当前位置: 首页 > 期刊 > 《内分泌检查杂志》 > 2005年第1期 > 正文
编号:11372796
Fibroblast Growth Factor Signaling during Early Vertebrate Development
http://www.100md.com 《内分泌检查杂志》
     Division of Molecular Embryology, Deutsches Krebsforschungszentrum, D-69120 Heidelberg, Germany

    Correspondence: Address all correspondence and requests for reprints to: Ralph T. B?ttcher, Department of Molecular, Cellular and Developmental Biology, Yale University, Kline Biology Tower-244, 266 Whitney Avenue, New Haven, Connecticut 06511. E-mail: ralph.boettcher@yale.edu

    Abstract

    Fibroblast growth factors (FGFs) have been implicated in diverse cellular processes including apoptosis, cell survival, chemotaxis, cell adhesion, migration, differentiation, and proliferation. This review presents our current understanding on the roles of FGF signaling, the pathways employed, and its regulation. We focus on FGF signaling during early embryonic processes in vertebrates, such as induction and patterning of the three germ layers as well as its function in the control of morphogenetic movements.

    I. Introduction

    II. FGFs and FGF Receptors (FGFRs)

    III. FGFR Signal Transduction

    A. Ras/MAPK pathway

    B. PLC/Ca2+ pathway

    C. PI3 kinase/Akt pathway

    IV. The Biological Activities of FGFs

    A. FGF signaling pathway in mesoderm formation and gastrulation movements

    B. Chemotactic action of FGFs in morphogenetic movements

    C. FGF signaling in neural induction and AP patterning

    D. FGF signaling in endoderm formation

    V. Regulation of FGF Signaling

    A. Regulation by HSPGs

    B. Transmembrane regulators

    C. Intracellular regulators

    VI. Concluding Remarks

    I. Introduction

    FIBROBLAST GROWTH FACTORS (FGFs) constitute a large family of polypeptide growth factors found in a variety of multicellular organisms, including invertebrates. The first member of this family, FGF2, was identified by its ability to stimulate proliferation of mouse 3T3 fibroblasts (1); however, the function of FGFs is not restricted to cell growth. Instead, FGFs are involved in diverse cellular processes including chemotaxis, cell migration, differentiation, cell survival, and apoptosis. The common feature of FGFs is that they are structurally related and generally signal through receptor tyrosine kinases. FGFs play an important role in embryonic development in invertebrates and vertebrates. This review focuses on the role of FGF signaling during early vertebrate development. Excellent reviews dealing with the role of FGFs in Drosophila (2, 3, 4), Caenorhabditis elegans (5, 6), vertebrate limb development (7, 8, 9), central nervous system development (10, 11), skeleton formation (12), and cancer (13, 14, 15) are available.

    II. FGFs and FGF Receptors (FGFRs)

    The human FGF protein family consists of 22 members that share a high affinity for heparin as well as a high-sequence homology within a central core domain of 120 amino acids (16), which interacts with the FGFR (17, 18, 19) (Fig. 1A). FGFs can be classified into subgroups according to structure, biochemical properties, and expression (for reviews see Refs.16 and 20). For example, members of the FGF8 subfamily (FGF8, FGF17, FGF18) share 70–80% amino acid sequence identity and similar receptor-binding properties and have overlapping expression patterns. FGFs appear to be diffusible (21), and they are able to act in a dose-dependent fashion, e.g. to induce different marker genes at different concentrations (22, 23). This raises the question of whether FGFs act physiologically as morphogens. Indeed, in neural progenitors, graded FGF signaling appears to be translated into a distinct motor neuron Hox pattern (24).

    FIG. 1. Domain structure of generic FGF and FGFR proteins. A, Structure of a generic FGF protein containing a signal sequence and the conserved core region that contains receptor- and HSPG-binding sites. B, The main structural features of FGFRs including Ig domains, acidic box, heparin-binding domain, CAM-homology domain (CHD), transmembrane domain, and a split tyrosine kinase domain are illustrated with respective functions. CAM, Cell adhesion molecule; ECM, extracellular matrix; PKC, protein kinase C.

    FGFs induce their biological responses by binding to and activating FGFRs, a subfamily of cell surface receptor tyrosine kinases (RTKs). The vertebrate Fgfr gene family consists of four highly related genes, Fgfr1–4 (25). These genes code for single spanning transmembrane proteins with an extracellular ligand-binding region and an intracellular domain harboring tyrosine kinase activity. The extracellular region is composed of Ig-like domains that are required for FGF binding. The Ig-like domains also regulate binding affinity and ligand specificity (Fig. 1B). Located between Ig-like domains I and II is a stretch of acidic amino acids (acidic box domain) followed by a heparin-binding region and a cell adhesion homology domain. These domains are required for the interaction of the receptor with components of the extracellular matrix, in particular heparan sulfate (HS) proteoglycans (HSPGs), and cell adhesion molecules (CAMs). The intracellular part of the receptor includes the juxtamembrane domain, the split tyrosine kinase domain, and a short carboxy-terminal tail. In addition to the enzymatic activity, the intracellular domain harbors protein binding and phosphorylation sites (including protein kinase C and FRS2 sites) as well as several autophosphorylation sites that interact with intracellular substrates.

    Alternative splicing of Fgfr transcripts generates a diversity of FGFR isoforms (25, 26). In vitro assays show that different isoforms have distinct FGF-binding specificities (27, 28). The tissue-specific alternative splicing encompassing the carboxy-terminal half of Ig domain III strongly affects ligand-receptor binding specificity (29, 30, 31). In vivo, FGFR-FGF complex assembly and signaling is further regulated by the spatial and temporal expression of endogenous HSPGs (32, 33). HSPGs are required to activate effectively the FGFR and determine the interaction between specific FGF-FGFR pairs (see Section V.A.).

    III. FGFR Signal Transduction

    FGFRs, like other receptor tyrosine kinases, transmit extracellular signals after ligand binding to various cytoplasmic signal transduction pathways through tyrosine phosphorylation. Binding of FGFs causes receptor dimerization and triggers tyrosine kinase activation leading to autophosphorylation of the intracellular domain (34). Tyrosine autophosphorylation controls the protein tyrosine kinase activity of the receptor but also serves as a mechanism for assembly and recruitment of signaling complexes (35). These phosphorylated tyrosines function as binding sites for Src homology 2 and phosphotyrosine binding domains of signaling proteins, resulting in their phosphorylation and activation (36, 37). A subset of Src homology 2-containing proteins such as Src-kinase and phospholipase C (PLC) possesses intrinsic catalytic activities, whereas others are adapter proteins. FGF signal transduction, as analyzed during early embryonic development, can proceed via three main pathways described below (Fig. 2).

    FIG. 2. Intracellular signaling pathways activated through FGFRs. For simplicity, only those proteins are indicated that are discussed in the main text. Formation of a ternary FGF-heparin-FGFR complex leads to receptor autophosphorylation and activation of intracellular signaling cascades, including the Ras/MAPK pathway (shown in blue), PI3 kinase/Akt pathway (shown in green), and the PLC/Ca2+ pathway (shown in yellow). Proteins connected with two pathways are striped. Members of the FGF synexpression group are illustrated in red. The Ras/MAPK cascade is activated by binding of Grb2 to phosphorylated FRS2. The subsequent formation of a Grb2/SOS complex leads to the activation of Ras. Three routes by which FGFRs can activate PI3 kinase/Akt pathway are indicated. First, Gab1 can bind to FRS2 indirectly via Grb2, resulting in tyrosine phosphorylation and activation of the PI3-kinase/Akt pathway via p85 (41 201 ). Second, the PI3 kinase-regulatory subunit p85 can bind to a phosphorylated tyrosine residue of the FGFR, as shown in Xenopus cell extracts (202 ). Alternatively, activated Ras can induce membrane localization and activation of the p110 catalytic subunit of PI3 kinase (203 204 ). AA, Arachidonic acid; DAG, diacylglycerol; IP3, inositol-1,4,5-triphosphate.

    A. Ras/MAPK pathway

    The most common pathway employed by FGFs is the MAPK pathway. This involves the lipid-anchored docking protein FRS2 (also called SNT1) that constitutively binds FGFR1 even without receptor activation (38, 39, 40). Several groups have demonstrated the importance of FRS2 in FGFR1-mediated signal transduction during embryonic development (41, 42, 43). After activation of the FGFR, tyrosine-phosphorylated FRS2 functions as a site for coordinated assembly of a multiprotein complex activating and controlling the Ras-MAPK signaling cascade and the phosphatidylinositol 3 (PI3)-kinase/Akt pathway. The FRS2 tyrosine phosphorylation sites are recognized and bound by the adapter protein Grb2 and the protein tyrosine phosphatase (PTP) Shp2 (39, 44). Grb2 forms a complex with the guanine nucleotide exchange factor Son of sevenless (SOS) via its SH3 domain. Translocation of this complex to the plasma membrane by binding to phosphorylated FRS2 allows SOS to activate Ras by GTP exchange due to its close proximity to membrane-bound Ras. Once in the active GTP-bound state, Ras interacts with several effector proteins, including Raf leading to the activation of the MAPK signaling cascade. This cascade leads to phosphorylation of target transcription factors, such as c-myc, AP1, and members of the Ets family of transcription factors (reviewed in Ref.45).

    B. PLC/Ca2+ pathway

    The PLC/Ca2+ pathway involves binding of PLC to phosphorylated tyrosine 766 of FGFR1 (46, 47). Upon activation, PLC hydrolyzes phosphatidylinositol-4,5-diphosphate to form two second messengers, inositol-1,4,5-triphosphate and diacylglycerol. Diacylglycerol is an activator of protein kinase C, whereas inositol-1,4,5-triphosphate stimulates the release of intracellular Ca2+. This cascade has been implicated in the FGF2-stimulated neurite outgrowth (48, 49) and in the caudalization of neural tissue by FGFR4 in Xenopus (50).

    C. PI3 kinase/Akt pathway

    The PI3 kinase/Akt pathway can be activated by three mechanisms after FGFR activation (Fig. 2), and the phospholipids thereby generated regulate directly or indirectly the activity of target proteins such as Akt/PKB.

    Among other processes, the PI3 kinase signaling branch is involved during Xenopus mesoderm induction acting in parallel to the Ras/MAPK pathway. Overexpression of a dominant negative form of the PI3 kinase-regulatory subunit p85 interferes with Xenopus mesoderm formation. Conversely, coexpression of activated forms of MAPK and PI3 kinase leads to synergistic mesoderm induction (51).

