Open access peer-reviewed chapter

Novel Cellular and Molecular Interactions During Limb Development, Revealed from Studies on the Split Hand Foot Congenital Malformation

Written By

Daniele Conte, Luisa Guerrini and Giorgio R. Merlo

Submitted: 29 September 2014 Reviewed: 12 February 2015 Published: 21 October 2015

DOI: 10.5772/60402

From the Edited Volume

New Discoveries in Embryology

Edited by Bin Wu

Chapter metrics overview

1,946 Chapter Downloads

View Full Metrics

Abstract

The embryonic development of the limbs is widely used as a paradigm for the comprehension of the cellular processes and molecular mechanisms underlying organogenesis and pattern formation. The chick, mouse and (recently), zebrafish embryos are excellent models, for the ease of experimental manipulation and the availability of several mutant strains with limb malformation defects.

Keywords

  • Limb
  • Embryonic ectoderm
  • AER
  • SHFM
  • EEC
  • p63
  • Dlx5
  • Wnt5a
  • FGF8

1. Introduction

The limbs are projecting paired appendages of an animal body used especially for movement and grasping, for example, wings, arms, and legs. The development of the limb bud is often taken as a paradigm for a cellular and molecular comprehension of the common principles of organogenesis and pattern formation. Embryonic patterning implies that cells acquire positional information, usually by interpreting concentration gradient of signalling molecules. Accordingly, limb pattern is specified along three principal axes: anterior-posterior (A-P) (e.g., thumb to little finger), dorsal-ventral (D-V) (e.g., back of hand to palm) and proximal-distal (P-D) (e.g., shoulder to nails). Digit pattern across the A-P axis is a classic example of a signalling gradient that specifies positional values, linked to a gradient of Sonic-Hedgehog (SHH). D-V patterning is less studied and involves signals from dorsal and ventral ectoderm. The specification of P-D positional values has long been considered to involve a timing mechanism, under the control of ligands of the fibroblast growth factor (FGF) family. A concentration gradient of molecules can also give cells polarity information, recently shown to be critical for patterning and morphogenesis.

The limbs are not essential for life, thus a large number of mutant strains are available for studies on the genetic determinants of limb development, in normal and pathological conditions. Manipulation of chicken limb buds has been widely used in the past, mainly because of the ease of examination and manipulation, to postulate the first models of limb bud development leading to the identification of important regulatory genes and interactions. In addition to the chicken model, functional genetics has made great advances thanks to spontaneous and engineered loss- and gain-of-function mutant mouse strains, and recently with the advent of the zebrafish embryos as animal models.

In this chapter, we illustrate the pathways centred on the master transcription factor p63, and discuss the mechanisms by which these pathways impact on the regulation of signalling molecules controlling growth and patterning of the normal limb bud. Based on available knowledge, we propose how signalling networks are misregulated in the split hand foot malformation (SHFM) and related developmental conditions, and indicate emerging functions of the FGF8 and Wnt5a diffusible molecules.

1.1. Limb initiation

Around the embryonic age E8.0 in the mouse, limb buds are initiated as four lumps of mesenchymal cells covered by ectoderm, protruding from the main body axis at approximately the position of somites 6–11 (the forelimbs, FL) and somites 24–27 (the hindlimbs, HL). The limb buds are paired along the cephalo-caudal axis and develop at the same fixed locations on this body axis (Figure 1A). How are their positions defined?

Figure 1.

Schematic representation of limb development with embryonic timeline for chick wing and mouse forelimb. A) Representation of the prospective limb territories in a stage 14/8 chick (Hamburger-Hamilton stages, HH))/mouse embryo. The forelimbs (FL) and hindlimbs (HL) derive from discrete regions of the lateral plate mesoderm (LPM). Tbx5 (red) is expressed in the prospective FL, whereas Tbx4 (red) is expressed in the prospective HL. In this stage, Pitx1 (yellow) is expressed in a caudal domain that overlaps with Tbx4. The somites (blue) are numbered and serve as reference for the axial position of the FL and HL fields. B) Model of initiation of FL bud. Hox protein gradients establish the condition for the synthesis of retinoic acid (RA) in the LPM. RA causes induction of the transcription factors Tbx5 (or Tbx4 for the HL). Tbx5-expressing mesechymal cells express FGF10 and induce the ectodermal cells of the surface (yellow square) to activate epithelial-mesenchymal transition (EMT). C) Newly generated mesenchymal cells express FGF10 that induces the overlaying ectodermal cells to express FGF8 giving rise to the apical ectodermal ridge (AER). Expression of FGF8 by the AER induces the mesenchymal cells to express FGF10, thus establishing a positive feedback loop for the initial phases of limb outgrowth. D) Schematic representation of the progress zone (PZ) model. The AER maintains cells of the PZ in an unspecified state. For a detailed description of the proposed models of P-D patterning see the text. E) AER-derived FGF8 maintain the expression of SHH in the ZPA cells, which in turn gives feedback on the AER cells to maintain FGFs expression, via Grem1 and BMP inhibition. This signalling between the AER and ZPA contributes to co ordinate the patterning along the P-D and A-P axes.

It has been proposed that the position of several dorsal organs along the cephalo-caudal axis, their identity and timing of appearance depend on the expression of specific sets of Hox genes. The 39 vertebrate Hox genes code for homeodomain transcription factors, homologous with the genes of the Drosophila HOM-C complex, and are combinatorially expressed along the main body axis. Hox genes are serially organized in four clusters (a, b, c, and d), each located on a different chromosome. Within each cluster, Hox genes are organized in a physical order collinear with the cephalo-caudal axis of the growing organism so that the genes lying at the 3’ end are expressed earlier and are localized in the most anterior domains. Moving toward the 5’ direction, each next gene is expressed in a progressively more posterior territory. Thus, each Hox gene has a specific anterior limit of expression, and each A-P embryonic territory expresses a specific combination of Hox genes, utilized as positional information.

A key signalling molecule for limb initiation is FGF10, a member of the FGF family of diffusible peptides. The FGF10 mRNA is detected quite early in the presumptive limb mesenchyme and promotes AER induction (a key organizer and regulator of the P-D limb extension; see below) and initiation of ectopic limb development when applied exogenously [1]. Conversely, in FGF10-null mice, limb buds are initiated but the AER does not form, resulting in complete truncation of all four limbs [2 - 4].

The expression of FGF10 coincides with the time when the trunk is only competent to form an ectopic limb, for example, the time at which the trunk mesenchyme becomes determined and can no longer be redirected to a limb fate (the HH stage 16–17 of the chick embryo) [5, 6]. The initial assumption was that limbs originate from a pre-existing mesenchymal population undergoing a localized regulation of proliferation. In fact, at the HH stage 17–18, there is a substantial decrease in proliferation of the flank mesoderm, while instead higher rates are maintained within the emerging limb buds.

The current model considers that, shortly after gastrulation, a re-epithelization of mesodermal cells occurs so that the entire embryo is essentially epithelial, including also the notochord, the somites, the intermediate mesoderm and the lateral plate mesoderm (LPM). At stage HH 13 in the chick, before limb initiation, the somatopleure displays epithelial rather than mesenchymal features. The LPM of the limb field starts out as an epithelium and ultimately generates limb-bud mesenchyme through a process termed epithelial-mesenchymal transition (EMT) [7] (Figure 1B). In embryos null for FGF10 and Tbx5 the proportion of mesenchymal cells compared to the proportion of epithelial cells was significantly lower than that of wild-type (WT). These mutants show hyperplasia of the somatopleure epithelium, in support of failure of these cells to undergo EMT. These new data show the time in which the trunk is competent to form an ectopic limb, is precisely the time at which the trunk mesenchyme is initially generated. The old experiments of ectopic application of FGF10 need to be re-interpreted, as induction of limb-bud formation from epithelial trunk somatopleure cells and not from mesenchymal cells of the same A-P level.

1.2. T-box genes and limb-type specification

The FL and HL of all vertebrate species are evidently different (e.g., wing vs. leg in the chick embryo, pectoral vs. pelvic fins in fish embryos, arms vs. legs in primates, etc.). The specification of limb-type identity and morphology is established before overt limb initiation. A large body of evidence indicates that two transcription factors of the T-box family participate in the early definition of limb-type identity: Tbx5 for the presumptive territories of the mouse FL (wing of the chick, pectoral fins of fishes) and of Tbx4 in the presumptive territory of the mouse HL (leg of the chick, pelvic fins of fishes) (Figure 1A). These genes are expressed very early in the prospective limb mesoderm and, in addition to define limb-type identity, also appear to be necessary and sufficient for early limb induction [8 - 15].

However, although only expressed in FL and HL, respectively, Tbx5 and Tbx4 appear not to be the master directors of limb-bud identity/morphology design. Instead, the Pitx1 gene, which codes for a paired-type homeodomain transcription factor and which is expressed in the HL bud, is the upstream regulator of Tbx4, and is directly implicated in HL specification. Multiple independent cis-regulatory elements of Tbx4 expression have been identified, including the HL-specific enhancers [16]. Both may be targets of Pitx1 and other unknown upstream factors. Structural changes of these regulators might be some of the multiple factors responsible for the hind/lower limb morphology specification. However, limb-bud identity/morphology determination remains to a large extent unexplained.

1.3. Role of retinoic acid

A signalling molecule known to act upstream of Tbx5 during forelimb/pectoral fin development is retinoic acid (RA) [17 - 21]. According to old observations, patterning along the proximal-distal axis of the vertebrate limb is controlled by opposing diffusible signals, in which RA functions as the proximal signal and FGF as the distal one [22, 23].

The mechanism through which RA controls limb development has been widely debated [21], but clear results have only been produced in recent years [17, 19, 22, 24 - 27]. Mouse and zebrafish embryos null for the gene aldehyde dehydrogenase-1a2 (aldh1a2) fail to synthesize RA and do not develop limbs/fins. These mutants fail to express Tbx5, and the exogenous application of RA can rescue both limb/fin development and Tbx5 expression [18, 19, 25, 27, 28]. However, a recent paper shows that RA signalling is not required for P-D limb patterning, and instead provides genetic evidence that RA-FGF antagonism occur only along the trunk lateral plate mesoderm, prior to FL budding, to permit induction of Tbx5 [29]. This study shows that RA controls limb development in a manner much different than that originally envisioned (see below).

1.4. Proximal-distal axis

The limb skeleton is laid down as five cartilage skeletal elements, not just the three referred to as stylopod (humerus/femur), zeugopod (radius-ulna/tibia-fibula), and autopod (digits); in fact two carpal regions between zeugopod and autopod are present, that initially have the same size as the other segments but then grow substantially less.

P-D extension and patterning is strictly linked to the signalling activity of the apical ectodermal ridge (AER), a morphologically distinct ectodermal thickening, extending along the entire A-P length, and lining the D-V border. The AER is present between E9 and E11 in the mouse embryo, consists of a pseudo-stratified epithelium in the chick and pluri-stratified epithelium in the mouse, and is a dynamic structure constantly undergoing morphogenetic changes [30 - 32].

The AER plays a fundamental role in promoting and regulating the outgrowth and patterning of the P-D limb axis. Experimental removal of the AER in chicken limb buds, causes a developmental arrest, and truncation of wing skeleton [33], meanwhile grafts of an AER to a recipient limb bud induces ectopic P-D outgrowth [34]. In 1993, Niswander identified FGFs as the relevant signals produced by the AER to induce P-D limb axis formation and extension. P-D extension and outgrowth is rescued by exogenous application of FGFs on AER removal [35]. This study provided the first molecular insights into how AER-FGF signalling controls in P–D extension and patterning. Four FGF ligands (4, -8, -9 and -17), are expressed by the AER cells with redundant functions during P–D patterning of mouse limb buds. Inactivation of the three FGFs expressed predominantly by the posterior AER (FGF4, -9, -17) does not alter limb-bud development [36]. In contrast, loss of FGF8, which is the first and only FGF ligand expressed by the entire AER from early stages onward, disrupts formation of the proximal-most limb skeletal element [37 - 40]. This early and transient disruption of P-D extension is rescued by the activation of FGF4 in the FGF8-deficient AER, which results in almost normal development of the more distal limb skeleton [41, 42]. Combined inactivation of both FGF8 and FGF4 causes a complete arrest of limb-bud development and limb agenesis [39, 42]. In addition, transient expression of FGF8 and FGF4 during initiation of limb-bud outgrowth is sufficient for specification of the entire PD axis, but the progressive proliferative expansion of such specified limb segments is disrupted [42].

Other AER-expressed FGFs, in particular FGF9, contribute to the proliferative expansion of the limb mesenchymal progenitors in a P-D sequential order, so that higher AER-expressed FGF levels are required for formation of more distal limb structures [36]. Taken together, this genetic analysis reveals an instructive role of AER-FGF signalling in the specification and proliferative expansion along the PD axis.

The AER is first induced by the expression of FGF10 in the prospective limb-bud mesenchyme. FGF10 is expressed in the same territories as Tbx4 and Tbx5 and interestingly, FGF10 is a direct transcriptional target of these transcription factors. The expression of FGF10 is essential to establish AER-expressed FGF signalling, which in turn is required to maintain Ffg10 expression [1, 42]. The reciprocal induction of FGF10-FGF8 requires the expression of Wnt3a, coding for a ligand of the Wnt family acting through the β-catenin pathway (described in the following sections). The activation of Wnt3 expression couples FGF8 and FGF4 expression from cells of the AER with FGF10 expression [43, 44]. Thus, in these early phases, a positive feedback loop between FGF10 and FGF8 is established in adjacent territories, and is required for reciprocal maintenance (Figure 1A-C).

1.5. Limb extension: The progress zone

Old experiments showed that removing the AER at progressively earlier stages resulted in truncations of the limb skeleton at progressively more proximal levels [33]. Thus, the acquisition of a P-D positional identity seemed to depend on the time that proliferating/unspecified cells spend near the AER (the progress zone, PZ) under the influence of AER signals. According to the model proposed by Summerbell and Wolpert [45] the mesenchymal progenitor cells leaving the PZ early would acquire proximal identities, whereas the same cells leaving the PZ later would acquire progressively more distal identities (Figure 1D).

