pangolin


DEVELOPMENTAL BIOLOGY

Embryonic

pangolin is already found in embryos of 0-2 hours, indicative of maternal expression (van de Wetering, 1997). pangolin is expressed ubiquitously throughout embryogenesis. By the end of germ band retraction, pan transcriptions are arrayed in a segmentally modulated pattern, with high accumulation in the visceral mesoderm and endoderm of the gut, as well as in part of the procephalon (Brunner, 1997).

Pangolin function during wing development: Nemo antagonizes Wingless signaling

The cellular events that govern patterning during animal development must be precisely regulated. This is achieved by extrinsic factors and through the action of both positive and negative feedback loops. Wnt/Wg signals are crucial across species in many developmental patterning events. Drosophila nemo (nmo) acts as an intracellular feedback inhibitor of Wingless (Wg) and it is a novel Wg target gene. Nemo antagonizes the activity of the Wg signal, as evidenced by the finding that reduction of nmo rescues the phenotypic defects induced by misexpression of various Wg pathway components. In addition, the activation of Wg-dependent gene expression is suppressed in wing discs ectopically expressing nmo and enhanced cell autonomously in nmo mutant clones. nmo itself is a target of Wg signaling in the imaginal wing disc. nmo expression is induced upon high levels of Wg signaling and can be inhibited by interfering with Wg signaling. Finally, alterations are observed in Arm stabilization upon modulation of Nemo. These observations suggest that the patterning mechanism governed by Wg involves a negative feedback circuit in which Wg induces expression of its own antagonist Nemo (Zeng, 2004).

In Drosophila, several examples of Wg feedback inhibition have been identified. (1) It has been shown that Wg downregulates its own transcription in the wing pouch to narrow the RNA expression domain at the DV boundary. (2) Wg signaling can repress the expression of its receptor Dfz2 in the wg-expressing cells of the wing disc. Wg regulation of Dfz2 creates a negative feedback loop in which newly secreted Wg is stabilized only once it moves away from the DV boundary to cells expressing higher levels of Drosophila Fz2. (3) The Wg target gene naked cuticle (nkd) acts through Dsh to limit Wg activity. (4) Wingful (Wf), an extracellular inhibitor of Wg, is itself induced by Wg signaling (Zeng, 2004).

This research adds Nemo to this list of inducible antagonists participating in Wg signaling. Nemo antagonizes the Wg signal in wing development, as evidenced by phenotypic rescue, suppression of Wg-dependent gene expression in discs ectopically expressing nmo, and ectopic expression of a Wg-dependent gene in nmo mutant clones (Zeng, 2004).

In support that Wg signaling regulates the transcription of nmo, several dTCF consensus binding sites have been found in the 5' region of the nmo gene that may represent enhancer elements. Indeed, two sites match 9 out of 11 bp (GCCTTTGAT) of the T1 site (GCCTTTGATCT) in the dpp BE enhancer that has been shown both in vitro and in vivo to bind and respond to dTCF. The presence of these sites suggests that the observed transcriptional regulation of nmo by Wg may involve direct binding to the nmo DNA sequence by dTCF (Zeng, 2004).

As a result of comparing the endogenous expression pattern of nmo with stabilized Arm, it was noticed that the highest levels of Nemo exclude Arm stabilization, while high levels of Arm are present in cells in which nmo levels are lower. Since Arm protein stabilization is a direct consequence of Wg pathway activation, attempts were made to examine whether Nemo may function to inhibit Wg by promoting Arm destabilization and subsequent breakdown. Indeed, ectopic expression of Nemo can lead to cell-autonomous reduction in Arm protein levels. This preliminary result suggests a mechanism in which Nemo may contribute to the destabilization of Arm that involves the Axin/APC/GSK3 complex. One explanation to account for such a finding would concern the interaction with TCF in the nucleus and the role of dTCF as an anchor for Arm. Given what is known about NLKs, it is likely that Nemo may act on the ability of the dTCF/Arm complex to bind DNA and activate transcription. It has been proposed that dTCF acts as an anchor for Arm in the nucleus. It remains to be determined how efficient this anchor is and whether there are conditions in which the interaction may become compromised, such as is seen with elevated Nemo. NLKs have been shown to affect the DNA-binding ability of TCF/ß-catenin. Perhaps in the absence of DNA binding, this complex is less stable and Arm could be free to shuttle to the cytoplasm where it could associate with Axin or APC and become degraded. It is proposed that the ectopic nmo leads to destabilization of the dTCF/Arm/DNA complex, thus causing Arm to exit the nucleus and be degraded through interaction with Axin, APC and GSK3. The observation that ectopic expression of full-length Arm cannot induce any activated Wg phenotypes has been explained by the hypothesis that even these high levels of protein are not sufficient to overcome the degradation machinery. Thus, the finding that there is no elevated Arm in nmo clones is consistent with an inability to overcome the endogenous degradation machinery; even though less Nemo could lead to more stabilized DNA interactions, this would not lead to higher levels of stabilized Arm than are normally found (Zeng, 2004).