    IV. The Biological Activities of FGFs

    Early developmental processes during gastrulation stages of Xenopus, zebrafish, chicken, and mouse include mesoderm formation and gastrulation movements, neural induction and AP patterning, and endoderm formation (Fig. 3). FGF signaling plays important roles in early vertebrate development during these processes as discussed below and summarized in Table 1. Genes that function in FGF signaling and for which loss-of-function analysis revealed an early embryonic phenotype are summarized in Table 2.

    FIG. 3. Early developmental events during Xenopus, zebrafish, chick, and mouse gastrulation. A, In Xenopus, the three germ layers form along the animal-vegetal axis: the pigmented animal pole becomes ectoderm, and the yolk-rich vegetal pole becomes endoderm. Mesoderm is induced in an equatorial region (marginal zone) between the two other layers by signals emitted from the underlying prospective endoderm (black arrows). The dorsal mesoderm, the Spemann organizer, emits neural inducing signals (blue arrow). In the anamniotes, Xenopus and zebrafish (A and B), gastrulation is triggered by the involution of prospective mesendodermal cells starting at the organizer region of the embryos (red arrows). Convergence of cells toward the embryonic midline and anterior-posterior elongation of the body axis is achieved by cellular rearrangements (convergent extension). B, In zebrafish, the yolk syncitial layer (YSL) emits meso- and endoderm-inducing signals to the blastoderm (black arrows). Close to the margin, mesodermal and endodermal progenitors are interspersed, whereas further away from the margin only mesodermal cells are present. The zebrafish organizer, the shield, emits neural inducing signals acting within the plane of the epiblast (blue arrow). C, The chick embryo is a bilayered structure composed of the epiblast and the extraembryonic hypoblast. . Gastrulation starts with the formation of the primitive streak with Hensen’s node (the avian organizer, Hn) forming at its tip. During gastrulation, epiblast cells undergo epithelial to mesenchymal transition and ingress as individual cells through the primitive streak (red arrows). In the space between the epiblast and the hypoblast the migrating cells form axial and lateral mesoderm as well as definitive endoderm. Ectodermal cells become specified as either neural or epidermal cells before the onset of gastrulation. D, The mouse gastrula embryo has a cylindrical shape and consists of an outer and inner epidermal layer (for simplicity, only the inner layer is shown) As gastrulation starts, epiblast cells at the posterior side undergo an epithelial-mesenchymal transformation to generate mesoderm precursors and form the primitive streak. The streak elongates during gastrulation while mesodermal and endodermal progenitors start migrating through the primitive streak (red arrows).

    TABLE 1. Processes involving FGF signaling in early vertebrate embryos

    TABLE 2. Early developmental defects caused by loss of function (LOF) of components of the FGF signaling pathway

    A. FGF signaling pathway in mesoderm formation and gastrulation movements

    Induction and patterning of mesoderm is one of the earliest events during vertebrate body axis formation. Although TGF-? signaling plays a pivotal role in mesoderm induction and patterning and has received most attention in this context recently, FGF signaling is also essential. Indeed, the first identified mesoderm inducer was basic FGF (52), and this represents an evolutionarily conserved role, because FGF signaling is required for mesoderm induction in the primitive chordate Ciona (53). In Xenopus, mouse, and zebrafish, disruption of FGF signaling by Fgfr knock out or by overexpression of a dominant negative FGFR1 strongly affects body axis formation (54, 55, 56, 57). In these embryos, phenotypic changes are observed mostly in posterior regions, such as defects of trunk and tail structures. Despite some organism-specific differences, two functions for FGFs are conserved. First, FGFs control the specification and maintenance of mesoderm by regulation of T box transcription factors in Xenopus (58, 59), mouse (60, 61), and zebrafish (57, 62, 63). Second, they have a direct and/or indirect role in morphogenetic movements during gastrulation.

    The effect of FGFs on mesoderm formation was first demonstrated and most extensively studied in Xenopus (64). Since the first studies that discovered the mesoderm-inducing ability of FGFs (52, 65, 66), many components of the FGF signaling pathway were shown in Xenopus to be required for mesoderm development. These components include FGFRs (54, 67), HSPGs (68), adaptor proteins FRS2/SNT-1, Nck, and Grb2 (42, 43, 69), Src-kinase Laloo (70), PTPs (71, 72), Ras, Raf, MAPK kinase, ERK (73, 74, 75, 76), and the transcription factors AP1 and Ets2 (77, 78). Their inhibition blocks mesoderm formation and induces posterior and gastrulation defects. Conversely, in gain-of-function experiments they induce mesodermal markers and phenocopy FGF overexpression. The role of FGF signaling during mesoderm formation is that of a competence factor. It is required for cells to respond to TGF-?-like mesoderm inducers, because activin-mediated mesoderm induction in Xenopus animal caps, for example, is blocked by a dominant negative FGFR1 (79, 80, 81). Similarly, the trunk-inducing activity of dorsal mesoderm, which depends on Nodal signaling, is inhibited by a dominant negative FGFR1 (82). The ability of FGFs to regulate Nodal signaling may be due to phosphorylation of Smad2 by MAPKs (83).

    FGF signaling needs to continue during gastrulation for mesoderm maintenance as shown by transgenic Xenopus embryos overexpressing a dominant negative FGFR1 (84). One of the FGF mesodermal targets is the T box transcription factor brachyury (Xbra), required for posterior mesoderm and axis formations in mouse, zebrafish, and Xenopus (58, 85, 86, 87). Xbra and eFGF form an autoregulatory loop in which Xbra induces eFGF and vice versa (88, 89). eFGF is also required for XmyoD expression in the myogenic cell lineage of Xenopus. eFGF is a candidate factor mediating the community effect, whereby cells differentiate into muscle once a critical cell mass is reached, due to the accumulation of diffusible factors (90, 91).

    In Xenopus embryos, FGF signaling has both direct and indirect roles on gastrulation movements. Nutt et al. (92) provided evidence for a direct role of FGF signaling in cell movements during convergence extension. They identified and characterized XSprouty2 as an intracellular antagonist of FGF-dependent calcium signaling. Overexpression of XSprouty2 inhibits convergent extension movements without affecting MAPK activity, mesoderm induction, and patterning. FGF signaling also affects gastrulation movements indirectly via regulation of Xbra induction and maintenance. Xbra acts as a switch promoting convergent extension and actively repressing cell migration by inhibiting adhesion of migrating cells to fibronectin (93, 94). Also, Xbra directly activates the expression of Xwnt11 and prickle (95, 96), which regulate convergent extension movements via the noncanonical Wnt pathway (95, 96, 97).

    In zebrafish, as in Xenopus, FGF signaling acts as competence factor to control mesoderm induction by TGF-?s, such as activin and Nodal (63, 98). Through the transcriptional regulation of three mesodermal T box transcription factors ntl (86), tbx6 (99), and spadetail (spt), a regulator of mesoderm formation in the trunk (100), FGF signaling regulates tail and trunk development in zebrafish (57, 62, 101, 102). As in Xenopus and mouse, the T box genes also serve as an important link between cell fate and morphogenesis. One important role of spt is the regulation of convergence movements of paraxial mesoderm during gastrulation (103) mediated by paraxial protocadherin (papc) (104). Additionally, FGFs may affect gastrulation movements via snail. Consistent with results in mouse (61), overexpression of eFGF causes abnormal widespread expression of snail1 throughout the hypoblast of the late zebrafish gastrula (57).

    In addition to its effects on anterior-posterior (AP) patterning of the embryo and gastrulation movements, FGF signaling affects dorsoventral patterning of the early zebrafish embryo. Activation of the FGF8/FGFR1/Ras signaling pathway leads to the expansion of dorsolateral derivatives at the expense of ventral and posterior domains by an early inhibition of ventral bone morphogenetic protein (BMP) expression (105, 106, 107). It was therefore suggested that FGFs act in concert with secreted BMP antagonists (Spemann organizer factors Follistatin, Noggin, and Chordin) in dorsoventral patterning (105).

    FGFR1 knockouts gave the first evidence that FGF signaling is also essential for proper gastrulation in the mouse (55, 56). Fgfr1–/– mutant embryos die late in gastrulation, exhibiting defects in cell migration, cell fate specification, and patterning. Chimera analysis indicate that Fgfr1-deficient cells are impaired in epithelial to mesenchymal transition (EMT) and, as a consequence, fail to cross the primitive streak (108). The mutant cells accumulate in the primitive streak and overexpress E-cadherin, and this expression is accompanied by the down-regulation of mSnail. mSnail is a key mediator of EMT in development and cancer by repressing E-cadherin (reviewed in Ref.109). Thus, Fgfr1–/– cells may fail to migrate properly because they maintain high levels of E-cadherin in the streak as a result of mSnail down-regulation. The regulation of E-cadherin by FGFR1 via mSnail also provides a molecular link between FGF and Wnt signaling pathways because ectopic E-cadherin expression in Fgfr1–/– embryos attenuates Wnt signaling by sequestering cytoplasmic ?-catenin to the plasma membrane (61).

    Similar to Fgfr mutants, Fgf8–/– embryos display severe gastrulation defects, lacking mesodermal and endodermal tissues. However, epiblast cells move into the streak and undergo EMT in the Fgf8–/– embryos, but fail to move away from the streak (60). This indicates that FGF family members are involved in different stages of mesoderm formation from EMT to subsequent migration and that these two processes are independent.

    B. Chemotactic action of FGFs in morphogenetic movements

    The activity of FGF signaling during gastrulation movements in Xenopus, zebrafish, and mouse has been mostly linked to the regulation of target gene expression, which in turn regulates the migratory behavior of gastrulating cells. However, there is increasing evidence for a direct chemotactic response of migrating cells in response to FGFs both in vertebrates and invertebrates (3, 5). FGF2 and FGF8 function as potent chemoattractants during migration of mesencephalic neural crest cells (110), and FGF2 and FGF4 attract mesenchymal cells during limb bud development in the mouse (111, 112). FGF10, a key regulator of lung-branching morphogenesis, exerts a powerful chemoattractive effect to direct lung epithelial buds to their destinations during development (113). Recently, an elegant study by Yang et al. (114) showed that FGFs act as chemotactic signals to directly coordinate cell movements during gastrulation in the chick embryo. They analyzed the trajectories of cells emerging from different AP positions from the fully extended primitive streak and found that cells show different movement patterns depending on their position along the streak. The cells are guided by signals from the surrounding tissue, and beads soaked in FGFs alter their trajectories. FGF4 acts as a chemoattractant and FGF8b acts as a repellent for migrating streak cells, whereas FGF2 does not elicit a strong chemotactic response (114). In line with the expression of Fgf8 in the primitive streak and Fgf4 in Hensen’s node and forming head process, the movement away from the streak may be the result of a chemorepellent effect of FGF8, whereas the inward movement of cells results from attraction by FGF4 secreted by cells in the head process (114).