The great merit of this model has been to introduce the notion of time as an important factor in morphogenetic signalling; however, as a result of extensive molecular and cellular analyses, the original PZ model has been largely abandoned. First, the loss of proximal but not distal skeletal elements in FGF8-deficient mouse limb buds [40] are difficult to reconcile. Second, fate mapping studies in chicken embryos provide good evidence for the presence of pools of progenitor cells with distinct P-D identities, specified very early and then expanded sequentially by proliferation [46]. Removal of the AER at progressively later stages simply eliminates the distal mesenchyme containing the specified but not yet expanded progenitor pools. These studies provide a straightforward alternative explanation for the loss of distal skeletal elements following AER removal [46]. These and other results led to the proposal of the early specification/expansion model as an alternative to the PZ model [42, 46]. This model proposes that AER-expressed FGF signalling controls survival and sequential proliferative expansion of a pool of progenitor/stem cells in P-D territories, in a dose- and time-dependent highly regulated fashion.

A third model has been proposed, in which P-D patterning is controlled by opposing diffusible signals, with RA functioning as a proximal signal and FGF acting as a distal signal [26]. Chick FL or HL ectopically exposed to RA or FGF8, or to antagonists of RAR or FGF receptor, display P-D fate changes that either expand or contract expression of proximal limb markers [23]. Further evidence has been recently provided, indicating that RA is needed for P–D patterning of both FL and HL [47, 48]. Using recombinant heterotopic chick limb transplantations they propose that the exposure to the activities of Wnt3a, FGF8 (distal molecule), and RA (proximal molecule) maintains the potential to form both proximal and distal structures. While these studies report the ability of RA treatment to reprogram distal limb mesenchyme to a proximal fate and to maintain early limb mesenchyme in a Meis1-expressing proximal fate alongside Wnt and FGF treatment [47], they do not address a requirement for endogenous RA in proximal limb mesenchyme. The ability of RA to increase Meis1/2 (a proximal marker) could be a consequence of loss of FGF signalling, known to repress Meis1/2 distally [36].

Recently, Cunningham and colleagues [29] provide convincing evidence that RA is not required for limb patterning and that RA-FGF antagonism does not occur along the limb P-D axis, as originally proposed [26]. They suggest that both the initial expression of Meis1/2 in the LPM and later in the proximal limb bud do not require RA signalling, while the downregulation of Meis1/2 expression in the distal limb requires AER-derived FGF8. They suggest that since FGF8 expression in the AER appears after limb-bud formation [37], the proximal most limb domain is out of range of early AER-derived FGF signals, leading to maintenance of proximal Meis1/2 expression and restricted distal Meis1/2 expression (Figure 1D).

1.6. Anterior-posterior pattern

The mammalian limb bud is typically pentadactylous, for example, the autopodium gives rise to five skeletal elements. The digit organization, from anterior (pre-axial, the thumbs) to posterior (post-axial, the little finger) is referred to as the A-P pattern. It has long been recognized that the embryonic tissue mainly implicated in the regulation of the A-P pattern is the zone of polarizing activity (ZPA) (Figure 1E). In 1956, a region within the posterior-proximal limb mesenchyme was identified, that when grafted in the anterior margin of host chicken wing buds results in mirror image duplications of all digits [49, 50]. The ZPA acts as a signalling centre and specifies positional information in the limb-bud mesenchyme by secreting the diffusible molecule Sonic Hedgehog (SHH). Within the limb bud mesenchyme, SHH is present in a posterior (high) to anterior (low) gradient [51, 52]. Genetic studies indicate that the time spent expressing SHH provides cells with a kinetic memory relevant to specification of their A-P identities [53 - 56].

SHH signalling is translated into an intracellular, anterior (high) to posterior (low), gradient of the transcriptional repressor Gli3R within the limb mesenchyme [67]. Upon binding to the receptor Patched, SHH counteracts the conversion of Gli3 full-length into its cleaved repressor form. The Gli3R gradient is then required to establish the polarized expression of other genes involved in A-P patterning, and ultimately is translated into digit pattern, in ways not fully clarified [24, 57 - 61].

Further genetic studies in mouse and zebrafish embryos have implicated also HAND2 in the activation of SHH expression in both limb and fin buds [62]. Moreover, in the mouse embryo a mutual antagonistic interaction between Hand2 and Gli3, prior to SHH expression, establishes a A-P pre-pattern [60, 61, 63]. Finally, at later stages of development, the expression of the 5’ most members of the Hoxd gene cluster is activated within the posterior limb-bud mesenchyme. Cell biochemical studies have revealed a direct interaction of Hoxd proteins with the cis-regulatory limb-bud enhancer region of the SHH gene [64].

1.7. Hox genes and digit identity

An exhaustive illustration of this topic is beyond the scope of this chapter. Digit patterning has commonly been interpreted in the context of a gradient of expression of SHH preventing the processing of Gli3 to its repressor form (Gli3R) [65, 66].

Thus, a SHH gradient is translated into an inverse Gli3R gradient [24, 67]. However, between Gli3 and SHH;Gli3 null mutant mice display identical polydactylous limb phenotypes, indicating that an iterative series of (non-patterned) digits can form in the absence of SHH [24, 60], suggesting the existence of a SHH-independent prepatterning.

This observation, rather than supporting the SHH gradient model, is consistent with a Turing-type model of digit patterning [68 - 70]. According to this model, dynamic interactions between activator and inhibitor molecules produce periodic patterns of spots or stripes, serving as a molecular pre-pattern for chondrogenesis. Although the core molecules of a self-organizing mechanism remain poorly known, potential candidates for molecular modulators of the system include the Hox genes [70]. Distal Hoxa and Hoxd genes have a well-known impact on digit number, though their specific role remains unclear. Hoxd genes interact with the SHH-Gli3 pathway; these include the mutual transcriptional regulation between Hox genes and SHH and the binding of Hoxd12 to Gli3R, resulting in a blockage of Gli3R repressor activity [71 - 73]. In general, gain- and loss-of-function experiments suggest a positive relation between Hox genes and digit number [72, 74] that is also indicated by the ectopic anterior up-regulation of distal Hoxd genes in the Gli3-dependent polydactyly [24, 61]. Interestingly the disruption of various Hox genes combined with loss of Gli3 results in polydactyly; more Hox genes are removed – more digits are formed [75]. Thus, losing Hox genes seemed to shorten the spacing between digits – the wavelength in Turing’s mathematical language.

The Turing’s model implies the activity of two diffusing and interacting molecules; however, Hox genes code for non-diffusible transcription factors, for example, they cannot directly participate. However, evidence that distal Hox genes are necessary for correct limb development is overwhelming. Indeed, in addition to a correct digit formation via a Turing-type regulation of SHH signalling, a second role of Hox genes in limb P-D patterning has been studied in depth. Caudal, late-expressed paralogs of the Hox gene clusters display a P-D as well as A-P gradient of expression within the limb mesenchyme [76, 77] suggesting a combinatorial role of these genes in patterning the limb skeletal elements. Experimental evidence leads to the conclusion that the paralogs 9–13 of the Hox gene clusters -a and -d specify individual limb segments [78]. Indeed Hoxa11 -/-;Hoxd11 -/- double mutant embryos lack radius and ulna [79] while Hoxa13 -/-;Hoxd13 -/- double mutants lack digits [80].

Finally, in spite of the major role played by posterior Hox genes, little is known about the cellular and/or molecular bases for the observed developmental defects. Attempts in this directions [81] show that malformation of the FL zeugopod in Hoxa11/Hoxd11 double mutant mice results from multiple defects during the formation of the zeugopod, including reduced FGF8 and FGF10 expression, formation of smaller mesenchymal condensations, and failure to form normal growth plates at the proximal and distal ends of the zeugopod bones. As a consequence, growth and maturation of these bones is highly disorganized.

1.8. AER and ZPA interaction

The maintenance and propagation of SHH expression requires AER-derived FGF signalling as part of a positive epithelial-mesenchymal (E-M) feedback loop operating between the ZPA and the AER [82, 83] (Figure 1E). The BMP antagonist Gremlin1 (Grem1) was identified as a crucial mesenchymal component in this E-M feedback signalling system [59, 66, 84]. Grem1 is required to up-regulate AER-FGF signalling and to establish SHH/Grem1/FGF E-M feedback signalling. In Grem1-null limb buds, the establishment of this E-M feedback signalling loop is interrupted, and this in turn interferes with specification and expansion of the distal compartment (zeugopod and autopod) [59, 84].

1.9. Dorsal-ventral axis

Dorsal-ventral (D-V) patterning is mainly organized via signalling by Wnt7a, a diffusible molecule of the Wnt-family expressed in the dorsal ectoderm. Wnt7a is both necessary and sufficient to dorsalize the limb, indeed the loss of Wnt7a causes the dorsal side of limbs to acquire a ventral side identity, accompanied by missing posterior digits. Wnt7a is required to maintain expression of SHH, explaining the digit loss. Restoring the Wnt7a signal rescues both of these defects [85, 86].

Wnt7a induces the expression of Lmx1, coding for a Lim-family homeodomain-containing transcription factor. Lmx-1 is involved in dorsalization of the limb, which was shown by deleting Lmx1 in mice: Lmx1 null embryos produced ventral skin on both sides of their paws [87, 88]. There are other factors known to control the D-V patterning; on the ventral side the transcription factor-coding Engrailed-1 gene, exclusively expressed in the ventral ectoderm, has been shown to repress the dorsalizing effect of Wnt7a in this territory [89] (Figure 1E).

Advertisement

2. Distal Limb Malformations in Human

Congenital limb malformations are relatively common, and are genetically and clinically heterogeneous, with a diverse spectrum in their epidemiology, aetiology and anatomy. They are often difficult to diagnose and categorize, because of their complex phenotypes and their association with other malformations and clinical symptoms. Many etiological factors have been suggested for limb anomalies, including inheritance of mutated genes, teratogenic drugs, environmental chemicals, ionizing radiation (atomic weapons, radioiodine and radiation therapy), infections, metabolic imbalance (e.g., maternal diabetes), or mechanical factors like amniotic band syndrome. With the advent of functional genetics, molecular pathways centred on disease genes are being unravelled.

A wide set of human congenital limb malformations can be attributed to defects in P-D development. In this chapter, we will attempt to link known disease-causing genes with their known or presumed function in the maintenance of the AER. We will focus on the genes for which more functional data are available: namely Dlx5, FGF8, p63 and Wnts. These genes are co-expressed in the AER cells of the mouse limb [90] as well as in the fins of the zebrafish embryos [91, 92] and are known or proposed diseases-genes for the SHFM and EEC congenital limb malformations.

P-D defect refer to absence or hypoplasia of distal structure of the limb with more or less normal proximal structures. The spectrum of P-D limb reduction anomalies ranges from very mild disorders, such as syndactyly, to very severe forms, such as phocomelia or amelia. The most frequent congenital limb malformations are syndactylies, characterized by the fusion of the soft tissues of fingers and toes with or without bone fusion. Syndactylies are due to the lack of apoptosis in the interdigital mesenchyme and may also occur isolated or with other symptoms in a syndrome [93].

Polydactylies are distinguished by the appearance of supernumerary digits or parts of them, which may be present as a complete duplication of a whole limb or as a duplication of single digits [94]. Pre-axial polydactyly with extra digits located on the side of the hand or of the thumb or postaxial polydactyly where the extra digit is found on the side of the hand or foot of the fifth digit are common isolated limb malformation traits. On molecular level, many forms of polydactyly have been shown to be more or less directly linked to the SHH signal transduction pathway, which play a major role in A-P patterning of the limb [95, 96].

Brachydactylies are defined by shortened digits and are classified on an anatomic and genetic background into five groups from A to E [93]. Isolated brachydactylies are often inherited in an autosomal dominant manner and are characterized by a high degree of phenotypic variability. Type-B brachydactylies are associated to mutation in the Ror2 gene, and Ror2 mutations are also associated with the Robinow syndrome in which brachydactyly is a common feature [97 - 101].

A severe P-D arrest of the developing limb bud gives rise to phocomelia, characterized by undeveloped limbs [102]. Usually the upper limbs are not fully formed and sections of the “hands and arms may be missing”. Short arm bones, fused fingers and missing thumbs will often occur. Legs and feet are also affected. Individuals with phocomelia will often lack thigh bones, and the hands or feet may be abnormally small or appear as stumps due to their close attachment to the body. Phocomelia is a known negative effect of the administration of thalidomide to pregnant women, in use in the late 1950s/early 1960s, to treat morning sickness, although the mechanism of action of this teratogen remains controversial [103, 104].

Failure of formation of limb buds gives rise to amelia, the complete absence of one or more limbs. The most severe form of amelia is the tetra-amelia, characterized by the absence of all four limbs, associated with craniofacial, pulmonary and urogenital defects. This autosomal recessive disorder has been linked to mutations of the WNT3 gene (see the following section).

2.1. SHFM and EEC

SHFM, also known as ectrodactyly or lobster-claw malformation, is a congenital defect affecting predominantly the central rays of hands and/or feet. It may manifest either as an isolated trait or as part of syndromic conditions comprising other developmental disorders [105]. SHFM occurs with the incidence of about 1 in 18,000 live born infants and accounts for 8–17% of all limb malformations [106, 107]. SHFM is clinically heterogeneous, ranging from a relatively mild defect, such as hypoplasia of a single phalanx or syndactyly, to aplasia of one or more central digits (i.e., classical cleft, also known as lobster-claw anomaly).

Inter-individual and intra-familial variability of the SHFM is very high. Furthermore, variable expressivity of this feature can be so significant, that a different pattern of anomaly is seen in each limb of the same individual patient [93]. SHFM is mostly sporadic, although familial forms are known: in these cases an autosomal dominant transmission with reduced penetrance is the most common mode, but autosomal recessive and X-linked forms have been reported.