Developmental roles of the Mi-2/NURD-associated protein p66 in Drosophila: Overexpression of p66 can repress activation of a TCF reporter by ß-catenin

The NURD and Sin3 histone deacetylase complexes are involved in transcriptional repression through global deacetylation of chromatin. Both complexes contain many different components that may control how histone deacetylase complexes are regulated and interact with other transcription factors. In a genetic screen for modifiers of wingless signaling in the Drosophila eye, mutations were isolated in the Drosophila homolog of p66, a protein previously purified as part of the Xenopus NURD/Mi-2 complex. p66 encodes a highly conserved nuclear zinc-finger protein that is required for development and it is proposed that the p66 protein acts as a regulatory component of the NURD complex. Animals homozygous mutant for p66 display defects during metamorphosis possibly caused by misregulation of ecdysone-regulated expression. Although heterozygosity for p66 enhances a wingless phenotype in the eye, loss-of-function clones in the wing and the eye discs do not have any detectable phenotype, possibly due to redundancy with the Sin3 complex. Overexpression of p66, in contrast, can repress wingless-dependent phenotypes. Furthermore, p66 expression can repress multiple reporters in a cell culture assay, including a Wnt-responsive TCF reporter construct, implicating the NURD complex in repression of Wnt target genes. By co-immunoprecipitation, p66 associates with dMi-2, a known NURD complex member (Kon, 2005).

Thus, loss of p66 enhances the sev-wgts phenotype in the eye; i.e., the resulting phenotype resembles that of increased wingless signaling. Since loss of p66 does not affect other transgenes expressed using the same promoter as sev-wgts, it is believed that removal of p66 does not affect wingless expression through the sevenless promoter, but instead represses wingless target genes involved in bristle formation. The results are in agreement with experiments that have implicated the NURD complex in Wnt signaling. It has been reported that Lef-1 repression involves HDAC-1 function. This interaction takes place in the absence of mSin3A, leading to the hypothesis that Lef-1 repression involves recruitment of the NURD complex. Three pieces of evidence are presented to further support this hypothesis: (1) loss of p66 enhances a wg overexpression phenotype; (2) overexpression of p66 can repress activation of a TCF reporter by ß-catenin in tissue culture cells; (3) overexpression of p66 in vivo represses the formation of wingless-dependent scutellar bristles without repressing wingless expression. Together, these experiments provide additional evidence that the NURD complex is involved in repression of Wnt target genes (Kon, 2005).

The lethality of p66 mutants is caused by misregulation of ecdysone-regulated genes during larval stages. If p66 was involved in repression of ecdysone-induced targets such as E74 and DHR3, then loss of p66 should lead to ectopic gene expression. In contrast, it was found that ecdysone-induced genes are not activated, suggesting that the role of p66 is more complex and indirect. In microarray studies of ecdysone response, 44% of genes that changed expression were repressed. It is speculated that repression of one or more of these genes is required for ecdysone-induced expression and that p66 is required for this process (Kon, 2005).

An additional connection between the NURD complex and ecydsone response is made through Bonus (bon), the Drosophila homolog of TIF1. TIF1 was identified as a protein that interacts with HP1, a heterochromatin-associated protein. The NURD complex may be involved in histone modification to allow HP1 binding. Similar to p66 mutants, bon mutants die during pupal stages due to misregulation of ecdysone-induced genes. Furthermore, E74 expression is also reduced in bon mutants. Thus it is speculated that p66 and the NURD complex may be involved in regulation of ecdysone response through HP1-mediated repression (Kon, 2005).

p66 mutations can affect both wg and ecdysone-induced gene expression. Furthermore, in cell culture reporter assays, it was found that expression of p66 inhibits activation of a TCF reporter by ß-catenin, of an SRE reporter by the M1 receptor, and of an NF-AT reporter by high levels of intracellular calcium. In addition, two human p66 homologs, hp66alpha and hp66ß, function as transcriptional repressors when tethered to a promoter, suggesting that transcriptional repression is a shared activity of p66 proteins. However, p66 did not repress an albumin luciferase reporter, implying that expression of p66 does not cause a defect in general transcription. Therefore, it is concluded that p66 can repress multiple signaling pathways, and it is hypothesized that this repression is mediated by recruitment of the NURD histone deacetylase complex. This conclusion is also supported by previous reports that expression of p66 can change the localization of MBD3, a component of the NURD complex (Kon, 2005).

