.zfin.org/cgibin/webdriver?MIval=aapubview2.apg_and_OID=ZDB-PUB-010810-1http://%20http//www.zfin.org/cgibin/webdriver?MIval=aapubview2.apg_and_OID=ZDB-PUB-010810-1