SHFM has been linked to (at least) six distinct loci [106] (Table 1). SHFM-I (MIM #183600) is the most frequent type and is linked to mutations and/or deletions/rearrangements of the DLX5;DLX6 bigenic locus. Deletions, inversions and rearrangements affecting chromosome 7q21 have long been reported [108 - 112]. The smallest region of overlapping deletions encompasses several other genes in addition to DLX5, and DLX6: as DYNC1l1, SLC25A13, DSS1, but only Dlx5 and Dlx6 have been shown to be specifically expressed in the AER of the developing mouse limbs [113 - 115]. Recently, a point mutation in the DNA-binding domain of DLX5 (Q178P) has been reported in a SHFM-I family with a recessive transmission, co-segregating with the limb malformations [116]. In the mouse, the combined disruption of Dlx5;Dlx6 leads to an ectrodactyly phenotype affecting the HLs [114, 115], fully confirming that the human orthologs DLX5 (and presumably DLX6) are the disease genes for this malformation. Interestingly, SHFM type-V (MIM #606708) is linked to deletions of a region on chromosome 2 encompassing the HOXD gene cluster, near DLX1 and DLX2 [117 - 119]. Although no clear evidence for the involvement of DLX1 and DLX2 in this malformation is available, it is tempting to imagine that similar to DLX5, misregulated expression of DLX1/2 in the human embryonic limb bud could be the molecular mechanism leading to SHFM type-V.

SHFM type-III (MIM# 600095) is associated with duplications/rearrangements around the DACTYLIN (FBXW4) locus on chromosome 10q and the synthetic one in mice [120, 121]. The genomic lesion involves the DACTYLIN, LBX1, βTRCP and other more distant genes, but none of these is directly disrupted by the rearrangement and no point mutation has been reported. Interestingly, the FGF8 locus is located in the proximity of the rearrangement breakpoints [122], and considering its importance for P-D limb development, it represents a valid candidate SHFM type-III disease gene.

Mutations of p63 are associated to SHFM type-IV (MIM #605289), a condition in which ectrodactyly appears as an isolated non-syndromic disorder linked to mutations or chromosomal anomalies in the DBD or in the C-terminal domain of p63α [123 - 125]. The α tail of p63 contains a sumoylation site, inactivated by p63 mutations found in SHFM-IV (E639X). Sumoylation can modulate p63 half-life [126] and naturally occurring mutated p63 proteins often display altered stability, suggesting that the final effect of the mutations could be the persistence of the mutated protein and consequent misexpression of p63 targets [125, 127, 128]. p63 mutations also cause the ectodermal dysplasia-ectrodactyly-cleft lip/palate syndrome type-III (EEC-3) syndrome (MIM #604292) [129] in which ectrodactyly is a common feature.

SHFM type-VI (MIM #225300) is the only autosomal recessive form of this malformation, and is due to homozygous point mutations of the WNT10B gene [130 - 133]. Finally, the X-linked SHFM type-II form (MIM #313350) has been mapped to chromosome Xq26.3 [134] but no disease gene has yet been identified.

SHFM locus Chromosome/gene affected Case reported Inheritancepattern Limb phenotypes Additional phenotypes References
SHFM-I Rearrangements 7q21.3-q22.1DLX5 mutation 1 family Autosomal dominant
Autosomal recessive
SHFM EEC, mental retardation, sensorineural deafness Crackower et al. (1996)
Marinoni et al. (1995)
Shamseldin et al. (2012)
SHFM-II Xq26 1 family X-linked recessive SHFM, syndactyly, metacarpalhypoplasia, phalangeal hypoplasia Faiyaz ul
Haque et al. (2005)
SHFM-III Duplication 10q24 20% Autosomal dominant SHFM, triphalangeal and/
orduplicated thumbs
de Mollerat et al. (2003)
SHFM-IV TP63 mutations 10% non syndromic93% EEC syndromes Autosomal dominant SHFM EEC, ADULT, LADD, CHARGE,VATER/mental retardation van Bokhoven et al. (2001)
Ianakiev et al. (2000)
SHFM-V Deletion 2q31 Autosomal dominant SHFM Mental retardation, ectodermal and craniofacialfindings, orofacial clefting Goodman et al. (2002)
Del Campo et al. (1999)
SHFM-VI WNT10B mutation 3 family1 sporadic case Autosomal recessive SHFM, tibial aplasia/
hypoplasia
Ugur and Tolun (2008)
Blattner et al. (2012)
Khan et al. (2012)

Table 1.

Genetic alterations and SHFM-related phenotypes

SHFM – Split Hand/Foot Malformation, EEC – Ectrodactyly-Ectodermal dysplasia-Cleft lip/palate, ADULT – Acro-Dermato-Ungual-Lacrimal-Tooth syndrome, LADD – Lacrimo-Auriculo-Dento-Digital syndrome,


CHARGE syndrome (Coloboma of the eye, Heart defects, atresia of the nasal choanae, Retardation of growth and/or development, Genital and/or urinary abnormalities, Ear abnormalities and deafness), VATER association - vertebral anomalies, anal atresia, cardiovascular anomalies, tracheoesophageal fistula, renal and/or radial anomalies, limb defects.


2.2. p63-Dlx5;Dlx6 Regulation

SHFM type-IV and EEC are caused by mutations in the p63 gene, which codes for a highly conserved transcription factor related to the p53 and p73 tumour-suppressor genes [129, 135 - 137]. A common feature of these disorders is ectodermal dysplasia, consisting in abnormal maturation and stratification of the skin and abnormal development of hairs, teeth, nails, exocrine glands and cornea. The other two consistent features of p63-linked disorders are cleft lip/palate and ectrodactyly.

p63 is expressed in the basal or progenitor layers of many epithelial tissues [138, 139], and is able to promote the epithelial stratification program typical of the mammalian skin, as well as to control proliferation and exit from the cell cycle of epidermal stem cells. For these activities p63 has been proposed as a master regulator of epidermal stem cell maintenance, proliferation and stratification [140]. The TP63 gene is translated into ten protein isoforms [141]: the trans-activating (TA) isoforms, closely resembling p53, and the delta-N (ΔN) isoforms, devoid of the TA-domain-1 (TA1). Although the TA isoforms were initially thought to be the ones to possess transcriptional regulatory functions, it has well been established that the ΔN isoforms can activate transcription of a distinct set of target genes via a second TA-domain-2 (TA2) [142]. Five TA and ΔN isoforms are generated by two transcripts which are subjected to alternative splicing, thus the final protein products differ at the carboxyl termini (α, β, γ, δ and ε). In addition to TA1 or TA2 domain, the p63 proteins contain a DNA-binding domain (DBD) and an oligomerization domain (OD). The α-isoforms (either TA or ∆N) also contain a sterile alpha motif (SAM) domain, a protein-protein interaction module found in developmentally relevant proteins [143, 144]. Recent studies have identified a transcriptional inhibitory (TI) domain located between the SAM domain and the C-terminus of p63α; this domain is believed to be responsible for the lower transcriptional activity of TAp63α compared to the -β and the -γ isoforms [145]. ∆Np63α is the most expressed isoform in the embryonic ectoderm.

Attempts to establish genotype–phenotype correlations are hampered by the variable clinical expressivity observed within families: SHFM type-IV and the EEC syndromes are due to mutations in the DNA-binding domain of p63 [129]. In these cases, all p63 isoforms are affected by the mutations. DBD mutants usually act as dominant-negative effectors and render the WT protein unable to bind DNA [129], explaining the dominant transmission of EEC. In contrast, the Hay Wells or ankyloblepharon-ectodermal dysplasia-cleft palate syndrome (AEC, MIM #106260) manifests with normal limbs but severe skin defects, and is typically associated with heterozygous missense mutations in the SAM domain of p63. The acro-dermato-ungual lacrimal tooth (ADULT, MIM #103285) syndrome is associated with a specific gain-of-function mutation R298Q/G in exon 8, affecting the DNA-binding domain of p63. Finally, both limb-mammary syndrome (LMS, MIM #603543), very similar to ADULT and EEC syndromes, and Rapp-Hodgkin syndrome (RHS, MIM #129400), resembling AEC, are due to p63 mutations. p63 mutations causing EEC are usually not found in AEC, LMS and SHFM [146 - 148].

Mice null for p63 have been generated by two groups independently [136, 137]; at birth these mice show severe defects affecting their skin, limb and craniofacial skeleton, teeth, hair, and mammary glands. Specifically, the skin appears thin, mostly single layered and translucent, unable to prevent water loss. The HLs fail to form altogether, while the FLs are severely truncated and lack most of their distal skeletal elements. The altered phenotypes observed in these mutant mice are a direct consequence of altered cellular properties affecting the same tissues and organs as in EEC patients [90, 136, 137, 149, 150]. While in the null embryos the p63 protein is missing altogether (i.e., both the TA and ΔN isoforms), in EEC, AEC, LMS, and SHFM-IV patients the mutated p63 protein coexists with half of the normal dose of the wild-type protein. To better model the disease, the group of Dr. A. Mills (CSHL, USA) has generated mice bearing the R279H mutation (found in EEC patients). Homozygous embryos and newborn animals show a global phenotype similar, but not identical, to that of p63-/- [90], consisting in the absence of the HL, severely truncated FL, a thin translucent skin and craniofacial and palatal defects. The HL defects in both the p63 null and in the p63-R279H homozygous embryos are evident as early as E9.5, and are accompanied by loss of AER stratification and FGF8 expression [90, 136, 137]. Interestingly, mild limb defects are observed in heterozygous p63-R279H mice, the mouse model closer to EEC.

Mouse models of the AEC syndrome have also been generated. Compared to EEC patients, AEC patients suffer of extreme skin fragility but have normal limbs. The AEC-mutant p63 proteins appear to act in a dominant-negative fashion. Mice were generated in which either ∆Np63a is down regulated in the skin, as a way to mimic the dominant negative action of mutant p63 in the AEC patients, or an AEC-mutant p63 was introduced [151 - 153]. These mice show severe skin erosion resembling the AEC phenotype, characterized by suprabasal epidermal proliferation, delayed terminal differentiation and altered basement membrane.

p63 mutations cause limb congenital phenotypes due to their impact on the AER Animal models show p63 is essential for epidermal stratification [90, 139, 154 - 156]. Considering that the AER is one of the earliest attempt of the embryonic (non-neural) ectoderm to organize into a multilayered epithelial tissue [157], it is not surprising that in p63 null or p63 R279H homozygous mice the AER is thinner and poorly stratified. Failure to maintain AER stratification and FGF8 expression is a common feature of various ectrodactyly phenotypes [90, 157 - 159].

p63 is expected to control AER functions via transcriptional regulation of AER-restricted target genes [122, 154 - 156], indeed failure of AER stratification has also been associated with loss of expression of key morphogens for limb development, such as FGF8 and Dlx5;Dlx6 [122]. Dlx genes are the vertebrate homologs of Drosophila Dll, a homeodomain transcription factors required for the specification of distal limb elements in the fly embryo. In Dll hypomorphic mutant flies, a variable set of phenotypes is observed depending on the mutation, ranging from fusion of the distal segments (weak mutants) to complete loss of distal and medial leg segments (severe mutants) [160, 161]. In mice Dlx genes have a prominent role in specifying the mandible and maxillary skeletal structures [162, 163], as well as controlling normal limb development [114, 115]. Point mutations of DLX5 have been found to co-segregate in familiar cases of SHFM [116] and the combined deletion of Dlx5 and Dlx6 leads to ectrodactyly of the HLs, that is, a true mouse model of SHFM type-I. There is evidence that until E11.5 the AER appears and functions normally, including a normal morphology and normal expression of AER markers (FGF4, FGF8, Msx2). On the contrary, at E11.5-E12 the expression of AER markers indicate that the central wedge of the AER fails to function. At about the same time the first signs of dysmorphology are visible. The expression of FGF8 and other markers declines in the central sector of the limb bud, accompanied by loss of stratification in the same territory [158], while the expression of SHH, Hand2 and Tbx4 in the mutant limbs is unchanged. Considering the expression pattern of Dlx genes in the limb, the Dlx5;Dlx6 null defect can be summarized as a cell-autonomous failure of the central AER to maintain and express morphogenetic molecules.

p63 and Dlx proteins are co-expressed in the AER cells [90] as well as in the fins of the zebrafish embryos [91, 92]. In homozygous p63 null and p63EEC (R279H) mutant limbs, the expression of four Dlx genes is strongly reduced. Functionally, when the p63+/EEC (heterozygous) mutation is combined with an incomplete loss of Dlx5 and Dlx6 alleles, severe limb phenotypes are observed, not present in mice with either mutation alone [90]. Together, there is a clear evidence for p63-Dlx regulatory cascade that is functional for distal limb development.

In vitro, ∆Np63α induces transcription from the Dlx5 and Dlx6 promoters, an activity abolished by EEC and SHFM-IV mutations, but not by AEC-associated mutations. ChIP analysis shows that p63 occupies the Dlx5 and Dlx6 promoters. This regulation takes place both at the proximal promoter level [90] and via a conserved cis-acting genomic element, located 250 kb centromeric to DLX5, an element that is specifically deleted in few SHFM patients [164]. Recent studies have identified a tissue-specific enhancer located within the coding exons 15 and 17 of the Dync1/1 gene (near the Dlx5;Dlx6 locus). This genomic element is characterized by an enhancer-type chromatin signature and physically interacts with a DLX5/6 promoter region 900 kb distal to DYNC1/1, specifically in the limb [165, 166]. Using copy number variation (CNV) analyses in SHFM patients, combined with whole genome sequencing to map deletion and translocation breakpoints, a recent study shows that the DYNC1I1 enhancers are also critical for limb development in humans [167]. An additional enhancer was identified in an intron of the Slc25a13 locus, close to Dlx5;Dlx6, and was shown to drive Dlx gene expression in the otic vesicle, forebrain, branchial arch and limbs of the developing embryo [165, 166]. It is plausible that the SHFM phenotype linked with mutations in these enhancers is caused by an altered regulation of Dlx5/6 transcription.

2.3. Downstream of Dlx5;Dlx6

Sp8 is a transcription factor of the Sp1 zinc-finger family [168, 169], homologous to the Drosophila D-Sp1 gene that has been implicated in appendage development [170]. In the developing limbs Sp8 shows restricted expression in the ectoderm, including the AER cells [168]. Mouse embryos null for Sp8 show severe developmental defects affecting the distal portion of the limbs, associated with a strongly reduced expression of FGF8 [168, 169, 171], Sp8 is co-expressed with Dlx genes in the murine AER and forebrain [172] and appears in the top 1% of a list of conserved/co-expressed genes in microarray data [173]. Furthermore, conserved Dlx5 DNA-binding sites are predicted near the Sp8 locus, thus Sp8 is a likely direct Dlx5 transcriptional target. A Dlx5-Sp8 transcriptional cascade could be upstream of FGF8 expression, which in turn maintains p63 protein stability.