Although overexpression of p66 can repress multiple signaling pathways, no loss-of-function phenotypes were detected that would be consistent with this hypothesis. In Drosophila, relatively mild and tissue-specific phenotypes for repressors have also been found for Pangolin/dTCF and naked cuticle. With respect to Pangolin/dTCF, this could be due to its dual role as both a repressor and an activator, but in other cases, a lack of phenotypes is possibly due to redundancy among parallel pathways (not necessarily among related genes). This suggestion is in analogy to the SynMuv genes in C. elegans. Animals mutant in either a synMuvA or a synMuvB gene alone have a normal vulva; animals mutant for both a synMuvA gene and a synMuvB gene have a multi-vulval (Muv) phenotype. Indeed, in the C. elegans vulva, p66 functions in a redundant pathway (Kon, 2005).

P66 is present when NURD complex is purified from Xenopus oocytes and is found to be associated with the NURD complex in mammalian cells in association with the methyl DNA-binding protein MBD2, as part of the MeCP1 complex. Is p66 a component of the NURD complex or an accessory factor? WdMi-2 and p66 co-purify, suggesting that they form a complex. However, if all of the complex members participate in the same processes, then the corresponding mutants should also have the same phenotypes. To ascertain whether p66 and rpd3 (HDAC homolog) function together is difficult since rpd3 also participates in the Sin3 complex. Thus, it is likely that rpd3 will display a wider range of phenotypes than a mutant of the NURD complex alone (Kon, 2005).

However, the p66 mutant can be compared to the other NURD complex mutant characterized in Drosophila, dMi-2. dMi-2 is required for oogenesis and for cell viability. rpd3 is also likely important for oogenesis; a hypomorphic rpd3 allele produces very few eggs. In contrast, p66 mutant germlines produce normal embryos, and p66 mutant clones survive in third instar imaginal discs. dMi-2 zygotic mutants die during larval stages, which coincides with the lethality of strong allele combination of p66 mutants. It is possible that dMi-2 may also be required for ecdysone response in larval stages. Since the p66 mutant phenotype does not completely mimic Mi-2 mutant phenotypes, it is suggested that p66 is not a core component of the NURD complex, but could have a regulatory function. However, it is equally possible that differences in allelic strength or perdurance of maternal contributions obscures the full range of phenotypes on the several components (Kon, 2005).

These results are further supported by experiments demonstrating that dMi-2 and the Drosophila MBD2/3 protein do not colocalize in nuclei. dMi-2 is distributed ubiquitously in embryonic nuclei, while Drosophila MBD2/3 is localized in a speckled pattern. This result suggests that Drosophila MBD2/3 is not an integral component of all dMi-2 complexes (Kon, 2005).

How might p66 function? In mammalian cells, human p66 protein is associated only with the NURD complex as part of the MeCP1 complex, which additionally contains methylated DNA-binding activity through the MBD2 protein. Similarly, the Xenopus NURD complex, which copurifies with p66, in contrast to mammalian NURD complexes, also has methylated DNA-binding activity. The NURD complex, through Mi-2, interacts with the zinc-finger proteins Hunchback, Tramtrak, and Ikaros. It is hypothesized that p66, also a zinc-finger-containing protein, functions similarly to these transcription factors to recruit the NURD complex to methylated DNA (Kon, 2005).

This hypothesis is supported by experiments that demonstrate that human p66 interacts with MBD2 and MBD3 in vitro. Furthermore, overexpression of p66 can change localization of MBD3. On the basis of these results, it is hypothesized that p66 may function as a link between the NURD complex and the methylated DNA-binding proteins MBD2 and MBD3. It is proposed that p66 recruits the NURD complex to mediate methylation-mediated silencing. Since both DNA methyl transferases and DNA-methylated binding proteins exist in Drosophila, it is likely that there is some vestige of a methylation system in Drosophila, although the function is unknown at this time (Kon, 2005).