    C. FGF signaling in neural induction and AP patterning

    Neural induction is the process whereby naive ectodermal cells are instructed to adopt a neural fate. Several findings, mainly from experiments in Xenopus, led to the default model of neural induction (reviewed in Ref.115): BMP signals in the early embryo instruct ectoderm cells to form epidermis whereas ectodermal cells will adopt a neural fate (their default state) in the absence of cell-cell signaling. In this model, neural induction is mediated during gastrulation by diffusible BMP antagonists released by Spemann’s organizer. However, this model is certainly oversimplified, and increasing evidence suggests a pivotal role of FGF in neural induction and places the onset of neural induction before gastrulation. FGFs act as neural inducers in lower chordates such as ascidians (116) and planarians (117), indicating that this role is evolutionarily conserved.

    The best case for FGFs as neural inducer is the chick embryo. Here, induction of neural tissue precedes the formation of the organizer (Hensen’s node) (118, 119). FGFs are sufficient to induce preneural markers (118, 120, 121, 122), and inhibition of FGF signaling using various inhibitors blocks neural induction by Hensen’s node (118). FGF signaling functions by repressing BMP expression because Fgf3 overexpression leads to a down-regulation of BMP mRNA and the subsequent induction of neural tissue. Conversely, when FGF signaling is inhibited, BMP mRNA expression is maintained and neural fate is blocked (119). A recent study by Sheng et al. (123) suggests a mechanism by which FGFs may block BMP signaling during neural induction in chick. FGFs induce the transcriptional activator Churchhill (ChCh), which is required for the expression of Sip1 (Smad-interacting protein 1) in the neural plate. Sip1 binds to and represses Smad1/5 transcriptional activation of target genes, thereby blocking BMP signaling. ChCh/Sip1 sensitizes cells to neural-inducing signals and represses mesodermal genes (123). However, FGFs promote neural fates also independent of BMP inhibition because beads soaked with the BMP antagonists Noggin or Chordin are insufficient for neural induction under most experimental conditions (119, 124).

    In Xenopus, the role of FGF signaling in neural induction and patterning has been studied extensively. However, conflicting results were obtained concerning the role of FGFs in neural induction. Some reports showed that FGFs could directly induce neural tissue in animal cap explants. However, in most cases a sensitized background was used in which BMP signaling was partially attenuated (23, 125, 126). In addition, inhibition of FGF signaling by dominant negative FGFR1 blocks neural induction by Noggin or Spemann organizer (127), Chordin (128), or notochord tissue (129). On the other hand, several studies using dominant negative FGFR1 obtained contradictory results showing that neural tissue forms apparently even when FGF signaling is inhibited (84, 126, 129, 130, 131). However, these experiments were carried out using the dominant negative form of the FGFR1, but FGF signaling required for anterior neural induction may be mediated through FGFRs other than FGFR1. Indeed, Hongo et al. (132) showed that anterior neural development is more sensitive to inhibition of FGFR4a than to FGFR1. Likewise, FGF8 stimulates neuronal differentiation preferentially through FGFR4a (133), and FGF signaling through FGFR1, but not FGFR4a, is required for neural crest formation (134). The two receptors preferentially activate different downstream pathways, suggesting that they may have distinct functions during development. FGFR4a can signal through PLC in neural induction (50) whereas FGFR1 preferentially activates the Ras-MAPK pathway, involved in posteriorization of already neuralized ectoderm (50, 126).

    In addition to their early effects in neural induction, FGFs are potent posteriorizing agents during AP patterning of the neural plate (135). They are able to convert anterior neural tissue to more posterior neural cell types (50, 125, 126, 131, 136). For example, when FGF-soaked beads are placed in the anterior neural plate, they inhibit anterior neural marker expression (137, 138), and this may involve a cross-talk with the Wnt pathway (130, 139, 140, 141). Similarly, overexpression of FGF3 in zebrafish expands posterior neural tissue (142). Conversely, dominant negative FGFR1 or dominant negative Ras block induction of posterior neural markers (126, 131).

    Patterning of neuroectoderm and mesoderm is linked to the patterned expression of Hox genes along the AP axis of the embryo (143). There is strong evidence from studies in mouse, chick, and Xenopus that FGF signaling regulates early AP patterning via the regulation of Hox genes. Ectopic application or overexpression of FGFs results in the up-regulation of posterior members of the Hox gene family (137, 144, 145, 146, 147). Conversely, overexpression of dominant negative FGFR1 represses posterior Hox gene expression (137, 148). Caudals are immediate-early response genes to FGF signaling and mediate Hox regulation by FGFs (137, 147, 149, 150). A recent study demonstrates that graded FGF signaling is translated into a distinct motor neuron Hox pattern (24).

    An alternative mode of action for FGFs as posteriorizing factor was described for chick neural tube in which they mediate the maintenance of neural stem cells at Hensen’s node. Therefore, FGFs might act to prolong the time window during which cells are exposed to other posteriorizing factors, such as retinoic acid or Wnts, thereby promoting the continuous development of the posterior nervous system (151).

    D. FGF signaling in endoderm formation

    Although endodermal organ development is widely studied, little is known about how endoderm is initially formed and regionally specified in vertebrates. Accumulating data, especially in the mouse, point toward a function of FGF signaling in early endoderm maintenance and patterning. In chimeric embryos, Fgfr1–/– cells are unable to populate endodermal derivatives (61, 108). Similarly, embryoid bodies derived from Fgfr1–/– embryonic stem cells lack visceral endoderm. This is associated with a lowered expression of the endodermal marker -fetoprotein. However, some endoderm is eventually formed, as indicated by the unchanged expression of GATA4 (152). Endoderm deficiency was also obtained in embryoid bodies overexpressing a dominant negative FGFR2, where inhibition of three endodermal markers (-fetoprotein, HNF4, and Evx1) was observed (153). Fgf8–/– embryos develop no embryonic mesoderm- and endoderm-derived tissue, and this is accompanied by inhibition of Fgf4 expression (60). Fgf4–/– embryos die shortly after implantation and fail to make detectable mesoderm and endoderm (154), a phenotype similar to the targeted disruption of Fgfr2 (155, 156). A study by Wells and Melton (157) further supports the role of FGF4 in the specification of primitive endoderm in the mouse because recombinant FGF4 affects the patterning in isolated endoderm in a concentration-dependent manner. Other analyzed FGFs (aFGF, FGF2, 5, 8b) or other growth factors (TGF, TGF?, epidermal growth factor, IGF) had no activity in this assay (157). In conclusion, FGF signaling appears not to be required for induction but rather for the maintenance and patterning of endoderm, because some endodermal markers continue to be expressed even when FGF signaling is inhibited.

    An important caveat in analyzing endoderm development is the potent influence of mesoderm on its differentiation and patterning (158, 159). Thus, factors may regulate endoderm via an effect on mesoderm unless this is rigorously controlled. In Xenopus this can be achieved by explanting vegetal caps, which are devoid of mesoderm and only develop into endoderm. FGF signaling-dependent MAPK activity is detectable in the prospective endodermal cells of the Xenopus blastula embryo (160). However, studies with such vegetal caps gave conflicting results regarding the role of FGF signaling in endoderm. Expression of the dorsal endodermal marker Xlhbox8 and the pan-endodermal marker Mix1 can be blocked by injection of dominant negative FGFR1 (160, 161). In contrast, other studies failed to observe the requirement of FGF signaling for Xlhbox8 expression but rather report an inhibition of Xlhbox8 after FGF treatment (162, 163). Modulation of FGF signaling does not affect the expression of pan-endodermal markers in the majority of vegetal explants experiments, and FGF does not induce a number of endodermal genes in animal caps (128, 164). It is therefore suggested that FGF signaling is not essential for endoderm formation. However, defined levels of FGF activity may be required for endodermal patterning.

    In summary, FGF signaling plays simultaneous roles during gastrulation to induce different cell fates and morphogenetic movements, raising the question of how the specific response to FGFs in different germ layers is mediated. On the one hand, different FGFs may mediate neural or mesodermal induction (e.g. Ref.133). On the other hand, two recent papers show that in the ascidian Ciona and in chick the competence of cells to respond to FGF signals with neural induction is regulated by the neural-specific transcription factors GATA and ChCh, respectively (123, 165). Thus, in addition to qualitative differences among different FGF family members, the spatial and temporal restricted expression of cofactors provides another mechanism to separate the different functions of FGF during gastrulation.

    V. Regulation of FGF Signaling

    Developmental signaling that involves intercellular communication and regulates cell fate must be precisely regulated regarding timing, duration, and spread of such a signal. Precision and robustness of these features can be achieved by both positive and negative feedback loops (reviewed in Ref.166). Negative feedback stabilizes a system when confronted with environmental fluctuations (167). Furthermore, it can limit the absolute level or duration of a signal and can lead to signal oscillations (168). FGF signaling is regulated at many levels in the extracellular space and within the cell. Many of these regulatory mechanisms are not specific for FGF signaling but rather activate or attenuate RTK signaling in general (169). Yet, a growing number of proteins were identified in recent years that specifically regulate the activity of FGFs or signaling pathways activated through FGFRs. Interestingly, many of these regulatory components are themselves regulated by FGF signaling and are tightly coexpressed with Fgfs, forming a synexpression group (107, 168, 170). Synexpression groups are genetic modules in which expression of different genes are closely correlated and whose members function in a common pathway (171). Most of the FGF synexpression group members inhibit FGF signaling and thereby establish negative feedback loops.