A number of observations suggest that p63 and Dlx proteins may regulate FGF8 expression by acting directly on the genomic region corresponding to the SHFM type-III critical region [120, 121]: indeed p63-binding sites are present within the region, as demonstrated by ChIP-seq screening [164], and several predicted Dlx5 binding sites cluster around the FGF8 locus, in genomic regions conserved across mammalian species [158] (unpublished data).

Considering that the AER of Dac heterozygous embryos shows reduced FGF8 expression and defective cell layering [159], and considering that rearrangements/duplications around Dactylin do not disrupt or interrupt the gene, and since Dactylin is ubiquitously expressed in mouse tissues, the role of Dactylin as disease-gene is doubtful [122]. In alternative, FGF8 and components of the NFkB pathway might be the disease-genes. It is tempting to speculate that the complex duplication rearrangement modifies the position/organization of cis-acting control elements, which in turn affect expression of FGF8 and components of the NFkB pathway. Thus SHFM type-III could be a genome-misorganization type of genetic disease.

In further support of this, genome-wide CNV analyses on a Chinese family with SHFM type-III revealed a micro-duplication on chromosome 10q24 co-segregating with the SHFM phenotype [174]. This novel duplication contains two discontinuous DNA fragments: the minimal centromeric duplicated segment involves LBX1, POLL and a disrupted BTRC; the telomeric duplication encompasses DPCD and part of FBXW4. No coding and splice-site mutations of candidate genes in the region were found. Interestingly, the second duplicated fragment comprises Dlx5 and p63 DNA binding sites [164].

Another pathway that links p63 and Dlx5 in the regulation of the FGF8 locus implicates the gene IKKα, a direct transcriptional target of p63 relevant for ectoderm development and limb morphogenesis [175 - 177]. Interestingly, while mutations of p63 and loss of Dlx5;6 lead to a reduced FGF8 expression in the AER, in IKKα mutant embryos the AER shows an increased FGF8 expression [178], which nevertheless results in distal limb truncations and severe malformations.

From the above considerations, it appears that numerous players in the p63-Dlx5 cascade may contribute to regulate FGF8 expression in the AER. The possibility that the FGF8 locus is a common target of the p63 and the Dlx5 networks during limb development is in agreement with the well-known functions of FGF8 to sustain epithelial mesenchymal signalling and assure the timely generation of mesenchymal progenitors [36].

2.4. Post-translational p63 protein regulations

Several biochemical observations indicate that the ∆N- and TAp63 proteins are tightly regulated at post-translational level, via protein modification (phosphorylation, sumoylation and ubiquitination) and protein-protein interactions [126, 158, 179, 180]. Such modifications modulate the stability of the p63 protein, regulate its transcriptional activity and ultimately modulate its ability to orchestrate the timing of exit from the cell cycle and the dynamic of stratification of mammalian ectoderm [156, 181, 182].

Among the interacting or modifying proteins, MDM2 and p53 have been previously recognized [179, 180]. Recently we have shown that the peptidyl-prolyl cis/trans isomerase NIMA-interacting-1, Pin1, is a regulator of ∆Np63α protein stability, inducing its proteasome-mediated degradation [158] resulting in diminished transcription of two p63 targets [183] (Figure 2).

Another modification is acetylation, catalyzed by histone acetyl-transferase on lysine residues, and known to finely regulate p53 and p73 stability and transcriptional activity [184 - 189]. p73 is acetylated by p300 on lysine residues in the DBD and Oligomerization Domain [190] enhancing p73 ability to bind and activate proapototic target genes [191]. The p73-p300 interaction requires the prolyl-isomerase Pin1, which induces conformational changes following phosphorylation by the tyrosine kinase c-Abl [192]. Acetylation of p53 correlates with its stabilization and activation by antagonizing the activity of the MDM2 ubiquitin-ligase. It is interesting to note that a naturally occurring p63 mutation found in SHFM type-IV patients changes lysine 193 into glutamic acid (K193E) [125, 146, 147, 193]. Our unpublished data show that ∆Np63α is acetylated by p300 on the K193 residue, and that the K193E mutation prevents this modification (Guerrini and Restelli, unpublished) (Figure 2).

Figure 2.

Schematic representation of the molecules and their interactions that regulate the stability of ΔNp63 during the AER stratification. p63 regulates its own stability via to the expression of FGF8; this pathway includes the Dlx5/Dlx6 disease genes. Fgf8 stabilizes p63 by counteracting the activity of Pin1 to induce proteasome-mediated degradation of ΔNp63α. Novel results indicate that FGF8 activates a signalling cascade leading to activation of c-Abl that promotes phosphorylation of ΔNp63α on tyrosine residues. This phosphorylation event is required for the interaction of ΔNp63α with the p300 acetyl-transferase, which modulated ΔNp63α stabilization and transcriptional activity. Although only shown in vitro, we speculate that regulation may also occur in the AER cells (dotted line).

2.5. Emerging roles of FGF8

Expression of FGF8 is strongly reduced in the AER of the p63 null, R279H p63 mutant, and Dlx5;Dlx6 mutant embryos [90, 115] as well as several other mouse strains with distal limb defects. The AER of these mutants appears poorly stratified. Thus, loss of AER stratification and reduced FGF8 expression are a common theme during the onset of this specific class of malformations. The link between FGF8 expression and AER stratification is not totally clear. When FGFR2 gene is deleted in the AER cells, via conditional genetics, the AER loses stratification as well as Fgf8 expression. In this case, the AER cells cannot respond to (AER-derived?) FGFs [194] and it can be concluded that AER-expressed FGFs are needed for AER maintenance, apparently in an autocrine fashion.

An emerging role of FGF8 is the control of p63 stability in the AER cells. The AER of Dlx5/Dlx6 null mice shows poor stratification as well as reduced FGF8 expression, similar to what is seen in p63 mutant mice. We have documented that Dlx5;Dlx6 are transcriptional targets of p63, and that in turn FGF8 is a target of Dlx5. As already said, ∆Np63α protein stability is negatively regulated by the interaction with Pin1, via proteasome-mediated degradation. Recently we have shown that FGF8 counteracts Pin1-∆Np63α interaction, thus indicating that FGF8 participates in a feedback loop which involves the p63-Dlx5 cascade [158] (Figure 2).

p63 stability might also be regulated by another post-translational modification, namely acetylation by the p300 histone acetylase. c-Abl is a key regulator of the p53 family members and is known to be activated by treatment with FGF2 [192, 195 - 198]. Recently we have collected new data showing that FGF8 is able to stabilize ∆Np63α also via a novel pathway that requires the c-Abl tyrosine kinase and the protein acetylation by p300 (Guerrini and Restelli, unpublished). Thus, Dlx5, p63, Pin1, p300 and FGF8 participate in a time- and location-restricted regulatory loop that seems to be able to self-maintain and whose normal functioning is necessary for AER stratification, hence for normal extension and patterning of the limb buds. These results shed new light on the general molecular mechanisms at the basis of the SHFM and EEC limb malformations (Figure 2).

In an interesting set of experiments using cultured embryonic limbs, it was recently shown that the FGF/MAPK pathway establishes a high-distal to low-proximal gradient that controls the migration velocity of mesenchymal cells [199]. These cell movements enable continuous rearrangement of the cells at the distal tip of the limb bud. The effect of FGF/MAPK signalling emanating from the AER is different than the effect induced by Wnt5a in the limb bud. While Wnt5a induces directional movement of cells, FGF8 acts to induce rapid, yet disorganized, movements. Ultimately, the activity of both Wnt5a and FGF results in distal elongation (Figure 3). These observations suggest that FGF8 acts by inducing random movements, but with a higher velocity as cells move close to the source. A study proposes that the FGF pathway drives tail-bud elongation in the chick embryo by promoting random cell movements [200]. According to these authors FGF creates a gradient of cell motility and that the tail bud elongates by mass action of random cell movement at the posterior end of the embryo. Although this data indicate a similar mode of FGF action, cells in the limb bud additionally undergo oriented processes of cell division and directional movements under the influence of Wnt5a. This study indicates that it is the combined action of non-canonical WNT and FGF that integrates orientation and movement, consequently driving limb-bud elongation and thereby establishing a progenitor field of the proper dimensions for the subsequent patterning and morphogenesis of limb anatomy.

Figure 3.

Schematic representation of the mesenchymal cells orientation and organization in the early limb bud. These cellular events are regulated by the combined activities of the WNT and FGF pathways. Wnt5A/Jnk/PCP pathway is necessary for the proper orientation of cell movements and cell division. In contrast, the FGF/MAPK signaling pathway, emanating from the AER establishes a gradient of cell velocity. The combination of oriented cell divisions and movements drives the P-D extension of the limb bud necessary for subsequent morphogenesis.

2.6. Wnt signalling and limb development

Wnt molecules are the vertebrate homologs of the Drosophila wingless gene, required for wing development. Wnt molecules are involved in all aspects of embryonic development, from patterning to morphogenesis and cell-tissue interactions [201 - 203].

Several members of the Wnt family of ligands are expressed in the ectoderm and mesenchyme of the developing limbs. At early stages, Wnt8c and Wnt2b are transiently expressed in the LPM and participate in the initiation of HL and FL outgrowth, respectively [43]. At later stages, Wnt3/Wnt3a and Wnt5a are expressed by the AER cells while Wnt7a is expressed in the dorsal ectoderm.

Wnt ligands signal through the Frizzled (Fz) seven-pass trans-membrane receptors. In the “canonical” pathway, binding of Wnt ligands to Fz receptors represses the axin/glycogen synthase kinase-3β (GSK3β) complex, which in the absence of the ligand promotes the degradation of β-catenin via the ubiquitin pathway (reviewed in reference [204]). In Wnt-activated cells, cytoplasmic β-catenin accumulates and translocates to the nucleus where, in conjunction with T cell-specific factor/lymphoid enhancer binding factor-1 (Tcf/Lef1) transcription factors, activates transcription of target genes.

A role of “canonical” Wnt signalling in limb development has long been recognized [205]. In the chick limb bud, the Wnt/β-catenin pathway is essential for the induction and maintenance of the AER. Indeed, ectopic expression of Wnts in the interflank region prior to limb outgrowth induces ectopic FGF10 expression and limb formation. FGF10 subsequently induces Wnt3a expression in the AER, which in turn switches on the expression of Fgf8, again via the β-catenin pathway, and promotes AER formation [43, 44, 206].

In the chick embryo Wnt3a mediates the Wnt/β-catenin signalling required for establishment of the AER. In the mouse, old data indicate that mouse embryos lacking the Wnt/β-catenin pathway component LRP6, or simultaneously lacking Lef1 and Tcf1, exhibit defective AER formation and limb defects, indicating that his pathway is indeed essential for AER formation [207, 208]. However, Wnt3a is not expressed in the limb ectoderm of the mouse embryo, and Wnt3a null embryos do not show limb defects [209, 210]. Instead, the closely related Wnt3 gene is expressed in the limb ectoderm [210] and the conditional removal of Wnt3 in the limb ectoderm leads to severe distal limb truncations and AER malfunction. Similar results were obtained by the conditional removal of β-catenin in the limb ectoderm [211], strongly suggesting that the murine Wnt3 is functionally homologous to chick Wnt3a, and that a pre-AER active Wnt3/β-catenin pathway in the embryonic ectoderm is essential for AER formation and maintenance. Notably, homozygous mutations of WNT3 in human are associated with a rare autosomal recessive congenital disorder known as tetra-amelia [212] characterized by the absence of all four limbs.

Wnt signalling has been implicated in removing “excess” tissue by programmed cell death and sculpting the limb shape. Indeed, the ability of BMP4 to induce cell death in the developing limb appears to be mediated by Dkk1 [213]. Loss of function of Dkk1 in mice results in the downregulation of Msx1, a component of the cell death pathway, in the anterior and posterior necrotic zones and the interdigital mesenchyme, whilst gain of Dkk1 function in chicks causes excessive cell death via activation of the c-jun pathway [213, 214]. The decrease in cell death in the mouse mutants contributes to the polydactyly and fusion of digits that occur in Dkk1 mutant mice [214]. In addition, Fz2, –3 and –4, and Dkk2, and –3, are expressed in the interdigit mesenchyme, suggesting that a fine balance of Wnt signalling controls cell death/survival in this region [215, 216].

2.7. Emerging role of Wnt5a and non-canonical signalling

Wnt ligands can also activate two other branches of “non-canonical” pathways; one of these is known as the planar cell polarity (PCP) pathway, involves Fz receptors and dishevelled (Dvl), which interact with a distinct set of “PCP proteins” such as Van Gogh (Vang) and Prickle [217]. The PCP pathway recruits the small GTPases Rho and Cdc42 and the c-Jun N-terminal kinase (JNK) [218 - 220]. Initially identified in Drosophila, PCP establishes cellular polarity in the plane of an epithelium, perpendicular to the apical-basal orientation [217]. Studies in vertebrate model systems, including Xenopus and zebrafish, indicate that the PCP pathway also regulates a morphogenetic process known as convergent extension (CE). CE was first demonstrated in gastrulating Xenopus embryos in which mesodermal cells underwent medio-laterally oriented intercalation, leading to concomitant tissue lengthening and narrowing [221]. Imaging experiments in zebrafish indicate that, in addition to polarized cell intercalation, the PCP pathway also regulates directional cell migration and oriented cell division underlying CE [222 - 225]. A second branch of “non-canonical” Wnt transduction pathways leads to the release of intracellular Ca2+ and the activation of protein kinase C (PKC) and Ca2+/Calmodulin-dependent Kinase-II (CamKII) [226 - 229]. The choice of the pathway being activated by a Wnt ligand appears to depend mostly on the receptor profile and on the intracellular signalling molecules available in a given cell type, and little on the Wnt ligand itself.

A role of “non-canonical” Wnt signalling during limb development has been recognized, although the cellular and molecular mechanisms are not fully clarified. The vertebrate Wnt5a gene, the homolog to Drosophila Dwnt-5 gene essential for limb and appendage development, is considered the typical non-canonical Wnt, involved in the establishment of PCP [230 - 232]. Wnt5a together with Wnt11 mediates the activation of PCP during the CE in frogs and zebrafish [223, 233, 234], and during mouse limb development Wnt5a is expressed in a gradient from the AER to the proximal mesechymal cells, is regulated by FGF signalling from the AER and has been shown to inhibit β-catenin degradation [235, 236].