Effects of Mutation or Deletion

A screening was carried out to identify components of the wingless signal transduction pathway; the screen sought dominant suppressors of the rough eye phenotype, caused by a transgene that drives ectopic expression of wingless during eye development. Essentially such a screen would identify cellular components that when mutated would fail to transduce signals from ectopically produced Wingless. Four strong suppressors were found that fall into two complementation groups. Two are recessive alleles of armadillo and two are recessive alleles of a locus on the fourth chromosome, designated as pangolin. A stronger allele of pan was isolated which shows strong genetic interactions when trans-heterozygous with either of the two isolated armadillo mutations, causing adult phenotypes characteristic of reduced wg activity. These results suggest that pan, like arm, encodes a component of the wingless transduction pathway, and also raise the possiblity that these components may physically interact (Brunner, 1997).

Adults mutant for pan show crippled or absent antennae and legs, and the wing is transformed into a duplicated notum. There is a loss of dorso-ventral polarity in the leg pattern. Wing notches are found in 20% to 30% of adults carrying one allele. The abdominal structure is altered with sternite structures affected in all animals. Typically, sternite bristles are lost and sternite tissues are replaced by pleura. Thus embryos mutant for pan display a segment polarity phenotype in which the larval epidermis forms a lawn of ventral denticles and lacks naked cuticle between the segmental denticle belts, similar to armadillo mutants (Brunner, 1997).

A constitutively active form of Armadillo was expressed in wild type flies and results in a naked phenotype in which the denticles normally positioned in the anterior portion of each segment are replaced by naked cuticle. In pan mutants no naked cuticle due to active Armadillo overexpression is observed, thus showing that in the absence of PAN, ARM activity is blocked and cannot cause a biological wingless response in vivo (Brunner, 1997).

An N-terminal deletion mutant of pangolin was ectopically expressed in flies. Ubiquitous expression of pangolin results in a segment polarity phenotype. In the most severe transgenics, the denticle belts are also narrowed in the dorsal-ventral axis. A decay in the stripes of engrailed expression is apparent in these mutants during late stage 9 (van de Wetering, 1997).

When a constitutively active from of armadillo is expressed in pangolin mutants, its action is largely inhibited. Such double mutants resemble pangolin mutants in that they have alternating denticles and naked cuticle, with portions of the lateral naked cuticle converted to denticles. However, they often have regions of naked cuticle intruding into the normal denticle belt at the ventral midline (van de Wetering, 1997).

The mutation ciD is mutant for two neighboring loci, cubitus interruptus and pan. In situ hybridication experiments with pan probes reveal that in ciD/+ heterozygote embryos, pan transcripts are detected in a pattern indistinguishable from that of ci transcripts, providing additional evidence that the molecular lesion of ciD disrupts both genes. It is a curious coincidence that pan and ci are adjacent genes, as ci encodes a transcription factor that is essential for transducing all examples of Hedgehog signaling, whereas the present evidence suggests an equivalent role for Pan in Wingless signal transduction. The fact that the ciD mutation abolishes both activities also calls for a reassessment of genetic epistasis experiments in which this allele is used to assay the relationship between Hedgehog and Wingless signaling (Brunner, 1997)

The Hedgehog (Hh) and Wingless (Wg) signaling pathways play important roles in animal development. The activities of the two pathways depend on each other during Drosophila embryogenesis. In the embryonic segment, Wg is required in anterior cells to sustain Hh secretion in adjacent posterior cells. In turn, Hh input is necessary for anterior cells to maintain wg expression. The Hh and Wg pathways are mediated by the transcription factors Cubitus interruptus (Ci) and Pangolin/TCF (Pan), respectively. Coincidentally, pan and ci are adjacent genes on the fourth chromosome in a head-to-head orientation. Genetic and in situ hybridization data indicate that ciD is a mutation affecting both ci and pan. Whereas pan is expressed ubiquitously during embryogenesis, in ciD mutant embryos pan is expressed in an intense, segmentally-repeated manner. This expression is intriguingly similar to that of the ci gene and it occurs both in homozygous and heterozygous mutant embryos. Moreover, it was found that in ciD/ciD homozyogous mutants, the typical striped expression of ci is absent; instead, ci transcripts are dispersed uniformly at low levels, like pan transcripts in wild type embryos. Molecular analysis reveals that the ciD allele is caused by an inversion event that swaps the promoter regions and the first exons of the two genes. The ci gene in ciD is controlled by the ubiquitous pan promoter and encodes a hybrid Ci protein that carries the N-terminal region of Pan. The predicted Ci fusion protein product consists of the first 246 amino acids of Pan fused in frame to the Ci protein, of which the first 13 N-terminal amino acids are missing. The N-terminal domain of Pan has previously been shown to bind to the beta-catenin homolog Armadillo (Arm), raising the possibility that Wg input, in addition to Hh input, modulates the activity of the hybrid CiD protein. Indeed, Wg signaling induces the expression of the Hh target gene patched (ptc) in ciD animals. Evidence is provided that this effect depends on the ability of the CiD protein to bind Arm. Genetic and molecular data indicate that wild-type Pan and CiD compete for binding to Arm, leading to a compromised transduction of the Wg signal in heterozygous ciD/+ animals and to a dramatic enhancement of the gain-of-function activity of ciD in homozygous mutants. Thus, the Hh and the Wg pathways are affected by the ciD mutation, and the CiD fusion protein integrates the activities of both (Schweizer, 1998).