    A. Regulation by HSPGs

    An important feature of FGF biology involves the extracellular interaction between FGF and heparin or HSPG (32). HSPGs are sulfated glycosaminoglycans covalently bound to a core protein. Enormous structural heterogeneity can be generated through specific heparin sulfate chain modifications during their biosynthesis, as well as from the diverse nature of their core proteins (reviewed in Ref.172). HSPGs are required for FGFs to effectively activate the FGFR and determine interaction between individual FGF family members with certain FGFRs (173, 174). The crystal structure of ternary FGF-heparin-FGFR complexes have been determined; heparin bridges two FGFRs to form FGFR dimers, which bind two FGFs (175). In an elegant study, a ligand and carbohydrate engagement assay was employed to examine the spatio-temporal changes in HS to promote ternary complex formation with FGFs and FGFRs. These changes depend on glycosyltransferases and sugar sulfotransferases. The analysis reveals that synthesis of HS domains is tightly regulated throughout development, and different FGF-FGFR pairs have unique HS-binding requirements (33). Other than promoting FGF binding to its receptors, HSPGs protect the FGFs from thermal denaturation and proteolysis (176). In addition, these interactions limit FGF diffusion and release into interstitial spaces, thereby regulating FGF activity within a tissue (177, 178). The importance of HSPGs and enzymes involved in HS biosynthesis for FGF signaling during development has been demonstrated in vivo in a number of studies (68, 179, 180, 181). For example, mice mutated in UDP-glucose dehydrogenase (Ugdh), one of the key enzymes for HS biosynthesis, fail to gastrulate properly and show defects of mesoderm and endoderm migration similar to FGF pathway mutants (181). In the Ugdh mutant embryos, FGF signaling is specifically blocked whereas signaling by Wnt3 and Nodal appears unaffected. Ugdh mutant Drosophila embryos also have phenotypes similar to FGFR mutants (180), indicating an evolutionary conserved requirement for HSPGs in FGF signaling.

    The nature of the proteoglycan core protein essential for FGF signal transduction has not been identified because single knock-out mice mutant for diverse core proteins (including Syndecan1, -3, and -4; Glypican2 and -3; and Perlecan) survive past gastrulation.

    B. Transmembrane regulators

    An emerging theme is the regulation of FGF signaling by transmembrane modulators. The transmembrane protein Sef is a member of the FGF synexpression group in zebrafish and mouse, expressed in the characteristic pattern of FGF8 and regulated by FGF signaling (107, 170, 182). Overexpression studies in zebrafish embryos show that Sef inhibits FGF signaling via interaction with FGFR1 and -4a (170) and that it functions at the level or downstream of MAPK kinase (107). Interfering with Sef function by antisense morpholino oligonucleotide (MO) injection induces phenotypic changes reminiscent of embryos dorsalized by ectopic expression of Fgf8 (107, 170). The molecular mechanism of Sef action is controversial. Sef may inhibit FGF signaling at the level of FGFR (183, 213). Overexpression of mouse Sef inhibits tyrosine-phosphorylation of FGFR1 and FRS2 and inhibits activation of the MAP kinase and Akt pathway, possibly due to reduced recruitment of Grb2 to FRS2 (183). In contrast, other reports argue Sef acts downstream of or at the level of MEK (107, 214, 215, 216). Torii et al. (216) showed that hSef acts to restrict ERK activity to the cytoplasm. hSef binds to activated forms of MEK and inhibits the dissociation of the MEK-ERK complexes. Thereby, nuclear translocation of ERK is blocked without inhibiting its activity in the cytoplasm (216).

    In Xenopus, XFLRT3, a member of a family of leucine-rich repeat transmembrane proteins, is another component of the FGF signaling pathway (184). Its expression is very similar to that of the FGF8 synexpression group, and it is regulated by FGFs. When overexpressed, FLRT3 phenocopies FGF signaling and activates the MAPK pathway. Inhibition of XFLRTs by MO injection interferes with FGF signaling. XFLRT3 binds to FGFRs and modulates FGFR signaling by an unknown mechanism. In cell culture, FLRT3 promotes neurite outgrowth of primary cerebella granule neurons (217). Since FGF promotes axon outgrowth (218, 219, 220), it is possible that potentiation of FGF signaling by FLRT3 underlies its effects on neurite growth.

    Finally, FGFRs interact with CAMs such as N-CAM, N-cadherin, and L1 via their extracellular domains. In neurons, N-CAM binding to FGFR1 activates the PLC/Ca2+ pathway (49, 185, 186). In tumor ?-cells and L cells, Cavallaro et al. (187) show a direct association of FGFR4 and N-CAM. This interaction induces the formation of a multiprotein signaling complex that mediates neurite outgrowth and cell-matrix adhesion (187).

    C. Intracellular regulators

    1. PTPs.

    It is well established that RTK signaling is controlled by PTPs, which act as positive or negative regulators on RTK signaling (reviewed in Refs.188 and 189). The dual-specificity PTP MAPK phosphatase 3 (MKP3, also called Pyst1) antagonizes the MAPK pathway via ERK1/2 inactivation. In vertebrates, its expression is very similar to that of the FGF8 synexpression group, and it is regulated by FGF8 (190, 191). In the developing chick limb MKP3 mediates the antiapoptotic signaling of FGF8 (191), and manipulation of MKP3 activity by short interfering RNA inhibition or overexpression disrupts limb morphology and outgrowth (190, 191). In contrast, the PTP SHP-2 and Xenopus low-molecular weight PTP1 (XLPTP1) act as positive regulators of FGF signaling upstream of Ras during Xenopus embryogenesis (71, 72). Inhibition of XLPTP1 by antisense MOs blocks FGF-induced MAPK activation, leading to a shortened AP axis (72).

    2. Sprouty protein.

    The feedback inhibitor Sprouty was identified in Drosophila as a negative regulator of FGF signaling during tracheal development (192). Four mammalian Sprouty proteins (Spry1–4) and three Spreds (Sprouty-related EVH1 domain proteins) that share a highly conserved cysteine-rich domain at the carboxy terminus have been identified. This domain is critical to target them to the plasma membrane and to inhibit the MAPK pathway (193). Sprouty2 and -4 are members of the FGF8 synexpression group and are regulated by FGF signaling (106, 194, 195, 196). Sprouty proteins interfere with FGF signaling through several mechanisms, depending on the cellular context and/or stimulation. Different Sprouty proteins exhibit different activities and have different interaction partners, including but not limited to Grb2 (197) and Raf1 (198) (reviewed in Ref.199).

    Gain- and loss-of-function experiments confirm the function of Sprouty proteins as intracellular antagonists of FGF signaling. Reduction of mSprouty2 expression during mouse lung development using antisense oligonucleotides established Sprouty as an inhibitor of branching morphogenesis, as in Drosophila (192, 200). Overexpressing Sprouty during chick limb development causes a reduction in limb bud outgrowth consistent with reduced FGF signaling (195). Furthermore, Fürthauer et al. (106) showed that inhibiting Sprouty4 in zebrafish by MO injection elicits phenotypes similar to that obtained after up-regulation of Ffg8. Sprouty4 is required for dorsoventral mesoderm patterning and for head development (106). In Xenopus embryos, XSprouty2 regulates gastrulation movements by a Ca2+-dependent pathway (92).

    VI. Concluding Remarks

    Substantial progress has been made in recent years in the understanding of the role of FGF signaling during early vertebrate development. FGFs play an important role in regulating cell fate and specification of all three germ layers. The role of FGF signaling during mesoderm formation and neural induction is now well documented and appears evolutionarily conserved in chordates. A surprising new role is the ability of FGFs to orchestrate gastrulation movements by acting as chemoattractants and repellents. This probably involves direct regulation of the cytoskeleton.

    FGF downstream pathways and the components employed are characterized in some detail. A growing number of these components are themselves tightly regulated by FGFs and form an FGF8 synexpression group. The FGF8 synexpression group is an autoregulatory genetic module, conserved from frogs to mice, which is employed, for example, in mesoderm formation and neural patterning during early development. Part of this group are the novel transmembrane regulators Sef and FLRT3, which bind to and modulate FGFRs, but their precise function needs further investigation. It has also become clear that different ligand receptor pairs may preferentially couple to different pathways but how this is regulated is unclear. In the future it will be important to assign the pathways employed by FGFs in different cells and define the determinants for signaling specificity. One emerging class of determinants are HSPGs and the enzymes involved in their synthesis. Two classes of downstream target genes, which mediate FGF signaling are T-box and Hox genes, and this is evolutionary conserved in all vertebrates. Furthermore, there is substantial cross-talk between FGFs with TGF-?- and Wnt signaling pathways. We are only beginning to understand the mechanisms of these cross-regulations, which are at the heart of the ability of FGFs to act as competence factor, e.g. during mesoderm induction by Nodals. Finally, do FGFs act as morphogens? They have many characteristics of morphogens, including their ability to induce marker genes in a dose-dependent manner. However, the proof that FGFs act directly in a concentration-dependent fashion rather than by relay and the demonstration of a physiological requirement for this is missing.

    Acknowledgments

    We thank Dr. Sonia Pinho for advice and critical reading of the manuscript.

    First Published Online October 26, 2004

    Abbreviations: AP, Anterior-posterior; BMP, bone morphogenetic protein; CAM, cell adhesion molecule; EMT, epithelial to mesenchymal transition; FGF, fibroblast growth factor; FGFR, FGF receptor; HS, heparan sulfate; HSPG, heparan sulfate proteoglycan; MO, morpholino oligonucleotide; PI3, phosphatidylinositol-3; PLC, phospholipase C; PTP, protein tyrosine phosphatase; RTK, receptor tyrosine kinase; SOS, Son of sevenless.