In addition to the PCP pathway, Wnt5a has been shown to activate at least two other non-canonical pathways. The first is known as the Wnt–Ca2+ pathway, in which Wnt5a stimulation induces Ca2+ release and subsequent activation of the Ca2+-sensitive kinases protein kinase C and Ca2+/calmodulin-dependent kinase [226, 227, 237, 238]. Over-expression of the core PCP proteins, Dvl and Pk, can also activate the Wnt–Ca2+ cascade in zebrafish and Xenopus, suggesting that the Wnt–Ca2+ and PCP pathways either overlap substantially or are components of the same signalling network [229, 239, 240]. Second, in mammalian cells Wnt5a has been shown to antagonize the canonical Wnt pathway by either promoting GSK3β-independent β-catenin degradation [236] or by inhibiting β-catenin-dependent transcription [241].

Wnt5a can signal through different Fz receptors and co-receptors, but also via non-conventional tyrosine-kinase like receptors (Ror2 and Ryk) and can activate both the canonical and the non-canonical Wnt pathways [241, 242]. Activation of the canonical pathway entails the Lrp5 and Lrp6 co-receptors, which through cytoplasmic Dvl promote stabilization of β-catenin, its nuclear translocation and the activation of gene transcription [243, 244]. However, the distinct phenotypes observed between Wnt3, β-catenin, Lrp5/Lrp6 and Wnt5a mutant mice [245] argues that during limb development Wnt5a does not signal through the β-catenin pathway [246].

In human, missense mutations in WNT5A have been documented in an autosomal dominant form of RRS (MIM #180700) [247 - 249] implying that a disruption of Wnt5a signalling may underlie both RRS and BDB1. Homozygous ROR2 mutations have been linked to the autosomal recessive form of Robinow syndrome (or COVESDEM syndrome) (MIM #268310), while heterozygous ROR2 mutations lead to type brachydactyly (MIM #113000) [250] and autosomal Dominant Brachydactyly type-B (BDB1, MIM #113000). BDB1 is the most severe form of brachydactyly and is characterized by loss of nails and varying number of phalanges [100, 251]. In contrast, RRS patients display broader skeletal dysplasia including mesomelic limb shortening and dwarfism, and may or may not display brachydactyly [97, 98, 101].

In mice, the disruption of Wnt5a results in short metacarpal elements, absence of phalanges and truncations of proximal elements [236, 252, 253]. The remaining limb skeletal elements are significantly shortened and the severity of the phenotype follows a gradient, with distal bones more affected than proximal ones, reminiscent of mesomelic limb shortening in RRS patients. Interestingly, the AER appears normally stratified and expresses FGF8 [252]. Strong evidence of the involvement of the Wnt5-dependent pathways in limb development is derived from phenotypes of mice with loss of Wnt5a receptors. In addition to Fz receptors, Wnt5a binds to both Ryk and Ror2 receptors and regulates PCP by promoting Vangl2 stability during limb extension [242, 254]. Ryk and Ror2 are single-pass tyrosine-kinase type of receptors [241, 255]. Ror2 (an orphan tyrosine kinase receptor) activates JNK [256] and in Xenopus has been shown to interact with Wnt11 and Fz7 to regulate CE, suggesting that it may be part of the PCP pathway [257]. Upon binding with Wnt5a, Ror2 inhibits the canonical Wnt signalling. Furthermore, Ror2 also plays an important role in chondrogenesis. Ror2 is selectively expressed in chondrocytes of cartilage anlagen, and is thus probably important in their initial growth and patterning. Mice mutant for Ror2 and double mutants for Ror1;Ror2 exhibit phenotypes that correspond to human RRS malformation, and bear similarities with the Wnt5a mutant mice [258, 259]. Ryk is another unconventional Wnt5a receptor, consisting in a single transmembrane pass, catalytically-inactive, tyrosine kinase molecule. Ryk mutant mice show limb truncation similar to those of Wnt5a null embryos [260]. Finally, disruption of PCP signalling as Vangl2 in mice causes limb morphogenesis and skeletal defects and may underlie the Robinow syndrome and brachydactyly type B [261]. Together, these observations indicate that Wnt5a, Ryk and Ror2 molecules produce similar phenotypes when lost, for example, the disruption of components of the Wnt non-canonical pathway causes similar limb developmental defects.

2.8. Wnt5a controls aspects of PCP and CE in limb development

Recent data [199] shed light on the cellular functions of Wnt5a during limb development. Inspired by the CE process and the PCP pathway, first described in lower organisms, the authors examined the proliferative expansion and migration of mesenchymal cells of the mouse limb bud; in particular, they examined the orientation of cell division and movements in response to Wnt5a. The combination of oriented cell divisions and movements drives the P-D elongation of the limb bud necessary to set the stage for subsequent morphogenesis. They show that Wnt5a via the JNK PCP pathway is needed for the proper orientation of mesenchymal cell movements and cell division reminiscent of CE in Xenopus and zebrafish [222 - 225] (Figure 3).

Although these recent studies implicate Wnt5a in the oriented migration and cell division of the mesenchymal cells, little is known about the ectoderm cells, and in particular the AER cells, in which Wnt5a is expressed. It is conceivable that the AER cells might be the prime (autocrine) cellular target of Wnt5a, and that the acquisition of a correct planar orientation is a requisite for correct AER formation. Wnt5a and Dlx5 have an overlapping expression pattern, and the phenotype of Wnt5a null mice, although not identical, is quite similar to that of Dlx5/Dlx6 mutant. One possibility is that a deregulation of Wnt5a expression, secondary to the disruption of Dlx5;Dlx6 may underlie ectrodactyly of the Dlx5;Dlx6 mutant embryos (Figure 4). In support of this, we have evidence that Dlx genes promote neuronal differentiation via Wnt5a, and that Dlx2 and Dlx5 physically occupy conserved genomic elements near the Wnt5a locus and activate its transcription [262]. This interaction and regulation is likely to occur also in the AER cells, a possibility that remains to be investigated.

Figure 4.

Proposed model of regulation of the AER cell orientation. Dlx5 is known to regulate the transcription of both FGF8 and Wnt5a. In turn, FGF8 is required for AER maintenance and stratification, via p63, while the function of Wnt5a for AER maintenance is still poorly known. Recent data of the regulation of orientation and velocity of mesenchymal cells by, respectively, the Wnt5a/PCP and the FGF8-MAP-Erk pathways open the possibility that Wnt5a may regulate the orientation/motility of the AER cell and assure a correct stratification (dotted line).

2.9. Quantitative and dynamic gene expression in limb development

An emerging theme in developmental biology is the importance of gene dosage and dynamic gene expression for correct morphogenesis [56]. Several Dlx (1, 2, 3, 5 and 6) and FGF (4, 8, 9 and 17) genes are co-expressed in the AER, and their expression is dynamically regulated, both with respect to time (embryonic age) and location (territory of expression). In addition, there is evidence that Dlx and FGF genes are functionally redundant, at least in part. For example, no limb phenotype is observed in mice null for only one Dlx gene, while ectrodactyly is observed in Dlx2;Dlx5 null mice [161] and the ectrodactyly of Dlx5;Dlx6 null mice is fully rescued by the re-expression of Dlx5 alone [115]. Likewise, an increased severity of craniofacial phenotypes correlates with progressive loss of Dlx gene [263, 264]. All these are indications of gene-dosage effects between functionally redundant genes.

We propose that the portion of the p63 network that (direct or indirect) regulates FGF8 expression is exerted in a quantitative and dynamic mode. To support this, we should consider that although p63 null and p63EEC homozygous mice show severe limb truncation or absence, the heterozygous mice appear to be normal. When heterozygous EEC mice are bred with heterozygous Dlx5;Dlx6 ones (the latter have normal limbs), anomalies are clearly observed [90].

A gene-dosage effect combined with the co-expression of functionally redundant genes implies the existence of a threshold level to be maintained to assure AER stratification and signalling functions. Indeed, we have noted that the expression of Dlx2 and Dlx5 is lower in the central portion of the AER, compared to the anterior or posterior segments [122]. Thus, the central AER might be more sensitive to reduced Dlx expression due to intrinsic lower expression. On the same line, there is evidence that certain amount of AER-derived pan-FGF is required to induce and maintain the underlying mesenchymal progenitors [36, 56, 157]. In fact, in the Dlx5; Dlx6 mutant limbs, the reduction of FGF8 expression is restricted to the central AER, the region where epithelial-mesenchymal signalling is primarily defective and the region where morphogenesis fails [114, 115]. Thus, the entire p63-Dlx-FGF cascade is sensitive to gene dosage and position of expression.

Advertisement

3. Concluding remarks

p63 is a master regulator of ectodermal cell proliferation, differentiation and stratification, and has a key role in the establishment of a positive loop that maintains FGF8 expression. In turn, our recent data reveal a novel role of FGF8 to (directly and indirectly) stabilize the p63 proteins and modulate their transcriptional activity. Thus, in the biology and development of the ectoderm, p63 post-translational modifications are as important as p63 gene expression and may reveal novel targets to be used in p63 modulation.

We illustrate that the p63-Dlx5 transcriptional regulation is at the centre of a pathway relevant for the SHFM malformation. The stability of p63 and the activation of the pathway appear to be under the regulation of FGF8, which in turn is regulated by the pathway. In addition to decipher this positive regulatory loop, these data support a model to attempt to explain the SHFM-III pathogenesis in terms of genome positional effects on the FGF8 locus.

FGF8 and Wnt5a provide instructions for mesoderm cells as to which direction and orientation to take, at the basis of AER formation and proper migration of mesenchymal cells. This instruction adopts molecules of the PCP pathway, most likely inducing convergent extension. While this has been recently demonstrated for the mesenchymal cells, the possibility that a Wnt5a-dependent PCP pathway is also functional for the organization and stratification of the AER cells remains to be addressed. Notably, data from the human malformation diseases and the corresponding animal models clearly suggest so.

The study of animal models of EEC and SHFM diseases has provided much of this knowledge, and will continue to do so. The big hope is that, once the pathways will be elucidated, we might be able to exploit diffusible molecules and attempt to correct the limb malformation defects. Preliminary attempts are being conducted on whole-organ cultured limbs.