It has been proposed that during embryogenesis ciD functions as a gain-of-function allele of ci. This is largely based on the striking finding that CiD can substute for Hh protein in driving expression of Hh-responsive genes. Specifically, in homozygous ciD mutant embryos, the ptc, wg and gooseberry genes are expressed in wider stripes, when compared to wild type, even in the absence of active Hh protein. However, homozygous ciD mutant embryos can be rescued to adult by introducing a duplication of the pan/ci genomic region (unpublished). Thus it is unlikely that the observed phenotypes can be ascribed merely to the dosage of CiD protein. Rather, it appears to be critical that no wild-type Pan protein accompanies the CiD protein in ciD homozygous embryos. This allows all Arm protein that accumulates in response to Wg to be unrestrained to bind to, and activate the CiD protein. It is predicted that the expression of an N-terminally complete Pan protein -- even if its DNA-binding activity has been impaired -- would abolish the extreme gain-of-function effect of CiD in homozygous mutant embryos. Thus it must be the simultaneous gain and loss of Arm binding sites by Ci and Pan, respectively, that confers these unusual properties to the ciD allele. Together with the observation that CiD interferes in a dominant-negative manner with Wg signaling, these results illustrate that the levels of free and accessible Arm protein critically determine the output of Wg signaling in wild-type and the output of CiD signaling in mutant situations (Schweizer, 1998 and references).

The signaling mediated by Arm requires not only its cytoplasmic accumulation but also the activity of the DNA-binding protein Drosophila TCF known as Pangolin (Pan). Arm directly interacts with Pangolin through its central Armadillo repeats, a region that is critical in Arm's ability to induce cell death. Pangolin function in an APC-like mutant background cannot be completely eliminated; however, flies heterozygous for two mutant alleles of pangolin, panciD and pan13, were examined to determine if Pangolin is required in the activation of cell death that results from Apc loss. panciD is mutant for two loci, ci and pan; it behaves as both a hypomorphic allele of ci and a null allele of pan. In flies heterozygous for panciD, some retinal neurons are rescued from cell death in the Apc mutant. As was seen with ectopic Zw3 expression, the rescued cells are detected solely at the apical surface of the eye. The same rescue is found in Apc mutants that are heterozygous for the null allele pan13. These findings suggest that elevated Arm levels activate a cell death pathway via the Arm/Pangolin complex (Ahmed, 1998).

The Drosophila cubitus interruptus (ci) gene encodes a sequence-specific DNA-binding protein that regulates transcription of Hedgehog (Hh) target genes. Activity of the Ci protein is posttranslationally regulated by Hh signaling. In animals homozygous for the ciD mutation, however, transcription of Hh target genes is regulated by Wingless (Wg) signaling rather than by Hh signaling. ciD is shown to encode a chimeric protein composed of the regulatory domain of dTCF/Pangolin (Pan) and the DNA binding domain of Ci. Pan is a Wg-regulated transcription factor that is activated by the binding of Armadillo (Arm) to its regulatory domain. Arm is thought to activate Pan by contributing a transactivation domain. A constitutively active form of Arm potentiates activity of a CiD transgene and coimmunoprecipitates with CiD protein. The Wg-responsive activity of CiD could be explained by recruitment of the Arm transactivation function to the promoters of Hh-target genes. It is suggestrd that wild-type Ci also recruits a protein with a transactivation domain as part of its normal mechanism of activation (Von Ohlen, 1999).