    References

    Armelin HA 1973 Pituitary extracts and steroid hormones in the control of 3T3 cell growth. Proc Natl Acad Sci USA 70:2702–2706

    Forbes A, Lehmann R 1999 Cell migration in Drosophila. Curr Opin Genet Dev 9:473–478

    Wilson R, Leptin M 2000 Fibroblast growth factor receptor-dependent morphogenesis of the Drosophila mesoderm. Philos Trans R Soc Lond B Biol Sci 355:891–895

    Zelzer E, Shilo BZ 2000 Cell fate choices in Drosophila tracheal morphogenesis. Bioessays 22:219–226

    Montell DJ 1999 The genetics of cell migration in Drosophila melanogaster and Caenorhabditis elegans development. Development 126:3035–3046

    Borland CZ, Schutzman JL, Stern MJ 2001 Fibroblast growth factor signaling in Caenorhabditis elegans. Bioessays 23:1120–1130

    Martin GR 1998 The roles of FGFs in the early development of vertebrate limbs. Genes Dev 12:1571–1586

    Capdevila J, Izpisua Belmonte JC 2001 Patterning mechanisms controlling vertebrate limb development. Annu Rev Cell Dev Biol 17:87–132

    Wilkie AO, Patey SJ, Kan SH, van den Ouweland AM, Hamel BC 2002 FGFs, their receptors, and human limb malformations: clinical and molecular correlations. Am J Med Genet 112:266–278

    Ford-Perriss M, Abud H, Murphy M 2001 Fibroblast growth factors in the developing central nervous system. Clin Exp Pharmacol Physiol 28:493–503

    Dono R 2003 Fibroblast growth factors as regulators of central nervous system development and function. Am J Physiol Regul Integr Comp Physiol 284:R867–R881

    Ornitz DM, Marie PJ 2002 FGF signaling pathways in endochondral and intramembranous bone development and human genetic disease. Genes Dev 16:1446–1465

    Dickson C, Spencer-Dene B, Dillon C, Fantl V 2000 Tyrosine kinase signalling in breast cancer: fibroblast growth factors and their receptors. Breast Cancer Res 2:191–196

    Moroni E, Dell’Era P, Rusnati M, Presta M 2002 Fibroblast growth factors and their receptors in hematopoiesis and hematological tumors. J Hematother Stem Cell Res 11:19–32

    Cronauer MV, Schulz WA, Seifert HH, Ackermann R, Burchardt M 2003 Fibroblast growth factors and their receptors in urological cancers: basic research and clinical implications. Eur Urol 43:309–319

    Ornitz DM, Itoh N 2001 Fibroblast growth factors. Genome Biol 2:REVIEWS3005

    Eriksson AE, Cousens LS, Weaver LH, Matthews BW 1991 Three-dimensional structure of human basic fibroblast growth factor. Proc Natl Acad Sci USA 88:3441–3445

    Zhu X, Komiya H, Chirino A, Faham S, Fox GM, Arakawa T, Hsu BT, Rees DC 1991 Three-dimensional structures of acidic and basic fibroblast growth factors. Science 251:90–93

    Plotnikov AN, Hubbard SR, Schlessinger J, Mohammadi M 2000 Crystal structures of two FGF-FGFR complexes reveal the determinants of ligand-receptor specificity. Cell 101:413–424

    Szebenyi G, Fallon JF 1999 Fibroblast growth factors as multifunctional signaling factors. Int Rev Cytol 185:45–106

    Christen B, Slack JM 1999 Spatial response to fibroblast growth factor signalling in Xenopus embryos. Development 126:119–125

    Green JB, New HV, Smith JC 1992 Responses of embryonic Xenopus cells to activin and FGF are separated by multiple dose thresholds and correspond to distinct axes of the mesoderm. Cell 71:731–739

    Kengaku M, Okamoto H 1995 bFGF as a possible morphogen for the anteroposterior axis of the central nervous system in Xenopus. Development 121:3121–3130

    Dasen JS, Liu JP, Jessell TM 2003 Motor neuron columnar fate imposed by sequential phases of Hox-c activity. Nature 425:926–933

    Johnson DE, Williams LT 1993 Structural and functional diversity in the FGF receptor multigene family. Adv Cancer Res 60:1–41

    Groth C, Lardelli M 2002 The structure and function of vertebrate fibroblast growth factor receptor 1. Int J Dev Biol 46:393–400

    Ornitz DM, Xu J, Colvin JS, McEwen DG, MacArthur CA, Coulier F, Gao G, Goldfarb M 1996 Receptor specificity of the fibroblast growth factor family. J Biol Chem 271:15292–15297

    Xu J, Liu Z, Ornitz DM 2000 Temporal and spatial gradients of Fgf8 and Fgf17 regulate proliferation and differentiation of midline cerebellar structures. Development 127:1833–1843

    Avivi A, Yayon A, Givol D 1993 A novel form of FGF receptor-3 using an alternative exon in the immunoglobulin domain III. FEBS Lett 330:249–252

    Gilbert E, Del Gatto F, Champion-Arnaud P, Gesnel MC, Breathnach R 1993 Control of BEK and K-SAM splice sites in alternative splicing of the fibroblast growth factor receptor 2 pre-mRNA. Mol Cell Biol 13:5461–5468

    Orr-Urtreger A, Bedford MT, Burakova T, Arman E, Zimmer Y, Yayon A, Givol D, Lonai P 1993 Developmental localization of the splicing alternatives of fibroblast growth factor receptor-2 (FGFR2). Dev Biol 158:475–486

    Ornitz DM 2000 FGFs, heparan sulfate and FGFRs: complex interactions essential for development. Bioessays 22:108–112

    Allen BL, Rapraeger AC 2003 Spatial and temporal expression of heparan sulfate in mouse development regulates FGF and FGF receptor assembly. J Cell Biol 163:637–648

    McKeehan WL, Wang F, Kan M 1998 The heparan sulfate-fibroblast growth factor family: diversity of structure and function. Prog Nucleic Acid Res Mol Biol 59:135–176

    Schlessinger J 2000 Cell signaling by receptor tyrosine kinases. Cell 103:211–225

    Pawson T, Olivier P, Rozakis-Adcock M, McGlade J, Henkemeyer M 1993 Proteins with SH2 and SH3 domains couple receptor tyrosine kinases to intracellular signalling pathways. Philos Trans R Soc Lond B Biol Sci 340:279–285

    Forman-Kay JD, Pawson T 1999 Diversity in protein recognition by PTB domains. Curr Opin Struct Biol 9:690–695

    Wang JK, Xu H, Li HC, Goldfarb M 1996 Broadly expressed SNT-like proteins link FGF receptor stimulation to activators of Ras. Oncogene 13:721–729

    Kouhara H, Hadari YR, Spivak-Kroizman T, Schilling J, Bar-Sagi D, Lax I, Schlessinger J 1997 A lipid-anchored Grb2-binding protein that links FGF-receptor activation to the Ras/MAPK signaling pathway. Cell 89:693–702

    Ong SH, Guy GR, Hadari YR, Laks S, Gotoh N, Schlessinger J, Lax I 2000 FRS2 proteins recruit intracellular signaling pathways by binding to diverse targets on fibroblast growth factor and nerve growth factor receptors. Mol Cell Biol 20:979–989

    Hadari YR, Gotoh N, Kouhara H, Lax I, Schlessinger J 2001 Critical role for the docking-protein FRS2 in FGF receptor-mediated signal transduction pathways. Proc Natl Acad Sci USA 98:8578–8583

    Kusakabe M, Masuyama N, Hanafusa H, Nishida E 2001 Xenopus FRS2 is involved in early embryogenesis in cooperation with the Src family kinase Laloo. EMBO Rep 2:727–735

    Akagi K, Kyun Park E, Mood K, Daar IO 2002 Docking protein SNT1 is a critical mediator of fibroblast growth factor signaling during Xenopus embryonic development. Dev Dyn 223:216–228

    Hadari YR, Kouhara H, Lax I, Schlessinger J 1998 Binding of Shp2 tyrosine phosphatase to FRS2 is essential for fibroblast growth factor-induced PC12 cell differentiation. Mol Cell Biol 18:3966–3973

    Lee Jr JT, McCubrey JA 2002 The Raf/MEK/ERK signal transduction cascade as a target for chemotherapeutic intervention in leukemia. Leukemia 16:486–507

    Mohammadi M, Dionne CA, Li W, Li N, Spivak T, Honegger AM, Jaye M, Schlessinger J 1992 Point mutation in FGF receptor eliminates phosphatidylinositol hydrolysis without affecting mitogenesis. Nature 358:681–684

    Peters KG, Marie J, Wilson E, Ives HE, Escobedo J, Del Rosario M, Mirda D, Williams LT 1992 Point mutation of an FGF receptor abolishes phosphatidylinositol turnover and Ca2+ flux but not mitogenesis. Nature 358:678–681

    Doherty P, Walsh FS 1996 CAM-FGF receptor interactions: a model for axonal growth. Mol Cell Neurosci 8:99–111

    Hall H, Williams EJ, Moore SE, Walsh FS, Prochiantz A, Doherty P 1996 Inhibition of FGF-stimulated phosphatidylinositol hydrolysis and neurite outgrowth by a cell-membrane permeable phosphopeptide. Curr Biol 6:580–587

    Umbhauer M, Penzo-Mendez A, Clavilier L, Boucaut J, Riou J 2000 Signaling specificities of fibroblast growth factor receptors in early Xenopus embryo. J Cell Sci 113:2865–2875

    Carballada R, Yasuo H, Lemaire P 2001 Phosphatidylinositol-3 kinase acts in parallel to the ERK MAP kinase in the FGF pathway during Xenopus mesoderm induction. Development 128:35–44

    Slack JM, Darlington BG, Heath JK, Godsave SF 1987 Mesoderm induction in early Xenopus embryos by heparin-binding growth factors. Nature 326:197–200

    Nishida H 2002 Patterning the marginal zone of early ascidian embryos: localized maternal mRNA and inductive interactions. Bioessays 24:613–624

    Amaya E, Musci TJ, Kirschner MW 1991 Expression of a dominant negative mutant of the FGF receptor disrupts mesoderm formation in Xenopus embryos. Cell 66:257–270

    Deng CX, Wynshaw-Boris A, Shen MM, Daugherty C, Ornitz DM, Leder P 1994 Murine FGFR-1 is required for early postimplantation growth and axial organization. Genes Dev 8:3045–3057

    Yamaguchi TP, Harpal K, Henkemeyer M, Rossant J 1994 fgfr-1 is required for embryonic growth and mesodermal patterning during mouse gastrulation. Genes Dev 8:3032–3044

    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 121:2983–2994

    Smith JC, Price BM, Green JB, Weigel D, Herrmann BG 1991 Expression of a Xenopus homolog of Brachyury (T) is an immediate-early response to mesoderm induction. Cell 67:79–87

    Strong CF, Barnett MW, Hartman D, Jones EA, Stott D 2000 Xbra3 induces mesoderm and neural tissue in Xenopus laevis. Dev Biol 222:405–419

    Sun X, Meyers EN, Lewandoski M, Martin GR 1999 Targeted disruption of Fgf8 causes failure of cell migration in the gastrulating mouse embryo. Genes Dev 13:1834–1846