Advertisement

Nomenclature

A-P, anterior-posterior

D-V, dorsal-ventral

P-D, proximal-distal

SHH, sonic hedgehog

FGF, fibroblast growth factor

FL, forelimb

HL, hindlimb

ZPA, zone of polarizing activity

AER, apical ectodermal ridge

PZ, progress zone

KO, knock-out

PCP, planar cell polarity

CE, convergent extension

LPM, lateral plate mesoderm

References

  1. 1. Ohuchi, H., et al. (1997) The mesenchymal factor, FGF10, initiates and maintains the outgrowth of the chick limb bud through interaction with FGF8, an apical ectodermal factor. Development 124, 2235–2244
  2. 2. 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, 753–765
  3. 3. Min, H., et al. (1998) Fgf-10 is required for both limb and lung development and exhibits striking functional similarity to Drosophila branchless. Genes Dev 12, 3156–3161
  4. 4. Sekine, K., et al. (1999) FGF10 is essential for limb and lung formation. Nat Genet 21, 138–141
  5. 5. Ohuchi, H., et al. (1995) An additional limb can be induced from the flank of the chick embryo by FGF4. Biochem Biophys Res Commun 209, 809–816
  6. 6. Vogel, A., et al. (1996) Involvement of FGF-8 in initiation, outgrowth and patterning of the vertebrate limb. Development 122, 1737–1750
  7. 7. Gros, J., and Tabin, C.J. (2014) Vertebrate limb bud formation is initiated by localized epithelial-to-mesenchymal transition. Science 343, 1253–1256
  8. 8. Agarwal, P., et al. (2003) Tbx5 is essential for forelimb bud initiation following patterning of the limb field in the mouse embryo. Development 130, 623–633
  9. 9. Ahn, D.G., et al. (2002) T-box gene tbx5 is essential for formation of the pectoral limb bud. Nature 417, 754–758
  10. 10. Begemann, G., et al. (2002) Cloning of zebrafish T-box genes tbx15 and tbx18 and their expression during embryonic development. Mech Dev 114, 137–141
  11. 11. Garrity, D.M., et al. (2002) The heartstrings mutation in zebrafish causes heart/fin Tbx5 deficiency syndrome. Development 129, 4635–4645
  12. 12. Minguillon, C., et al. (2005) Tbx5 and Tbx4 are not sufficient to determine limb-specific morphologies but have common roles in initiating limb outgrowth. Dev Cell 8, 75–84
  13. 13. Rallis, C., et al. (2003) Tbx5 is required for forelimb bud formation and continued outgrowth. Development 130, 2741–2751
  14. 14. Saito, D., et al. (2002) Specification and determination of limb identity: evidence for inhibitory regulation of Tbx gene expression. Development 129, 211–220
  15. 15. Takeuchi, J.K., et al. (2003) Tbx5 and Tbx4 trigger limb initiation through activation of the Wnt/Fgf signaling cascade. Development 130, 2729–2739
  16. 16. Menke, J. (2008) Contrast-enhanced magnetic resonance angiography in peripheral arterial disease: improving image quality by automated image registration. Magn Reson Med 60, 224–229
  17. 17. Mic, F.A., et al. (2004) Retinoic acid synthesis controlled by Raldh2 is required early for limb bud initiation and then later as a proximodistal signal during apical ectodermal ridge formation. J Biol Chem 279, 26698–26706
  18. 18. Begemann, G., and Meyer, A. (2001) Hindbrain patterning revisited: timing and effects of retinoic acid signalling. Bioessays 23, 981–986
  19. 19. Gibert, Y., et al. (2006) Induction and prepatterning of the zebrafish pectoral fin bud requires axial retinoic acid signaling. Development 133, 2649–2659
  20. 20. Maden, M. (1982) Vitamin A and pattern formation in the regenerating limb. Nature 295, 672–675
  21. 21. Tickle, C., et al. (1982) Local application of retinoic acid to the limb bond mimics the action of the polarizing region. Nature 296, 564–566
  22. 22. Mercader, N., et al. (2006) Prdm1 acts downstream of a sequential RA, Wnt and Fgf signaling cascade during zebrafish forelimb induction. Development 133, 2805–2815
  23. 23. Mercader, N., et al. (2000) Opposing RA and FGF signals control proximodistal vertebrate limb development through regulation of Meis genes. Development 127, 3961–3970
  24. 24. Litingtung, Y., et al. (2002) SHH and Gli3 are dispensable for limb skeleton formation but regulate digit number and identity. Nature 418, 979–983
  25. 25. Grandel, H., et al. (2002) Retinoic acid signalling in the zebrafish embryo is necessary during pre-segmentation stages to pattern the anterior-posterior axis of the CNS and to induce a pectoral fin bud. Development 129, 2851–2865
  26. 26. Tabin, C., and Wolpert, L. (2007) Rethinking the proximodistal axis of the vertebrate limb in the molecular era. Genes Dev 21, 1433–1442
  27. 27. Zhao, X., et al. (2009) Retinoic acid promotes limb induction through effects on body axis extension but is unnecessary for limb patterning. Curr Biol 19, 1050–1057
  28. 28. Niederreither, K., et al. (1999) Embryonic retinoic acid synthesis is essential for early mouse post-implantation development. Nat Genet 21, 444–448
  29. 29. Cunningham, T.J., et al. (2013) Antagonism between retinoic acid and fibroblast growth factor signaling during limb development. Cell Rep 3, 1503–1511
  30. 30. Bell, S.M., et al. (1998) The loss of ventral ectoderm identity correlates with the inability to form an AER in the legless hindlimb bud. Mech Dev 74, 41–50
  31. 31. Todt, W.L., and Fallon, J.F. (1986) Development of the apical ectodermal ridge in the chick leg bud and a comparison with the wing bud. Anat Rec 215, 288–304
  32. 32. Todt, W.L., and Fallon, J.F. (1987) Posterior apical ectodermal ridge removal in the chick wing bud triggers a series of events resulting in defective anterior pattern formation. Development 101, 501–515
  33. 33. Saunders, J.W., Jr. (1948) The proximo-distal sequence of origin of the parts of the chick wing and the role of the ectoderm. J Exp Zool 108, 363–403
  34. 34. Fallon, J.F., et al. (1986) Apical ectodermal ridge maintenance in ovo and in vitro. Prog Clin Biol Res 226, 103–113
  35. 35. Niswander, L., et al. (1993) FGF-4 replaces the apical ectodermal ridge and directs outgrowth and patterning of the limb. Cell 75, 579–587
  36. 36. Mariani, F.V., et al. (2008) Genetic evidence that FGFs have an instructive role in limb proximal-distal patterning. Nature 453, 401–405
  37. 37. Crossley, P.H., et al. (1996) Roles for FGF8 in the induction, initiation, and maintenance of chick limb development. Cell 84, 127–136
  38. 38. Moon, A.M., and Capecchi, M.R. (2000) Fgf8 is required for outgrowth and patterning of the limbs. Nat Genet 26, 455–459
  39. 39. Boulet, A.M., et al. (2004) The roles of FGF4 and Fgf8 in limb bud initiation and outgrowth. Dev Biol 273, 361–372
  40. 40. Lewandoski, M., et al. (2000) Fgf8 signalling from the AER is essential for normal limb development. Nat Genet 26, 460–463
  41. 41. Lu, P., et al. (2006) Increasing FGF4 expression in the mouse limb bud causes polysyndactyly and rescues the skeletal defects that result from loss of Fgf8 function. Development 133, 33–42
  42. 42. Sun, X., et al. (2002) Functions of FGF signalling from the apical ectodermal ridge in limb development. Nature 418, 501–508
  43. 43. Kawakami, Y., et al. (2001) WNT signals control FGF-dependent limb initiation and AER induction in the chick embryo. Cell 104, 891–900
  44. 44. McQueeney, K., et al. (2002) Beta-catenin-dependent Wnt signaling in apical ectodermal ridge induction and FGF8 expression in normal and limbless mutant chick limbs. Dev Growth Differ 44, 315–325
  45. 45. Summerbell, D., et al. (1973) Positional information in chick limb morphogenesis. Nature 244, 492–496
  46. 46. Dudley, A.T., et al. (2002) A re-examination of proximodistal patterning during vertebrate limb development. Nature 418, 539–544
  47. 47. Cooper, K.L., et al. (2011) Initiation of proximal-distal patterning in the vertebrate limb by signals and growth. Science 332, 1083–1086
  48. 48. Rosello-Diez, A., et al. (2011) Diffusible signals, not autonomous mechanisms, determine the main proximodistal limb subdivision. Science 332, 1086–1088
  49. 49. Tickle, C., et al. (1975) Positional signalling and specification of digits in chick limb morphogenesis. Nature 254, 199–202
  50. 50. Zwilling, E. (1956) Genetic mechanism in limb development. Cold Spring Harb Symp Quant Biol 21, 349–354
  51. 51. Riddle, R.D., et al. (1993) Sonic hedgehog mediates the polarizing activity of the ZPA. Cell 75, 1401–1416
  52. 52. Wolpert, L. (1969) Positional information and the spatial pattern of cellular differentiation. J Theor Biol 25, 1–47
  53. 53. Harfe, B.D., et al. (2004) Evidence for an expansion-based temporal SHH gradient in specifying vertebrate digit identities. Cell 118, 517–528
  54. 54. Scherz, P.J., et al. (2007) Extended exposure to Sonic hedgehog is required for patterning the posterior digits of the vertebrate limb. Dev Biol 308, 343–354
  55. 55. Zeller, R. (2004) It takes time to make a pinky: unexpected insights into how SHH patterns vertebrate digits. Sci STKE 2004, pe53
  56. 56. Zeller, R. (2010) The temporal dynamics of vertebrate limb development, teratogenesis and evolution. Curr Opin Genet Dev 20, 384–390
  57. 57. Panman, L., et al. (2006) Differential regulation of gene expression in the digit forming area of the mouse limb bud by SHH and gremlin 1/FGF-mediated epithelial-mesenchymal signalling. Development 133, 3419–3428
  58. 58. Robert, B., and Lallemand, Y. (2006) Anteroposterior patterning in the limb and digit specification: contribution of mouse genetics. Dev Dyn 235, 2337–2352
  59. 59. Michos, O., et al. (2004) Gremlin-mediated BMP antagonism induces the epithelial-mesenchymal feedback signaling controlling metanephric kidney and limb organogenesis. Development 131, 3401–3410
  60. 60. te Welscher, P., et al. (2002) Mutual genetic antagonism involving GLI3 and dHAND prepatterns the vertebrate limb bud mesenchyme prior to SHH signaling. Genes Dev 16, 421–426
  61. 61. te Welscher, P., et al. (2002) Progression of vertebrate limb development through SHH-mediated counteraction of GLI3. Science 298, 827–830
  62. 62. Cohn, M.J., et al. (2002) Branching, segmentation and the metapterygial axis: pattern versus process in the vertebrate limb. Bioessays 24, 460–465
  63. 63. Ros, M.A., et al. (1996) The limb field mesoderm determines initial limb bud anteroposterior asymmetry and budding independent of sonic hedgehog or apical ectodermal gene expressions. Development 122, 2319–2330
  64. 64. Capellini, T.D., et al. (2006) Pbx1/Pbx2 requirement for distal limb patterning is mediated by the hierarchical control of Hox gene spatial distribution and SHH expression. Development 133, 2263–2273
  65. 65. Bastida, M.F., et al. (2009) A BMP-SHH negative-feedback loop restricts SHH expression during limb development. Development 136, 3779–3789
  66. 66. Zeller, R., et al. (2009) Vertebrate limb bud development: moving towards integrative analysis of organogenesis. Nat Rev Genet 10, 845–858
  67. 67. Wang, B., et al. (2000) Hedgehog-regulated processing of Gli3 produces an anterior/posterior repressor gradient in the developing vertebrate limb. Cell 100, 423–434
  68. 68. Kondo, S., and Miura, T. (2010) Reaction-diffusion model as a framework for understanding biological pattern formation. Science 329, 1616–1620
  69. 69. Turing, A.M. (1990) The chemical basis of morphogenesis. 1953. Bull Math Biol 52, 153–197; discussion 119–152
  70. 70. Miura, T., et al. (2006) Mixed-mode pattern in Doublefoot mutant mouse limb--Turing reaction-diffusion model on a growing domain during limb development. J Theor Biol 240, 562–573
  71. 71. Kmita, M., et al. (2005) Early developmental arrest of mammalian limbs lacking HoxA/HoxD gene function. Nature 435, 1113–1116
  72. 72. Zakany, J., et al. (2004) A dual role for Hox genes in limb anterior-posterior asymmetry. Science 304, 1669–1672
  73. 73. Chen, Y., et al. (2004) Direct interaction with Hoxd proteins reverses Gli3-repressor function to promote digit formation downstream of SHH. Development 131, 2339–2347
  74. 74. Kmita, M., et al. (2002) Serial deletions and duplications suggest a mechanism for the collinearity of Hoxd genes in limbs. Nature 420, 145–150
  75. 75. Sheth, R., et al. (2012) Hox genes regulate digit patterning by controlling the wavelength of a Turing-type mechanism. Science 338, 1476–1480
  76. 76. Beauchemin, M., et al. (1998) Graded expression of Emx-2 in the adult newt limb and its corresponding regeneration blastema. J Mol Biol 279, 501–511
  77. 77. Geraudie, J., and Ferretti, P. (1998) Gene expression during amphibian limb regeneration. Int Rev Cytol 180, 1–50
  78. 78. Shubin, N., et al. (1997) Fossils, genes and the evolution of animal limbs. Nature 388, 639–648
  79. 79. Davis, A.P., et al. (1995) Absence of radius and ulna in mice lacking hoxa-11 and hoxd-11. Nature 375, 791–795
  80. 80. Fromental-Ramain, C., et al. (1996) Hoxa-13 and Hoxd-13 play a crucial role in the patterning of the limb autopod. Development 122, 2997–3011
  81. 81. Boulet, A.M., and Capecchi, M.R. (2004) Multiple roles of Hoxa11 and Hoxd11 in the formation of the mammalian forelimb zeugopod. Development 131, 299–309
  82. 82. Laufer, E., et al. (1994) Sonic hedgehog and Fgf-4 act through a signaling cascade and feedback loop to integrate growth and patterning of the developing limb bud. Cell 79, 993–1003
  83. 83. Niswander, L., et al. (1994) Function of FGF-4 in limb development. Mol Reprod Dev 39, 83–88; discussion 88–89
  84. 84. Khokha, M.K., et al. (2003) Gremlin is the BMP antagonist required for maintenance of SHH and Fgf signals during limb patterning. Nat Genet 34, 303–307
  85. 85. Parr, B.A., et al. (1998) The classical mouse mutant postaxial hemimelia results from a mutation in the Wnt 7a gene. Dev Biol 202, 228–234
  86. 86. Parr, B.A., and McMahon, A.P. (1995) Dorsalizing signal Wnt-7a required for normal polarity of D-V and A-P axes of mouse limb. Nature 374, 350–353
  87. 87. Chen, H., et al. (1998) Limb and kidney defects in Lmx1b mutant mice suggest an involvement of LMX1B in human nail patella syndrome. Nat Genet 19, 51–55