Epistasis analysis was carried out to position Adenomatous polypopsis coli tumor suppressor homolog 2 (Apc2) with respect to other components of the signal transduction pathway. wg; Apc2DeltaS double mutant embryos (with Apc2DeltaS mutant mothers) show a partial rescue of the wg phenotype, with restoration of the normal diversity of cuticular pattern elements and small expanses of naked cuticle, suggesting that Apc2 is downstream of wg. There are two possible explanations for the fact that the double mutant does not show the same phenotype as the Apc2 single mutant: either Apc2DeltaS is not null, or the negative regulatory machinery remains partially active in the absence of Apc2. If Apc2DeltaS is not null, it was reasoned that repeating the epistasis test with Apc2DeltaS in trans to a deficiency removing Apc2 (Df(3R)crb87-4) might further reduce Apc2 function, producing a double mutant phenotype more similar to that of Apc2DeltaS alone. However, when this was done, there was no change in the double mutant phenotype, suggesting that Apc2DeltaS may be genetically null for this function. Other components of the Wg signal transduction pathway act downstream of Apc2. Embryos maternally and zygotically mutant for both dishevelled (dsh) and Apc2 show a phenotype indistinguishable from the dsh single mutant, as do embryos maternally mutant for both dsh and Apc2 that are zygotically dsh/Y; Apc2DeltaS/Df(3R)crb87-4. Likewise, arm; Apc2 and Apc2; dTCF double mutants (derived from Apc2 homozygous mothers) are indistinguishable from arm or dTCF single mutants. Thus, dsh, arm, and dTCF all act genetically downstream of Apc2; this was expected for arm and dTCF, but was surprising for dsh (McCartney, 1999).

The murine Nemo homolog Nlk has been implicated in regulating Wnt signaling by repressing the Arm/ß-Catenin-TCF complex, a component of Wingless signaling. Whether such an interaction occurs in flies during wing vein formation was examined. The extra vein phenotypes observed in nmo mutant wings are very similar to those that have been previously described as resulting from overexpression of Armadillo and both vertebrate ß-catenin and plakoglobin in the wing. In addition, ectopic veins are produced as a result of ectopic wg and dsh expression. Constitutively active Armadillo (UAS-Arms10) expressed using the 1348-Gal4 driver leads to the formation of moderate ectopic veins emanating from the PCV, in addition to more severe ectopic veins along LII. These are both regions of the wing that are sensitive to nmo mutations and where similar ectopic veins are observed in nmoadk. The phenotypes seen with overexpression of wg and arm are consistent with the theory that Nemo is a negative regulator of Wingless signaling since loss of nemo mimics extra veins seen with overexpression of arm and wg (Verheyen, 2001).

Overexpression of both wildtype mouse Lef-1 (a TCF homolog) and a constitutive repressor form of dTCF (Pangolin) results in dominant negative phenotypes. The constitutive repressor form of dTCF (UAS-dTCFDeltaN) is unable to bind Arm and represses wg-dependent gene expression. These findings suggest that expression of wildtype dTCF (UAS-dTCFwt) somehow interferes with a wingless-targeted transcription factor in a dominant negative way. Consistent with this, it is found that ectopic expression of UAS-dTCFwt using vestigial-Gal4 results in defects in the posterior wing margin, a phenotype seen with loss of wg signaling. Ectopic expression of the UAS-dTCFDeltaN using the 1348-Gal4 driver is lethal. However, homozygosity for nmoadk is able to rescue the lethality and the flies that emerged had reduced ectopic wing veins. This finding can be interpreted by taking into account the dual roles TCF plays in the nucleus. In nmoadk mutants, the negative regulation of endogenous dTCF may be reduced, leading to more wg-dependent signaling and the induction of extra veins similar to those seen with constitutive Arm expression. Expression of UAS-dTCFDeltaN in the nmoadk background most likely interferes with the de-repressed endogenous dTCF and block the induction of extra veins seen in nmoadk (Verheyen, 2001).

Newly eclosed flies have wings that are highly folded and compact. Within an hour, each wing has expanded, the dorsal and ventral cuticular surfaces bonding to one another to form the mature wing. To initiate a dissection of this process, two mutant phenotypes were examined. (1) The batone (bae) mutant blocks wing expansion, a behavior that is shown to have a mutant focus anterior to the wing in the embryonic fate map. (2) Ectopic expression of protein kinase A catalytic subunit (PKAc) using certain GAL4 enhancer detector strains mimics the batone wing phenotype and also induces melanotic 'tumors'. Surprisingly, these GAL4 strains express GAL4 in cells, which seem to be hemocytes, found between the dorsal and ventral surfaces of newly opened wings. Ectopic expression of Ricin A in these cells reduces their number and prevents bonding of the wing surfaces without preventing wing expansion. It is proposed that hemocytes are present in the wing to phagocytose apoptotic epithelial cells and to synthesize an extracellular matrix that bonds the two wing surfaces together. Hemocytes are known to form melanotic tumors either as part of an innate immune response or under other abnormal conditions, including evidently ectopic PKAc expression. Ectopic expression of PKAc in the presence of the batone mutant causes dominant lethality, suggesting a functional relationship. It is proposed that batone is required for the release of a hormone necessary for wing expansion and tissue remodeling by hemocytes in the wing (Kiger, 2001).