    Ciruna B, Rossant J 2001 FGF signaling regulates mesoderm cell fate specification and morphogenetic movement at the primitive streak. Dev Cell 1:37–49

    Griffin KJ, Amacher SL, Kimmel CB, Kimelman D 1998 Molecular identification of spadetail: regulation of zebrafish trunk and tail mesoderm formation by T-box genes. Development 125:3379–3388

    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 22:6078–6088

    Isaacs HV 1997 New perspectives on the role of the fibroblast growth factor family in amphibian development. Cell Mol Life Sci 53:350–361

    Kimelman D, Kirschner M 1987 Synergistic induction of mesoderm by FGF and TGF-? and the identification of an mRNA coding for FGF in the early Xenopus embryo. Cell 51:869–877

    Kimelman D, Abraham JA, Haaparanta T, Palisi TM, Kirschner MW 1988 The presence of fibroblast growth factor in the frog egg: its role as a natural mesoderm inducer. Science 242:1053–1056

    Neilson KM, Friesel R 1996 Ligand-independent activation of fibroblast growth factor receptors by point mutations in the extracellular, transmembrane, and kinase domains. J Biol Chem 271:25049–25057

    Itoh K, Sokol SY 1994 Heparan sulfate proteoglycans are required for mesoderm formation in Xenopus embryos. Development 120:2703–2711

    Gupta RW, Mayer BJ 1998 Dominant-negative mutants of the SH2/SH3 adapters Nck and Grb2 inhibit MAP kinase activation and mesoderm-specific gene induction by eFGF in Xenopus. Oncogene 17:2155–2165

    Weinstein DC, Marden J, Carnevali F, Hemmati-Brivanlou A 1998 FGF-mediated mesoderm induction involves the Src-family kinase Laloo. Nature 394:904–908

    Tang TL, Freeman Jr RM, O’Reilly AM, Neel BG, Sokol SY 1995 The SH2-containing protein-tyrosine phosphatase SH-PTP2 is required upstream of MAP kinase for early Xenopus development. Cell 80:473–483

    Park EK, Warner N, Mood K, Pawson T, Daar IO 2002 Low-molecular-weight protein tyrosine phosphatase is a positive component of the fibroblast growth factor receptor signaling pathway. Mol Cell Biol 22:3404–3414

    Whitman M, Melton DA 1992 Involvement of p21ras in Xenopus mesoderm induction. Nature 357:252–254

    MacNicol AM, Muslin AJ, Williams LT 1993 Raf-1 kinase is essential for early Xenopus development and mediates the induction of mesoderm by FGF. Cell 73:571–583

    LaBonne C, Burke B, Whitman M 1995 Role of MAP kinase in mesoderm induction and axial patterning during Xenopus development. Development 121:1475–1486

    Umbhauer M, Marshall CJ, Mason CS, Old RW, Smith JC 1995 Mesoderm induction in Xenopus caused by activation of MAP kinase. Nature 376:58–62

    Dong Z, Xu RH, Kim J, et al 1996 AP-1/jun is required for early Xenopus development and mediates mesoderm induction by fibroblast growth factor but not by activin. J Biol Chem 271:9942–9946

    Kawachi K, Masuyama N, Nishida E 2003 Essential role of the transcription factor Ets-2 in Xenopus early development. J Biol Chem 278:5473–5477

    Cornell RA, Kimelman D 1994 Activin-mediated mesoderm induction requires FGF. Development 120:453–462

    LaBonne C, Whitman M 1994 Mesoderm induction by activin requires FGF-mediated intracellular signals. Development 120:463–472

    Cornell RA, Musci TJ, Kimelman D 1995 FGF is a prospective competence factor for early activin-type signals in Xenopus mesoderm induction. Development 121:2429–2437

    Mitchell TS, Sheets MD 2001 The FGFR pathway is required for the trunk-inducing functions of Spemann’s organizer. Dev Biol 237:295–305

    Kretzschmar M, Doody J, Timokhina I, Massague J 1999 A mechanism of repression of TGF?/ Smad signaling by oncogenic Ras. Genes Dev 13:804–816

    Kroll KL, Amaya E 1996 Transgenic Xenopus embryos from sperm nuclear transplantations reveal FGF signaling requirements during gastrulation. Development 122:3173–3183

    Herrmann BG, Labeit S, Poustka A, King TR, Lehrach H 1990 Cloning of the T gene required in mesoderm formation in the mouse. Nature 343:617–622

    Halpern ME, Ho RK, Walker C, Kimmel CB 1993 Induction of muscle pioneers and floor plate is distinguished by the zebrafish no tail mutation. Cell 75:99–111

    Conlon FL, Sedgwick SG, Weston KM, Smith JC 1996 Inhibition of Xbra transcription activation causes defects in mesodermal patterning and reveals autoregulation of Xbra in dorsal mesoderm. Development 122:2427–2435

    Isaacs HV, Pownall ME, Slack JM 1994 eFGF regulates Xbra expression during Xenopus gastrulation. EMBO J 13:4469–4481

    Schulte-Merker S, Smith JC 1995 Mesoderm formation in response to Brachyury requires FGF signalling. Curr Biol 5:62–67

    Standley HJ, Zorn AM, Gurdon JB 2001 eFGF and its mode of action in the community effect during Xenopus myogenesis. Development 128:1347–1357

    Fisher ME, Isaacs HV, Pownall ME 2002 eFGF is required for activation of XmyoD expression in the myogenic cell lineage of Xenopus laevis. Development 129:1307–1315

    Nutt SL, Dingwell KS, Holt CE, Amaya E 2001 Xenopus Sprouty2 inhibits FGF-mediated gastrulation movements but does not affect mesoderm induction and patterning. Genes Dev 15:1152–1166

    Conlon FL, Smith JC 1999 Interference with brachyury function inhibits convergent extension, causes apoptosis, and reveals separate requirements in the FGF and activin signalling pathways. Dev Biol 213:85–100

    Kwan KM, Kirschner MW 2003 Xbra functions as a switch between cell migration and convergent extension in the Xenopus gastrula. Development 130:1961–1972

    Tada M, Smith JC 2000 Xwnt11 is a target of Xenopus Brachyury: regulation of gastrulation movements via Dishevelled, but not through the canonical Wnt pathway. Development 127:2227–2238

    Takeuchi M, Nakabayashi J, Sakaguchi T, Yamamoto TS, Takahashi H, Takeda H, Ueno N 2003 The prickle-related gene in vertebrates is essential for gastrulation cell movements. Curr Biol 13:674–679

    Heisenberg CP, Tada M, Rauch GJ, Saude L, Concha ML, Geisler R, Stemple DL, Smith JC, Wilson SW 2000 Silberblick/Wnt11 mediates convergent extension movements during zebrafish gastrulation. Nature 405:76–81

    Rodaway A, Takeda H, Koshida S, Broadbent J, Price B, Smith JC, Patient R, Holder N 1999 Induction of the mesendoderm in the zebrafish germ ring by yolk cell-derived TGF-? family signals and discrimination of mesoderm and endoderm by FGF. Development 126:3067–3078

    Hug B, Walter V, Grunwald DJ 1997 tbx6, a Brachyury-related gene expressed by ventral mesendodermal precursors in the zebrafish embryo. Dev Biol 183:61–73

    Kimmel CB, Kane DA, Walker C, Warga RM, Rothman MB 1989 A mutation that changes cell movement and cell fate in the zebrafish embryo. Nature 337:358–362

    Goering LM, Hoshijima K, Hug B, Bisgrove B, Kispert A, Grunwald DJ 2003 An interacting network of T-box genes directs gene expression and fate in the zebrafish mesoderm. Proc Natl Acad Sci USA 100:9410–9415

    Griffin KJ, Kimelman D 2003 Interplay between FGF, one-eyed pinhead, and T-box transcription factors during zebrafish posterior development. Dev Biol 264:456–466

    Ho RK, Kane DA 1990 Cell-autonomous action of zebrafish spt-1 mutation in specific mesodermal precursors. Nature 348:728–730

    Yamamoto A, Amacher SL, Kim SH, Geissert D, Kimmel CB, De Robertis EM 1998 Zebrafish paraxial protocadherin is a downstream target of spadetail involved in morphogenesis of gastrula mesoderm. Development 125:3389–3397

    Fürthauer M, Thisse C, Thisse B 1997 A role for FGF-8 in the dorsoventral patterning of the zebrafish gastrula. Development 124:4253–4264

    Fürthauer M, Reifers F, Brand M, Thisse B, Thisse C 2001 Sprouty4 acts in vivo as a feedback-induced antagonist of FGF signaling in zebrafish. Development 128:2175–2186

    Fürthauer M, Lin W, Ang SL, Thisse B, Thisse C 2002 Sef is a feedback-induced antagonist of Ras/MAPK-mediated FGF signalling. Nat Cell Biol 4:170–174

    Ciruna BG, Schwartz L, Harpal K, Yamaguchi TP, Rossant J 1997 Chimeric analysis of fibroblast growth factor receptor-1 (Fgfr1) function: a role for FGFR1 in morphogenetic movement through the primitive streak. Development 124:2829–2841

    Hemavathy K, Ashraf SI, Ip YT 2000 Snail/slug family of repressors: slowly going into the fast lane of development and cancer. Gene 257:1–12

    Kubota Y, Ito K 2000 Chemotactic migration of mesencephalic neural crest cells in the mouse. Dev Dyn 217:170–179

    Webb SE, Lee KK, Tang MK, Ede DA 1997 Fibroblast growth factors 2 and 4 stimulate migration of mouse embryonic limb myogenic cells. Dev Dyn 209:206–216

    Li S, Muneoka K 1999 Cell migration and chick limb development: chemotactic action of FGF-4 and the AER. Dev Biol 211:335–347

    Park WY, Miranda B, Lebeche D, Hashimoto G, Cardoso WV 1998 FGF-10 is a chemotactic factor for distal epithelial buds during lung development. Dev Biol 201:125–134

    Yang X, Dormann D, Munsterberg AE, Weijer CJ 2002 Cell movement patterns during gastrulation in the chick are controlled by positive and negative chemotaxis mediated by FGF4 and FGF8. Dev Cell 3:425–437

    Weinstein DC, Hemmati-Brivanlou A 1999 Neural induction. Annu Rev Cell Dev Biol 15:411–433