  88. 88. Riddle, R.D., et al. (1995) Induction of the LIM homeobox gene Lmx1 by WNT7a establishes dorsoventral pattern in the vertebrate limb. Cell 83, 631–640
  89. 89. Loomis, C.A., et al. (1998) Analysis of the genetic pathway leading to formation of ectopic apical ectodermal ridges in mouse Engrailed-1 mutant limbs. Development 125, 1137–1148
  90. 90. Lo Iacono, N., et al. (2008) Regulation of Dlx5 and Dlx6 gene expression by p63 is involved in EEC and SHFM congenital limb defects. Development 135, 1377–1388
  91. 91. Bakkers, J., et al. (2002) Zebrafish DeltaNp63 is a direct target of Bmp signaling and encodes a transcriptional repressor blocking neural specification in the ventral ectoderm. Dev Cell 2, 617–627
  92. 92. Lee, H., and Kimelman, D. (2002) A dominant-negative form of p63 is required for epidermal proliferation in zebrafish. Dev Cell 2, 607–616
  93. 93. Temtamy, S.A., and McKusick, V.A. (1978) The genetics of hand malformations. Birth Defects Orig Artic Ser 14, i–xviii, 1–619
  94. 94. Schwabe, G.C., and Mundlos, S. (2004) Genetics of congenital hand anomalies. Handchir Mikrochir Plast Chir 36, 85–97
  95. 95. Lettice, L.A., et al. (2003) A long-range SHH enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly. Hum Mol Genet 12, 1725–1735
  96. 96. Tsukurov, O., et al. (1994) A complex bilateral polysyndactyly disease locus maps to chromosome 7q36. Nat Genet 6, 282–286
  97. 97. Afzal, A.R., and Jeffery, S. (2003) One gene, two phenotypes: ROR2 mutations in autosomal recessive Robinow syndrome and autosomal dominant brachydactyly type B. Hum Mutat 22, 1–11
  98. 98. Afzal, A.R., et al. (2000) Recessive Robinow syndrome, allelic to dominant brachydactyly type B, is caused by mutation of ROR2. Nat Genet 25, 419–422
  99. 99. Ho, H.Y., et al. (2012) Wnt5a-Ror-Dishevelled signaling constitutes a core developmental pathway that controls tissue morphogenesis. Proc Natl Acad Sci U S A 109, 4044–4051
  100. 100. Schwabe, G.C., et al. (2000) Distinct mutations in the receptor tyrosine kinase gene ROR2 cause brachydactyly type B. Am J Hum Genet 67, 822–831
  101. 101. Schwarzer, W., et al. (2009) A gradient of ROR2 protein stability and membrane localization confers brachydactyly type B or Robinow syndrome phenotypes. Hum Mol Genet 18, 4013–4021
  102. 102. Newman, C.G. (1986) The thalidomide syndrome: risks of exposure and spectrum of malformations. Clin Perinatol 13, 555–573
  103. 103. Ito, T., and Handa, H. (2012) Deciphering the mystery of thalidomide teratogenicity. Congenit Anom (Kyoto) 52, 1–7
  104. 104. Therapontos, C., et al. (2009) Thalidomide induces limb defects by preventing angiogenic outgrowth during early limb formation. Proc Natl Acad Sci U S A 106, 8573–8578
  105. 105. Duijf, P.H., et al. (2003) Pathogenesis of split-hand/split-foot malformation. Hum Mol Genet 12 Spec No 1, R51–60
  106. 106. Gurrieri, F., and Everman, D.B. (2013) Clinical, genetic, and molecular aspects of split-hand/foot malformation: an update. Am J Med Genet A 161A, 2860–2872
  107. 107. Sowinska-Seidler, A., et al. (2014) Split-hand/foot malformation – molecular cause and implications in genetic counseling. J Appl Genet 55, 105–115
  108. 108. Crackower, M.A., et al. (1996) Characterization of the split hand/split foot malformation locus SHFM1 at 7q21.3-q22.1 and analysis of a candidate gene for its expression during limb development. Hum Mol Genet 5, 571–579
  109. 109. Del Porto, G., et al. (1983) [Interstitial deletion of the long arm of chromosome 7 and its clinical correlations]. Pathologica 75 Suppl, 268–271
  110. 110. Marinoni, J.C., et al. (1995) Split foot and developmental retardation associated with a deletion of three microsatellite markers in 7q21.2-q22.1. Clin Genet 47, 90–95
  111. 111. Scherer, S.W., et al. (2003) Human chromosome 7: DNA sequence and biology. Science 300, 767–772
  112. 112. Scherer, S.W., et al. (1994) Fine mapping of the autosomal dominant split hand/split foot locus on chromosome 7, band q21.3-q22.1. Am J Hum Genet 55, 12–20
  113. 113. Acampora, D., et al. (1999) Craniofacial, vestibular and bone defects in mice lacking the Distal-less-related gene Dlx5. Development 126, 3795–3809
  114. 114. Merlo, G.R., et al. (2002) Mouse model of split hand/foot malformation type I. Genesis 33, 97–101
  115. 115. Robledo, R.F., et al. (2002) The Dlx5 and Dlx6 homeobox genes are essential for craniofacial, axial, and appendicular skeletal development. Genes Dev 16, 1089–1101
  116. 116. Shamseldin, H.E., et al. (2012) Identification of a novel DLX5 mutation in a family with autosomal recessive split hand and foot malformation. J Med Genet 49, 16–20
  117. 117. Del Campo, M., et al. (1999) Monodactylous limbs and abnormal genitalia are associated with hemizygosity for the human 2q31 region that includes the HOXD cluster. Am J Hum Genet 65, 104–110
  118. 118. Goodman, F.R., et al. (2002) A 117-kb microdeletion removing HOXD9-HOXD13 and EVX2 causes synpolydactyly. Am J Hum Genet 70, 547–555
  119. 119. Goodman, F.R. (2002) Limb malformations and the human HOX genes. Am J Med Genet 112, 256–265
  120. 120. Sidow, A., et al. (1999) A novel member of the F-box/WD40 gene family, encoding dactylin, is disrupted in the mouse dactylaplasia mutant. Nat Genet 23, 104–107
  121. 121. de Mollerat, X.J., et al. (2003) A genomic rearrangement resulting in a tandem duplication is associated with split hand-split foot malformation 3 (SHFM3) at 10q24. Hum Mol Genet 12, 1959–1971
  122. 122. Guerrini, L., et al. (2011) A symphony of regulations centered on p63 to control development of ectoderm-derived structures. J Biomed Biotechnol 2011, 864–904
  123. 123. van Bokhoven, H., et al. (2001) p63 Gene mutations in eec syndrome, limb-mammary syndrome, and isolated split hand-split foot malformation suggest a genotype-phenotype correlation. Am J Hum Genet 69, 481–492
  124. 124. Elliott, A.M., et al. (2006) Clinical and epidemiological findings in patients with central ray deficiency: split hand foot malformation (SHFM) in Manitoba, Canada. Am J Med Genet A 140, 1428–1439
  125. 125. Rinne, T., et al. (2007) p63-associated disorders. Cell Cycle 6, 262–268
  126. 126. Ghioni, P., et al. (2005) The protein stability and transcriptional activity of p63alpha are regulated by SUMO-1 conjugation. Cell Cycle 4, 183–190
  127. 127. Elliott, A.M., et al. (2005) Split hand foot malformation (SHFM). Clin Genet 68, 501–505
  128. 128. van Bokhoven, H., and Brunner, H.G. (2002) Splitting p63. Am J Hum Genet 71, 1–13
  129. 129. Celli, J., et al. (1999) Heterozygous germline mutations in the p53 homolog p63 are the cause of EEC syndrome. Cell 99, 143–153
  130. 130. Aziz, A., et al. (2013) Novel homozygous mutations in the WNT10B gene underlying autosomal recessive split hand/foot malformation in three consanguineous families. Gene 534, 7
  131. 131. Blattner, A., et al. (2012) Homozygous nonsense mutation in WNT10B and sporadic split-hand/foot malformation (SHFM) with autosomal recessive inheritance. Am J Med Genet A 152A, 2053–2056
  132. 132. Khan, S., et al. (2012) A novel homozygous missense mutation in WNT10B in familial split-hand/foot malformation. Clin Genet 82, 48–55
  133. 133. Ugur, S.A., and Tolun, A. (2008) Homozygous WNT10b mutation and complex inheritance in Split-Hand/Foot Malformation. Hum Mol Genet 17, 2644–2653
  134. 134. Faiyaz-Ul-Haque, M., et al. (2005) Fine mapping of the X-linked split-hand/split-foot malformation (SHFM2) locus to a 5.1-Mb region on Xq26.3 and analysis of candidate genes. Clin Genet 67, 93–97
  135. 135. Berdon-Zapata, V., et al. (2004) p63 gene analysis in Mexican patients with syndromic and non-syndromic ectrodactyly. J Orthop Res 22, 1–5
  136. 136. Yang, A., et al. (1999) p63 is essential for regenerative proliferation in limb, craniofacial and epithelial development. Nature 398, 714–718
  137. 137. Mills, A.A., et al. (1999) p63 is a p53 homologue required for limb and epidermal morphogenesis. Nature 398, 708–713
  138. 138. Senoo, M., et al. (2007) p63 Is essential for the proliferative potential of stem cells in stratified epithelia. Cell 129, 523–536
  139. 139. Candi, E., et al. (2008) p63 in epithelial development. Cell Mol Life Sci 65, 3126–3133
  140. 140. Romano, R.A., et al. (2012) DeltaNp63 knockout mice reveal its indispensable role as a master regulator of epithelial development and differentiation. Development 139, 772–782
  141. 141. Mangiulli, M., et al. (2009) Identification and functional characterization of two new transcriptional variants of the human p63 gene. Nucleic Acids Res 37, 6092–6104
  142. 142. Ghioni, P., et al. (2002) Complex transcriptional effects of p63 isoforms: identification of novel activation and repression domains. Mol Cell Biol 22, 8659–8668
  143. 143. Thanos, C.D., and Bowie, J.U. (1999) p53 Family members p63 and p73 are SAM domain-containing proteins. Protein Sci 8, 1708–1710
  144. 144. Yang, A., et al. (2006) Relationships between p63 binding, DNA sequence, transcription activity, and biological function in human cells. Mol Cell 24, 593–602
  145. 145. Serber, Z., et al. (2002) A C-terminal inhibitory domain controls the activity of p63 by an intramolecular mechanism. Mol Cell Biol 22, 8601–8611
  146. 146. Brunner, H.G., et al. (2002) P63 gene mutations and human developmental syndromes. Am J Med Genet 112, 284–290
  147. 147. Brunner, H.G., et al. (2002) The p63 gene in EEC and other syndromes. J Med Genet 39, 377–381
  148. 148. van Bokhoven, H., and McKeon, F. (2002) Mutations in the p53 homolog p63: allele-specific developmental syndromes in humans. Trends Mol Med 8, 133–139
  149. 149. Lopardo, T., et al. (2008) Claudin-1 is a p63 target gene with a crucial role in epithelial development. PLoS One 3, e2715
  150. 150. Radoja, N., et al. (2007) Homeobox gene Dlx3 is regulated by p63 during ectoderm development: relevance in the pathogenesis of ectodermal dysplasias. Development 134, 13–18
  151. 151. Clements, S.E., et al. (2012) Mutations in AEC syndrome skin reveal a role for p63 in basement membrane adhesion, skin barrier integrity and hair follicle biology. Br J Dermatol 167, 134–144
  152. 152. Ferone, G., et al. (2013) p63 control of desmosome gene expression and adhesion is compromised in AEC syndrome. Hum Mol Genet 22, 531–543
  153. 153. Ferone, G., et al. (2012) Mutant p63 causes defective expansion of ectodermal progenitor cells and impaired FGF signalling in AEC syndrome. EMBO Mol Med 4, 192–205
  154. 154. Koster, M.I., et al. (2004) p63 is the molecular switch for initiation of an epithelial stratification program. Genes Dev 18, 126–131
  155. 155. Koster, M.I., and Roop, D.R. (2004) p63 and epithelial appendage development. Differentiation 72, 364–370
  156. 156. Koster, M.I., and Roop, D.R. (2007) Mechanisms regulating epithelial stratification. Annu Rev Cell Dev Biol 23, 93–113
  157. 157. Fernandez-Teran, M., and Ros, M.A. (2008) The Apical Ectodermal Ridge: morphological aspects and signaling pathways. Int J Dev Biol 52, 857–871
  158. 158. Restelli, M., et al. (2014) DLX5, FGF8 and the Pin1 isomerase control ∆Np63α protein stability during limb development: a regulatory loop at the basis of the SHFM and EEC congenital malformations. Hum Mol Genet 23, 3830–3842
  159. 159. Crackower, M.A., et al. (1998) Defect in the maintenance of the apical ectodermal ridge in the Dactylaplasia mouse. Dev Biol 201, 78–89
  160. 160. Panganiban, G. (2000) Distal-less function during Drosophila appendage and sense organ development. Dev Dyn 218, 554–562
  161. 161. Panganiban, G., and Rubenstein, J.L. (2002) Developmental functions of the Distal-less/Dlx homeobox genes. Development 129, 4371–4386
  162. 162. Beverdam, A., et al. (2002) Jaw transformation with gain of symmetry after Dlx5/Dlx6 inactivation: mirror of the past? Genesis 34, 221–227
  163. 163. Depew, M.J., et al. (2002) Specification of jaw subdivisions by Dlx genes. Science 298, 381–385
  164. 164. Kouwenhoven, E.N., et al. (2010) Genome-wide profiling of p63 DNA-binding sites identifies an element that regulates gene expression during limb development in the 7q21 SHFM1 locus. PLoS Genet 6, e1001065
  165. 165. Birnbaum, R.Y., et al. (2012) Functional characterization of tissue-specific enhancers in the DLX5/6 locus. Hum Mol Genet 21, 4930–4938
  166. 166. Birnbaum, R.Y., et al. (2012) Coding exons function as tissue-specific enhancers of nearby genes. Genome Res 22, 1059–1068
  167. 167. Lango Allen, H., et al. (2014) Next generation sequencing of chromosomal rearrangements in patients with split-hand/split-foot malformation provides evidence for DYNC1I1 exonic enhancers of DLX5/6 expression in humans. J Med Genet 51, 264–267
  168. 168. Bell, S.M., et al. (2003) Sp8 is crucial for limb outgrowth and neuropore closure. Proc Natl Acad Sci U S A 100, 12195–12200
  169. 169. Treichel, D., et al. (2003) mBtd is required to maintain signaling during murine limb development. Genes Dev 17, 2630–2635
  170. 170. Schock, F., et al. (1999) Common and diverged functions of the Drosophila gene pair D-Sp1 and buttonhead. Mech Dev 89, 125–132
  171. 171. Kawakami, Y., et al. (2004) Sp8 and Sp9, two closely related buttonhead-like transcription factors, regulate Fgf8 expression and limb outgrowth in vertebrate embryos. Development 131, 4763–4774
  172. 172. Waclaw, R.R., et al. (2006) The zinc finger transcription factor Sp8 regulates the generation and diversity of olfactory bulb interneurons. Neuron 49, 503–516
  173. 173. Ala, U., et al. (2008) Prediction of human disease genes by human-mouse conserved coexpression analysis. PLoS Comput Biol 4, e1000043
  174. 174. Dai, L., et al. (2013) Discontinuous microduplications at chromosome 10q24.31 identified in a Chinese family with split hand and foot malformation. BMC Med Genet 14, 45
  175. 175. Candi, E., et al. (2006) Differential roles of p63 isoforms in epidermal development: selective genetic complementation in p63 null mice. Cell Death Differ 13, 1037–1047