Comparison of the effects of Ricin A and of PKAc on wing maturation indicates that ectopic PKAc does not simply inactivate hemocytes. Instead, it appears to substitute one normal function of hemocytes for another. Rather than carry out phagocytosis and ECM synthesis, hemocytes enter into an innate immune response in which lamellocytes are differentiated and crystal cells melanize target cells. Evidently, aggregation of lamellocytes within the wing blade interferes with wing expansion, and loss of normal hemocyte function interferes with bonding of dorsal and ventral surfaces. The observation that the effect of ectopic PKAc on the wing is suppressed by overexpression of Pan, the Drosophila homolog of mammalian blood cell transcription factors (lymphocyte enhancer-binding factor 1 and T cell factor), suggests that ectopic PKAc inhibits (or represses synthesis of) Pan, which in turn inhibits Wingless target gene expression. This conclusion is strengthened by the observation that ectopic expression of UAS-dCBP(nej+) using GAL4-30A produces phenotypes similar to those caused by ectopic PKAc. Pan is bound and its transcriptional activity inhibited by dCBP. Expression of PanDeltaN, a dominant-negative inhibitor of Wingless target gene expression, elicits what seems to be a massive induction of the cellular innate immune response. Thus, the Wingless signal transduction pathway may be involved in regulating a choice between the innate immune response and the apoptotic/ECM response (Kiger, 2001).

The dominant-lethal interaction between ectopic PKAc and bae is intriguing. When and how death occurs needs closer examination, as does the cellular focus of bae activity. What role PKAc normally plays in regulating hemocyte behavior remains to be investigated. The association of a wing phenotype with altered hemocyte behavior should provide a means of identifying additional genes involved in hemocyte function during wing maturation (Kiger, 2001).

Cellular interaction between the proximal and distal domains of the limb plays key roles in proximal-distal patterning. In Drosophila, these domains are established in the embryonic leg imaginal disc as a proximal domain expressing escargot, surrounding the Distal-less expressing distal domain in a circular pattern. The leg imaginal disc is derived from the limb primordium that also gives rise to the wing imaginal disc. Essential roles of Wingless in patterning the leg imaginal disc are described. (1) Wingless signaling is essential for the recruitment of dorsal-proximal, distal, and ventral-proximal leg cells. Wingless requirement in the proximal leg domain appears to be unique to the embryo, since it has previously been shown that Wingless signal transduction is not active in the proximal leg domain in larvae. (2) Downregulation of Wingless signaling in wing disc is essential for its development, suggesting that Wg activity must be downregulated to separate wing and leg discs. In addition, evidence is provided that Dll restricts expression of a proximal leg-specific gene expression. It is proposed that those embryo-specific functions of Wingless signaling reflect its multiple roles in restricting competence of ectodermal cells to adopt the fate of thoracic appendages (Kubota, 2003).

To confirm whether Wg signaling is required cell autonomously for leg disc development, the dominant-negative forms of Drosophila TCF (DTCFdeltaN) or Drosophila axin (Daxin) were expressed in the limb primordia. The Dll-Gal4 driver, which is turned on in the limb primordium at stage 11 and continues to be active in leg and wing discs, was used. In Dll-GAL4 embryos carrying UAS-DTCFdeltaN or UAS-Daxin, the overall size of leg discs was reduced. Expression of Esg was preferentially reduced in the dorsal side. The drastic reduction of Dll mRNA and protein in distal leg cells in the armH8.6 mutants as well as in DTCFdeltaN- and Daxin-expressing embryos demonstrate that Wg signaling is required for both proximal and distal leg cells. In arm mutants, Hth-expressing cells expand to the distal domain. This observation suggests that, upon loss of Wg signaling, prospective leg disc cells lose their identity and adopt the fate of trunk ectoderm. However, disc-specific reduction of Wg signaling does not affect wing disc formation, although the Dll-Gal4 driver is active in the wing primordium. It is concluded that late function of Wg signaling promotes formation of the leg disc with a higher requirement in the proximal domain, but is dispensable for wing disc formation (Kubota, 2003).