    Lemaire P, Bertrand V, Hudson C 2002 Early steps in the formation of neural tissue in ascidian embryos. Dev Biol 252:151–169

    Cebria F, Kobayashi C, Umesono Y, Nakazawa M, Mineta K, Ikeo K, Gojobori T, Itoh M, Taira M, Sanchez Alvarado A, Agata K 2002 FGFR-related gene nou-darake restricts brain tissues to the head region of planarians. Nature 419:620–624

    Streit A, Berliner AJ, Papanayotou C, Sirulnik A, Stern CD 2000 Initiation of neural induction by FGF signalling before gastrulation. Nature 406:74–78

    Wilson SI, Graziano E, Harland R, Jessell TM, Edlund T 2000 An early requirement for FGF signalling in the acquisition of neural cell fate in the chick embryo. Curr Biol 10:421–429

    Alvarez IS, Araujo M, Nieto MA 1998 Neural induction in whole chick embryo cultures by FGF. Dev Biol 199:42–54

    Storey KG, Goriely A, Sargent CM, et al 1998 Early posterior neural tissue is induced by FGF in the chick embryo. Development 125:473–484

    Muhr J, Graziano E, Wilson S, Jessell TM, Edlund T 1999 Convergent inductive signals specify midbrain, hindbrain, and spinal cord identity in gastrula stage chick embryos. Neuron 23:689–702

    Sheng G, dos Reis M, Stern CD 2003 Churchill, a zinc finger transcriptional activator, regulates the transition between gastrulation and neurulation. Cell 115:603–616

    Streit A, Lee KJ, Woo I, Roberts C, Jessell TM, Stern CD 1998 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

    Lamb TM, Harland RM 1995 Fibroblast growth factor is a direct neural inducer, which combined with noggin generates anterior-posterior neural pattern. Development 121:3627–3636

    Ribisi Jr S, Mariani FV, Aamar E, Lamb TM, Frank D, Harland RM 2000 Ras-mediated FGF signaling is required for the formation of posterior but not anterior neural tissue in Xenopus laevis. Dev Biol 227:183–196

    Launay C, Fromentoux V, Shi DL, Boucaut JC 1996 A truncated FGF receptor blocks neural induction by endogenous Xenopus inducers. Development 122:869–880

    Sasai Y, Lu B, Piccolo S, De Robertis EM 1996 Endoderm induction by the organizer-secreted factors chordin and noggin in Xenopus animal caps. EMBO J 15:4547–4555

    Barnett MW, Old RW, Jones EA 1998 Neural induction and patterning by fibroblast growth factor, notochord and somite tissue in Xenopus. Dev Growth Differ 40:47–57

    McGrew LL, Hoppler S, Moon RT 1997 Wnt and FGF pathways cooperatively pattern anteroposterior neural ectoderm in Xenopus. Mech Dev 69:105–114

    Holowacz T, Sokol S 1999 FGF is required for posterior neural patterning but not for neural induction. Dev Biol 205:296–308

    Hongo I, Kengaku M, Okamoto H 1999 FGF signaling and the anterior neural induction in Xenopus. Dev Biol 216:561–581

    Hardcastle Z, Chalmers AD, Papalopulu N 2000 FGF-8 stimulates neuronal differentiation through FGFR-4a and interferes with mesoderm induction in Xenopus embryos. Curr Biol 10:1511–1514

    Monsoro-Burq AH, Fletcher RB, Harland RM 2003 Neural crest induction by paraxial mesoderm in Xenopus embryos requires FGF signals. Development 130:3111–3124

    Doniach T 1995 Basic FGF as an inducer of anteroposterior neural pattern. Cell 83:1067–1070

    Cox WG, Hemmati-Brivanlou A 1995 Caudalization of neural fate by tissue recombination and bFGF. Development 121:4349–4358

    Pownall ME, Tucker AS, Slack JM, Isaacs HV 1996 eFGF, Xcad3 and Hox genes form a molecular pathway that establishes the anteroposterior axis in Xenopus. Development 122:3881–3892

    Lombardo A, Slack JM 1998 Postgastrulation effects of fibroblast growth factor on Xenopus development. Dev Dyn 212:75–85

    Kazanskaya O, Glinka A, Niehrs C 2000 The role of Xenopus dickkopf1 in prechordal plate specification and neural patterning. Development 127:4981–4992

    Domingos PM, Itasaki N, Jones CM, Mercurio S, Sargent MG, Smith JC, Krumlauf R 2001 The Wnt/?-catenin pathway posteriorizes neural tissue in Xenopus by an indirect mechanism requiring FGF signalling. Dev Biol 239:148–160

    Kudoh T, Wilson SW, Dawid IB 2002 Distinct roles for Fgf, Wnt and retinoic acid in posteriorizing the neural ectoderm. Development 129:4335–4346

    Koshida S, Shinya M, Nikaido M, Ueno N, Schulte-Merker S, Kuroiwa A, Takeda H 2002 Inhibition of BMP activity by the FGF signal promotes posterior neural development in zebrafish. Dev Biol 244:9–20

    Krumlauf R 1994 Hox genes in vertebrate development. Cell 78:191–201

    Cho KW, De Robertis EM 1990 Differential activation of Xenopus homeo box genes by mesoderm-inducing growth factors and retinoic acid. Genes Dev 4:1910–1916

    Kolm PJ, Sive HL 1995 Regulation of the Xenopus labial homeodomain genes, HoxA1 and HoxD1: activation by retinoids and peptide growth factors. Dev Biol 167:34–49

    Partanen J, Schwartz L, Rossant J 1998 Opposite phenotypes of hypomorphic and Y766 phosphorylation site mutations reveal a function for Fgfr1 in anteroposterior patterning of mouse embryos. Genes Dev 12:2332–2344

    Bel-Vialar S, Itasaki N, Krumlauf R 2002 Initiating Hox gene expression: in the early chick neural tube differential sensitivity to FGF and RA signaling subdivides the HoxB genes in two distinct groups. Development 129:5103–5115

    Godsave SF, Durston AJ 1997 Neural induction and patterning in embryos deficient in FGF signaling. Int J Dev Biol 41:57–65

    Isaacs HV, Pownall ME, Slack JM 1998 Regulation of Hox gene expression and posterior development by the Xenopus caudal homologue Xcad3. EMBO J 17:3413–3427

    Pownall ME, Isaacs HV, Slack JM 1998 Two phases of Hox gene regulation during early Xenopus development. Curr Biol 8:673–676

    Mathis L, Kulesa PM, Fraser SE 2001 FGF receptor signalling is required to maintain neural progenitors during Hensen’s node progression. Nat Cell Biol 3:559–566

    Esner M, Pachernik J, Hampl A, Dvorak P 2002 Targeted disruption of fibroblast growth factor receptor-1 blocks maturation of visceral endoderm and cavitation in mouse embryoid bodies. Int J Dev Biol 46:817–825

    Chen Y, Li X, Eswarakumar VP, Seger R, Lonai P 2000 Fibroblast growth factor (FGF) signaling through PI 3-kinase and Akt/PKB is required for embryoid body differentiation. Oncogene 19:3750–3756

    Feldman B, Poueymirou W, Papaioannou VE, DeChiara TM, Goldfarb M 1995 Requirement of FGF-4 for postimplantation mouse development. Science 267:246–249

    Arman E, Haffner-Krausz R, Chen Y, Heath JK, Lonai P 1998 Targeted disruption of fibroblast growth factor (FGF) receptor 2 suggests a role for FGF signaling in pregastrulation mammalian development. Proc Natl Acad Sci USA 95:5082–5087

    Goldin SN, Papaioannou VE 2003 Paracrine action of FGF4 during periimplantation development maintains trophectoderm and primitive endoderm. Genesis 36:40–47

    Wells JM, Melton DA 2000 Early mouse endoderm is patterned by soluble factors from adjacent germ layers. Development 127:1563–1572

    Horb ME 2000 Patterning the endoderm: the importance of neighbours. Bioessays 22:599–602

    Kimelman D, Griffin KJ 2000 Vertebrate mesendoderm induction and patterning. Curr Opin Genet Dev 10:350–356

    LaBonne C, Whitman M 1997 Localization of MAP kinase activity in early Xenopus embryos: implications for endogenous FGF signaling. Dev Biol 183:9–20

    Henry GL, Brivanlou IH, Kessler DS, Hemmati-Brivanlou A, Melton DA 1996 TGF-? signals and a pattern in Xenopus laevis endodermal development. Development 122:1007–1015

    Gamer LW, Wright CV 1995 Autonomous endodermal determination in Xenopus: regulation of expression of the pancreatic gene XlHbox 8. Dev Biol 171:240–251

    Kavka AI, Green JB 2000 Evidence for dual mechanisms of mesoderm establishment in Xenopus embryos. Dev Dyn 219:77–83

    Hudson C, Clements D, Friday RV, Stott D, Woodland HR 1997 Xsox17 and -? mediate endoderm formation in Xenopus. Cell 91:397–405

    Bertrand V, Hudson C, Caillol D, Popovici C, Lemaire P 2003 Neural tissue in ascidian embryos is induced by FGF9/16/20, acting via a combination of maternal GATA and Ets transcription factors. Cell 115:615–627

    Freeman M 2000 Feedback control of intercellular signalling in development. Nature 408:313–319

    Bhalla US, Iyengar R 1999 Emergent properties of networks of biological signaling pathways. Science 283:381–387

    Niehrs C, Meinhardt H 2002 Modular feedback. Nature 417:35–36

    Dikic I, Giordano S 2003 Negative receptor signalling. Curr Opin Cell Biol 15:128–135

    Tsang M, Friesel R, Kudoh T, Dawid IB 2002 Identification of Sef, a novel modulator of FGF signalling. Nat Cell Biol 4:165–169

    Niehrs C, Pollet N 1999 Synexpression groups in eukaryotes. Nature 402:483–487

    Bernfield M, Gotte M, Park PW, et al 1999 Functions of cell surface heparan sulfate proteoglycans. Annu Rev Biochem 68:729–777

    Nurcombe V, Ford MD, Wildschut JA, Bartlett PF 1993 Developmental regulation of neural response to FGF-1 and FGF-2 by heparan sulfate proteoglycan. Science 260:103–106

    Bonneh-Barkay D, Shlissel M, Berman B, Shaoul E, Admon A, Vlodavsky I, Carey DJ, Asundi VK, Reich-Slotky R, Ron D 1997 Identification of glypican as a dual modulator of the biological activity of fibroblast growth factors. J Biol Chem 272:12415–12421