  176. 176. Marinari, B., et al. (2008) The tumor suppressor activity of IKKalpha in stratified epithelia is exerted in part via the TGF-beta antiproliferative pathway. Proc Natl Acad Sci U S A 105, 17091–17096
  177. 177. Marinari, B., et al. (2009) IKKalpha is a p63 transcriptional target involved in the pathogenesis of ectodermal dysplasias. J Invest Dermatol 129, 60–69
  178. 178. Sil, A.K., et al. (2004) IkappaB kinase-alpha acts in the epidermis to control skeletal and craniofacial morphogenesis. Nature 428, 660–664
  179. 179. Galli, F., et al. (2010) MDM2 and Fbw7 cooperate to induce p63 protein degradation following DNA damage and cell differentiation. J Cell Sci 123, 2423–2433
  180. 180. Ratovitski, E.A., et al. (2001) p53 associates with and targets Delta Np63 into a protein degradation pathway. Proc Natl Acad Sci U S A 98, 1817–1822
  181. 181. Moretti, F., et al. (2010) A regulatory feedback loop involving p63 and IRF6 links the pathogenesis of 2 genetically different human ectodermal dysplasias. J Clin Invest 120, 1570–1577
  182. 182. Browne, G., et al. (2011) Differential altered stability and transcriptional activity of DeltaNp63 mutants in distinct ectodermal dysplasias. J Cell Sci 124, 2200–2207
  183. 183. Girardini, J.E., et al. (2011) A Pin1/mutant p53 axis promotes aggressiveness in breast cancer. Cancer Cell 20, 79–91
  184. 184. Brooks, C.L., and Gu, W. (2003) Ubiquitination, phosphorylation and acetylation: the molecular basis for p53 regulation. Curr Opin Cell Biol 15, 164–171
  185. 185. Gu, B., and Zhu, W.G. (2012) Surf the post-translational modification network of p53 regulation. Int J Biol Sci 8, 672–684
  186. 186. Luo, J., et al. (2004) Acetylation of p53 augments its site-specific DNA binding both in vitro and in vivo. Proc Natl Acad Sci U S A 101, 2259–2264
  187. 187. Marmorstein, R., and Roth, S.Y. (2001) Histone acetyltransferases: function, structure, and catalysis. Curr Opin Genet Dev 11, 155–161
  188. 188. Meek, D.W., and Anderson, C.W. (2009) Posttranslational modification of p53: cooperative integrators of function. Cold Spring Harb Perspect Biol 1, a000950
  189. 189. Tang, Y., et al. (2008) Acetylation is indispensable for p53 activation. Cell 133, 612–626
  190. 190. Zeng, X., et al. (2000) The N-terminal domain of p73 interacts with the CH1 domain of p300/CREB binding protein and mediates transcriptional activation and apoptosis. Mol Cell Biol 20, 1299–1310
  191. 191. Costanzo, A., et al. (2002) DNA damage-dependent acetylation of p73 dictates the selective activation of apoptotic target genes. Mol Cell 9, 175–186
  192. 192. Mantovani, F., et al. (2004) Pin1 links the activities of c-Abl and p300 in regulating p73 function. Mol Cell 14, 625–636
  193. 193. Ianakiev, P., et al. (2000) Split-hand/split-foot malformation is caused by mutations in the p63 gene on 3q27. Am J Hum Genet 67, 59–66
  194. 194. Lu, P., et al. (2008) The apical ectodermal ridge is a timer for generating distal limb progenitors. Development 135, 1395–1405
  195. 195. Agami, R., et al. (1999) Interaction of c-Abl and p73alpha and their collaboration to induce apoptosis. Nature 399, 809–813
  196. 196. Levav-Cohen, Y., et al. (2005) C-Abl as a modulator of p53. Biochem Biophys Res Commun 331, 737–749
  197. 197. Sanchez-Prieto, R., et al. (2002) Regulation of p73 by c-Abl through the p38 MAP kinase pathway. Oncogene 21, 974–979
  198. 198. Yan, W., et al. (2008) Distinct angiogenic mediators are required for basic fibroblast growth factor- and vascular endothelial growth factor-induced angiogenesis: the role of cytoplasmic tyrosine kinase c-Abl in tumor angiogenesis. Mol Biol Cell 19, 2278–2288
  199. 199. Gros, J., et al. (2010) WNT5A/JNK and FGF/MAPK pathways regulate the cellular events shaping the vertebrate limb bud. Curr Biol 20, 1993–2002
  200. 200. Benazeraf, B., et al. (2010) A random cell motility gradient downstream of FGF controls elongation of an amniote embryo. Nature 466, 248–252
  201. 201. Pandur, P., et al. (2002) Increasingly complex: new players enter the Wnt signaling network. Bioessays 24, 881–884
  202. 202. Huelsken, J., and Behrens, J. (2002) The Wnt signalling pathway. J Cell Sci 115, 3977–3978
  203. 203. Hendriks, B., and Reichmann, E. (2002) Wnt signaling: a complex issue. Biol Res 35, 277–286
  204. 204. Clevers, H., and Nusse, R. (2012) Wnt/beta-catenin signaling and disease. Cell 149, 1192–1205
  205. 205. Church, V.L., and Francis-West, P. (2002) Wnt signalling during limb development. Int J Dev Biol 46, 927–936
  206. 206. Kengaku, M., et al. (1997) Expression of Wnt and Frizzled genes during chick limb bud development. Cold Spring Harb Symp Quant Biol 62, 421–429
  207. 207. Galceran, J., et al. (1999) Wnt3a-/--like phenotype and limb deficiency in Lef1(-/-)Tcf1(-/-) mice. Genes Dev 13, 709–717
  208. 208. Pinson, K.I., et al. (2000) An LDL-receptor-related protein mediates Wnt signalling in mice. Nature 407, 535–538
  209. 209. Takada, S., et al. (1994) Wnt-3a regulates somite and tailbud formation in the mouse embryo. Genes Dev 8, 174–189
  210. 210. Roelink, H., and Nusse, R. (1991) Expression of two members of the Wnt family during mouse development--restricted temporal and spatial patterns in the developing neural tube. Genes Dev 5, 381–388
  211. 211. Barrow, J.R., et al. (2003) Ectodermal Wnt3/beta-catenin signaling is required for the establishment and maintenance of the apical ectodermal ridge. Genes Dev 17, 394–409
  212. 212. Niemann, S., et al. (2004) Homozygous WNT3 mutation causes tetra-amelia in a large consanguineous family. Am J Hum Genet 74, 558–563
  213. 213. Grotewold, L., and Ruther, U. (2002) The Wnt antagonist Dickkopf-1 is regulated by Bmp signaling and c-Jun and modulates programmed cell death. EMBO J 21, 966–975
  214. 214. Mukhopadhyay, M., et al. (2001) Dickkopf1 is required for embryonic head induction and limb morphogenesis in the mouse. Dev Cell 1, 423–434
  215. 215. Monaghan, A.P., et al. (1999) Dickkopf genes are co-ordinately expressed in mesodermal lineages. Mech Dev 87, 45–56
  216. 216. Nohno, T., et al. (1999) Differential expression of the frizzled family involved in Wnt signaling during chick limb development. Cell Mol Biol (Noisy-le-grand) 45, 653–659
  217. 217. Zallen, J.A. (2007) Planar polarity and tissue morphogenesis. Cell 129, 1051–1063
  218. 218. Boutros, M., et al. (1998) Dishevelled activates JNK and discriminates between JNK pathways in planar polarity and wingless signaling. Cell 94, 109–118
  219. 219. Li, L., et al. (1999) Dishevelled proteins lead to two signaling pathways. Regulation of LEF-1 and c-Jun N-terminal kinase in mammalian cells. J Biol Chem 274, 129–134
  220. 220. McEwen, D.G., and Peifer, M. (2000) Wnt signaling: Moving in a new direction. Curr Biol 10, R562–564
  221. 221. Keller, R. (2002) Shaping the vertebrate body plan by polarized embryonic cell movements. Science 298, 1950–1954
  222. 222. Gong, Y., et al. (2004) Planar cell polarity signalling controls cell division orientation during zebrafish gastrulation. Nature 430, 689–693
  223. 223. Heisenberg, C.P., et al. (2000) Silberblick/Wnt11 mediates convergent extension movements during zebrafish gastrulation. Nature 405, 76–81
  224. 224. Jessen, J.R., et al. (2002) Zebrafish trilobite identifies new roles for Strabismus in gastrulation and neuronal movements. Nat Cell Biol 4, 610–615
  225. 225. Yin, C., et al. (2008) Cooperation of polarized cell intercalations drives convergence and extension of presomitic mesoderm during zebrafish gastrulation. J Cell Biol 180, 221–232
  226. 226. Kuhl, M., et al. (2000) The Wnt/Ca2+ pathway: a new vertebrate Wnt signaling pathway takes shape. Trends Genet 16, 279–283
  227. 227. Kuhl, M., et al. (2000) Ca(2+)/calmodulin-dependent protein kinase II is stimulated by Wnt and Frizzled homologs and promotes ventral cell fates in Xenopus. J Biol Chem 275, 12701–12711
  228. 228. Sheldahl, L.C., et al. (1999) Protein kinase C is differentially stimulated by Wnt and Frizzled homologs in a G-protein-dependent manner. Curr Biol 9, 695–698
  229. 229. Sheldahl, L.C., et al. (2003) Dishevelled activates Ca2+ flux, PKC, and CamKII in vertebrate embryos. J Cell Biol 161, 769–777
  230. 230. Eisenberg, L.M., et al. (1992) Cloning and characterization of a novel Drosophila Wnt gene, Dwnt-5, a putative downstream target of the homeobox gene distal-less. Dev Biol 154, 73–83
  231. 231. Rao, T.P., and Kuhl, M. (2010) An updated overview on Wnt signaling pathways: a prelude for more. Circ Res 106, 1798–1806
  232. 232. van Amerongen, R., and Nusse, R. (2009) Towards an integrated view of Wnt signaling in development. Development 136, 3205–3214
  233. 233. Kilian, B., et al. (2003) The role of Ppt/Wnt5 in regulating cell shape and movement during zebrafish gastrulation. Mech Dev 120, 467–476
  234. 234. Wallingford, J.B., et al. (2001) Regulation of convergent extension in Xenopus by Wnt5a and Frizzled-8 is independent of the canonical Wnt pathway. Int J Dev Biol 45, 225–227
  235. 235. Kawakami, Y., et al. (1999) Involvement of Wnt-5a in chondrogenic pattern formation in the chick limb bud. Dev Growth Differ 41, 29–40
  236. 236. Topol, L., et al. (2003) Wnt-5a inhibits the canonical Wnt pathway by promoting GSK-3-independent beta-catenin degradation. J Cell Biol 162, 899–908
  237. 237. Slusarski, D.C., et al. (1997) Modulation of embryonic intracellular Ca2+ signaling by Wnt-5A. Dev Biol 182, 114–120
  238. 238. Slusarski, D.C., et al. (1997) Interaction of Wnt and a Frizzled homologue triggers G-protein-linked phosphatidylinositol signalling. Nature 390, 410–413
  239. 239. Veeman, M.T., et al. (2003) A second canon. Functions and mechanisms of beta-catenin-independent Wnt signaling. Dev Cell 5, 367–377
  240. 240. Westfall, T.A., et al. (2003) Wnt-5/pipetail functions in vertebrate axis formation as a negative regulator of Wnt/beta-catenin activity. J Cell Biol 162, 889–898
  241. 241. Mikels, A.J., and Nusse, R. (2006) Purified Wnt5a protein activates or inhibits beta-catenin-TCF signaling depending on receptor context. PLoS Biol 4, e115
  242. 242. Andre, P., et al. (2012) The Wnt coreceptor Ryk regulates Wnt/planar cell polarity by modulating the degradation of the core planar cell polarity component Vangl2. J Biol Chem 287, 44518–44525
  243. 243. Logan, C.Y., and Nusse, R. (2004) The Wnt signaling pathway in development and disease. Annu Rev Cell Dev Biol 20, 781–810
  244. 244. Sokol, S.Y., and Wharton, K.A., Jr. (2007) WNTers in La Jolla. Development 134, 3393–3399
  245. 245. Zylstra, C.R., et al. (2008) Gene targeting approaches in mice: assessing the roles of LRP5 and LRP6 in osteoblasts. J Musculoskelet Neuronal Interact 8, 291–293
  246. 246. Yang, Y. (2003) Wnts and wing: Wnt signaling in vertebrate limb development and musculoskeletal morphogenesis. Birth Defects Res C Embryo Today 69, 305–317
  247. 247. Patton, M.A., and Afzal, A.R. (2002) Robinow syndrome. J Med Genet 39, 305–310
  248. 248. Person, A.D., et al. (2010) WNT5A mutations in patients with autosomal dominant Robinow syndrome. Dev Dyn 239, 327–337
  249. 249. Roifman, M., et al. (2015) De novo WNT5A-associated autosomal dominant Robinow syndrome suggests specificity of genotype and phenotype. Clin Genet 87, 34–41
  250. 250. Jeffery, A. (2003) Insulin resistance. Nurs Stand 17, 47–53; quiz 54–45
  251. 251. Oldridge, M., et al. (2000) Dominant mutations in ROR2, encoding an orphan receptor tyrosine kinase, cause brachydactyly type B. Nat Genet 24, 275–278
  252. 252. Yamaguchi, T.P., et al. (1999) A Wnt5a pathway underlies outgrowth of multiple structures in the vertebrate embryo. Development 126, 1211–1223
  253. 253. Yang, Y., et al. (2003) Wnt5a and Wnt5b exhibit distinct activities in coordinating chondrocyte proliferation and differentiation. Development 130, 1003–1015
  254. 254. Gao, B., et al. (2011) Wnt signaling gradients establish planar cell polarity by inducing Vangl2 phosphorylation through Ror2. Dev Cell 20, 163–176
  255. 255. Mikels, A., et al. (2009) Ror2 receptor requires tyrosine kinase activity to mediate Wnt5A signaling. J Biol Chem 284, 30167–30176
  256. 256. Oishi, I., et al. (2003) The receptor tyrosine kinase Ror2 is involved in non-canonical Wnt5a/JNK signalling pathway. Genes Cells 8, 645–654
  257. 257. Hikasa, H., et al. (2002) The Xenopus receptor tyrosine kinase Xror2 modulates morphogenetic movements of the axial mesoderm and neuroectoderm via Wnt signaling. Development 129, 5227–5239
  258. 258. Nomi, M., et al. (2001) Loss of mRor1 enhances the heart and skeletal abnormalities in mRor2-deficient mice: redundant and pleiotropic functions of mRor1 and mRor2 receptor tyrosine kinases. Mol Cell Biol 21, 8329–8335
  259. 259. Schwabe, G.C., et al. (2004) Ror2 knockout mouse as a model for the developmental pathology of autosomal recessive Robinow syndrome. Dev Dyn 229, 400–410
  260. 260. Halford, M.M., et al. (2000) Ryk-deficient mice exhibit craniofacial defects associated with perturbed Eph receptor crosstalk. Nat Genet 25, 414–418
  261. 261. Wang, B., et al. (2011) Disruption of PCP signaling causes limb morphogenesis and skeletal defects and may underlie Robinow syndrome and brachydactyly type B. Hum Mol Genet 20, 271–285
  262. 262. Paina, S., et al. (2011) Wnt5a is a transcriptional target of dlx homeogenes and promotes differentiation of interneuron progenitors in vitro and in vivo. J Neurosci 31, 2675–2687
  263. 263. Depew, M.J., et al. (2005) Reassessing the Dlx code: the genetic regulation of branchial arch skeletal pattern and development. J Anat 207, 501–561
  264. 264. Vieux-Rochas, M., et al. (2010) Spatio-temporal dynamics of gene expression of the Edn1-Dlx5/6 pathway during development of the lower jaw. Genesis 48, 262–373

Written By

Daniele Conte, Luisa Guerrini and Giorgio R. Merlo

Submitted: 29 September 2014 Reviewed: 12 February 2015 Published: 21 October 2015