    Schlessinger J, Plotnikov AN, Ibrahimi OA, Eliseenkova AV, Yeh BK, Yayon A, Linhardt RJ, Mohammadi M 2000 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

    Gospodarowicz D, Cheng J 1986 Heparin protects basic and acidic FGF from inactivation. J Cell Physiol 128:475–484

    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 131:123–130

    Flaumenhaft R, Moscatelli D, Rifkin DB 1990 Heparin and heparan sulfate increase the radius of diffusion and action of basic fibroblast growth factor. J Cell Biol 111:1651–1659

    Walz A, McFarlane S, Brickman YG, Nurcombe V, Bartlett PF, Holt CE 1997 Essential role of heparan sulfates in axon navigation and targeting in the developing visual system. Development 124:2421–2430

    Lin X, Buff EM, Perrimon N, Michelson AM 1999 Heparan sulfate proteoglycans are essential for FGF receptor signaling during Drosophila embryonic development. Development 126:3715–3723

    Garcia-Garcia MJ, Anderson KV 2003 Essential role of glycosaminoglycans in Fgf signaling during mouse gastrulation. Cell 114:727–737

    Lin W, Furthauer M, Thisse B, Thisse C, Jing N, Ang SL 2002 Cloning of the mouse Sef gene and comparative analysis of its expression with Fgf8 and Spry2 during embryogenesis. Mech Dev 113:163–168

    Kovalenko D, Yang X, Nadeau RJ, Harkins LK, Friesel R 2003 Sef inhibits fibroblast growth factor signaling by inhibiting FGFR1 tyrosine phosphorylation and subsequent ERK activation. J Biol Chem 278:14087–14091

    B?ttcher RT, Pollet N, Delius H, Niehrs C 2004 The transmembrane protein XFLRT3 forms a complex with FGF receptors and promotes FGF signalling. Nat Cell Biol 6:38–44

    Saffell JL, Williams EJ, Mason IJ, Walsh FS, Doherty P 1997 Expression of a dominant negative FGF receptor inhibits axonal growth and FGF receptor phosphorylation stimulated by CAMs. Neuron 18:231–242

    Williams EJ, Walsh FS, Doherty P 2003 The FGF receptor uses the endocannabinoid signaling system to couple to an axonal growth response. J Cell Biol 160:481–486

    Cavallaro U, Niedermeyer J, Fuxa M, Christofori G 2001 N-CAM modulates tumour-cell adhesion to matrix by inducing FGF-receptor signalling. Nat Cell Biol 3:650–657

    Van Vactor D, O’Reilly AM, Neel BG 1998 Genetic analysis of protein tyrosine phosphatases. Curr Opin Genet Dev 8:112–126

    Ostman A, Bohmer FD 2001 Regulation of receptor tyrosine kinase signaling by protein tyrosine phosphatases. Trends Cell Biol 11:258–266

    Eblaghie MC, Lunn JS, Dickinson RJ, Munsterberg AE, Sanz-Ezquerro JJ, Farrell ER, Mathers J, Keyse SM, Storey K, Tickle C 2003 Negative feedback regulation of FGF signaling levels by Pyst1/MKP3 in chick embryos. Curr Biol 13:1009–1018

    Kawakami Y, Rodriguez-Leon J, Koth CM, Buscher D, Itoh T, Raya A, Ng JK, Esteban CR, Takahashi S, Henrique D, Schwarz MF, Asahara H, Izpisua Belmonte JC 2003 MKP3 mediates the cellular response to FGF8 signalling in the vertebrate limb. Nat Cell Biol 5:513–519

    Hacohen N, Kramer S, Sutherland D, Hiromi Y, Krasnow MA 1998 sprouty encodes a novel antagonist of FGF signaling that patterns apical branching of the Drosophila airways. Cell 92:253–263

    Lim J, Yusoff P, Wong ES, Chandramouli S, Lao DH, Fong CW, Guy GR 2002 The cysteine-rich sprouty translocation domain targets mitogen-activated protein kinase inhibitory proteins to phosphatidylinositol 4,5-bisphosphate in plasma membranes. Mol Cell Biol 22:7953–7966

    de Maximy AA, Nakatake Y, Moncada S, Itoh N, Thiery JP, Bellusci S 1999 Cloning and expression pattern of a mouse homologue of Drosophila sprouty in the mouse embryo. Mech Dev 81:213–216

    Minowada G, Jarvis LA, Chi CL, Neubuser A, Sun X, Hacohen N, Krasnow MA, Martin GR 1999 Vertebrate sprouty genes are induced by FGF signaling and can cause chondrodysplasia when overexpressed. Development 126:4465–4475

    Chambers D, Mason I 2000 Expression of sprouty2 during early development of the chick embryo is coincident with known sites of FGF signalling. Mech Dev 91:361–364

    Hanafusa H, Torii S, Yasunaga T, Nishida E 2002 Sprouty1 and sprouty2 provide a control mechanism for the Ras/MAPK signalling pathway. Nat Cell Biol 4:850–858

    Sasaki A, Taketomi T, Kato R, Saeki K, Nomani A, Sasaki M, Kuriyama M, Saito N, Shibuya M, Yoshimura A 2003 Mammalian Sprouty4 suppresses Ras-independent ERK activation by binding to Raf1. Nat Cell Biol 5:427–432

    Christofori G 2003 Split personalities: the agonistic antagonist sprouty. Nat Cell Biol 5:377–379

    Tefft JD, Lee M, Smith S, Leinwand M, Zhao J, Bringas Jr P, Crowe DL, Warburton D 1999 Conserved function of mSpry-2, a murine homolog of Drosophila sprouty, which negatively modulates respiratory organogenesis. Curr Biol 9:219–222

    Ong SH, Hadari YR, Gotoh N, Guy GR, Schlessinger J, Lax I 2001 Stimulation of phosphatidylinositol 3-kinase by fibroblast growth factor receptors is mediated by coordinated recruitment of multiple docking proteins. Proc Natl Acad Sci USA 98:6074–6079

    Ryan PJ, Paterno GD, Gillespie LL 1998 Identification of phosphorylated proteins associated with the fibroblast growth factor receptor type I during early Xenopus development. Biochem Biophys Res Commun 244:763–767

    Rodriguez-Viciana P, Warne PH, Dhand R, Vanhaesebroeck B, Gout I, Fry MJ, Waterfield MD, Downward J 1994 Phosphatidylinositol-3-OH kinase as a direct target of Ras. Nature 370:527–532

    Pacold ME, Suire S, Perisic O, Lara-Gonzalez S, Davis CT, Walker EH, Hawkins PT, Stephens L, Eccleston JF, Williams RL 2000 Crystal structure and functional analysis of Ras binding to its effector phosphoinositide 3-kinase . Cell 103:931–943

    Locascio A, Nieto MA 2001 Cell movements during vertebrate development: integrated tissue behaviour versus individual cell migration. Curr Opin Genet Dev 11:464–469

    Beddington RS, Smith JC 1993 Control of vertebrate gastrulation: inducing signals and responding genes. Curr Opin Genet Dev 3:655–661

    Harland R, Gerhart J 1997 Formation and function of Spemann’s organizer. Annu Rev Cell Dev Biol 13:611–667

    Wilson SI, Edlund T 2001 Neural induction: toward a unifying mechanism. Nat Neurosci 4:1161–1168

    Meyers EN, Lewandoski M, Martin GR 1998 An Fgf8 mutant allelic series generated by Cre- and Flp-mediated recombination. Nat Genet 18:136–141

    Yokota C, Kofron M, Zuck M, Houston DW, Isaacs H, Asashima M, Wylie CC, Heasman J 2003 A novel role for a nodal-related protein; Xnr3 regulates convergent extension movements via the FGF receptor. Development 130:2199–2212

    Reifers F, Bohli H, Walsh EC, Crossley PH, Stainier DY, Brand M 1998 Fgf8 is mutated in zebrafish acerebellar (ace) mutants and is required for maintenance of midbrain-hindbrain boundary development and somitogenesis. Development 125:2381–2395

    Draper BW, Stock DW, Kimmel CB 2003 Zebrafish fgf24 functions with fgf8 to promote posterior mesodermal development. Development 130:4639–4654

    Xiong S, Zhao Q, Rong Z, Huang G, Chen P, Zhang S, Liu L, Chang Z 2003 hSef inhibits PC-12 cell differentiation by interfering with Ras-mitogen-activated protein kinase MAPK signaling. J Biol Chem 278:50273–50282

    Yang RB, Ng CK, Wasserman SM, Komuves LG, Gerritsen ME, Topper JN 2003 A novel interleukin-17 receptor-like protein identified in human umbilical vein endothelial cells antagonizes basic fibroblast growth factor-induced signaling. J Biol Chem 278:33232–33238

    Preger E, Ziv I, Shabtay A, Sher I, Tsang M, Dawid IB, Altuvia Y, Ron D 2004 Alternative splicing generates an isoform of the human Sef gene with altered subcellular localization and specificity. Proc Natl Acad Sci USA 101:1229–1234

    Torii S, Kusakabe M, Yamamoto T, Maekawa M, Nishida E 2004 Sef is a spatial regulator for Ras/MAP kinase signaling. Dev Cell 7:33–44

    Tsuji L, Yamashita T, Kubo T, Madura T, Tanaka H, Hosokawa K, Tohyama M 2004 FLRT3, a cell surface molecule containing LRR repeats and a FNIII domain, promotes neurite outgrowth. Biochem Biophys Res Commun 313:1086–1091

    Walicke P, Cowan WM, Ueno N, Baird A, Guillemin R 1986 Fibroblast growth factor promotes survival of dissociated hippocampal neurons and enhances neurite extension. Proc Natl Acad Sci USA 83:3012–3016

    Williams EJ, Furness J, Walsh FS, Doherty P 1994 Activation of the FGF receptor underlies neurite outgrowth stimulated by L1, N-CAM, and N-cadherin. Neuron 13:583–594

    Lin HY, Xu J, Ornitz DM, Halegoua S, Hayman MJ 1996 The fibroblast growth factor receptor-1 is necessary for the induction of neurite outgrowth in PC12 cells by aFGF. J Neurosci 16:4579–4587(Ralph T. B?ttcher and Christof Niehrs)