C-terminal-binding protein directly activates and represses Wnt transcriptional targets in Drosophila

Regulation of Wnt transcriptional targets is thought to occur by a transcriptional switch. In the absence of Wnt signaling, sequence-specific DNA-binding proteins of the TCF family repress Wnt target genes. Upon Wnt stimulation, stabilized β-catenin binds to TCFs, converting them into transcriptional activators. C-terminal-binding protein (CtBP) is a transcriptional corepressor that has been reported to inhibit Wnt signaling by binding to TCFs or by preventing -catenin from binding to TCF. This study shows that CtBP is also required for the activation of some Wnt targets in Drosophila. CtBP is recruited to Wnt-regulated enhancers in a Wnt-dependent manner, where it augments Armadillo (the fly β-catenin) transcriptional activation. CtBP is required for repression of a subset of Wnt targets in the absence of Wnt stimulation, but in a manner distinct from previously reported mechanisms. CtBP binds to Wnt-regulated enhancers in a TCF-independent manner and represses target genes in parallel with TCF. The data indicate dual roles for CtBP as a gene-specific activator and repressor of Wnt target gene transcription (Fang, 2006).

CtBP has previously been identified as a repressor of Wnt signaling, as measured by TCF reporter genes in cultured cells. Consistent with this, CtBP was identified in an overexpression screen via its ability to suppress Wg and Arm action in the developing eye. In wing imaginal discs, CtBP overexpression also inhibited the Wg target Senseless (Sens). Consistent with this overexpression data, the reduction of CtBP in cultured cells via RNAi is also consistent with a role for CtBP in repressing some Wnt targets (Fang, 2006).

The working model for CtBP repression of Wnt target gene expression holds that CtBP binds to the same area of the nkd and CG6234 loci as TCF, but this binding is TCF-independent. Consistent with this, knock down of CtBP and TCF or gro synergistically derepresses nkd expression. No synergism was seen with TCF/gro double depletions. The RNAi and ChIP data together favor a model where CtBP acts in parallel with TCF/Gro to repress nkd expression in the absence of Wg stimulation. Because CtBP has no detectable ability to bind nucleic acids, it is assumed that unknown DNA-binding protein(s) recruit CtBP to the WRE (Fang, 2006).

The existing models for CtBP antagonism of Wnt signaling cannot explain the data. TCF-independent recruitment of CtBP to WREs is not consistent with work suggesting direct binding of CtBP to TCF. The alternative mechanism, where a CtBP/APC complex diverts Arm/β-catenin away from TCF, also is inconsistent with the results. In this model, the activation of nkd expression after CtBP RNAi treatment would be dependent on TCF and arm. Because the derepression of nkd occurred when both CtBP and TCF were depleted and was not affected when arm was also inhibited, this model is not favored to explain the effects of CtBP depletion on nkd expression. These distinct mechanisms for CtBP repression are not mutually exclusive and may all occur in some contexts (Fang, 2006).

There is a qualitative difference in the amount of derepression found between the two Wg targets studied in Kc cells. Depletion of CtBP and TCF/gro causes a large (20- to 30-fold) increase in nkd basal expression, but has a much more modest (<3-fold) effect on CG6234. These differences may reflect a fundamental difference in the way TCF/Gro and CtBP act on various Wnt targets in unstimulated cells, but it is equally likely that the surrounding cis-elements in these targets have a strong influence on the degree of derepression that can be observed (Fang, 2006).

In addition to defining a novel mechanism for CtBP repression of Wg targets, strong evidence is provided for CtBP playing a role in Wg-mediated transcriptional activation. In the wing imaginal discs, loss of CtBP resulted in a lag in Wg-dependent activation of Sens and a reduction in Dll expression. In cultured Kc cells, CtBP depletion caused a two- to three-fold reduction in the ability of Wg to activate CG6234 expression. The ability of Gal4-Arm chimeras to activate a Gal4 reporter gene is also highly dependent on CtBP levels. In all these contexts, CtBP is not absolutely required for Wg signaling, but is necessary for maximal activation of Wg/Arm transcriptional activation (Fang, 2006).

The positive effect of CtBP on Wg signaling is direct, as judged by ChIP. Assuming that ChIP is measuring the degree of occupancy of CtBP on the chromatin, and not simply antigen accessibility, Wg stimulation promotes the association of CtBP with the CG6234 WRE. This increase in CtBP binding is not observed in TCF-depleted cells. Gal4-Arm recruits endogenous CtBP to a UASluc reporter. Taken together, these data support a model where TCF/Arm recruits CtBP to Wg targets. No binding between Arm and CtBP has been detected by co-immunoprecipitation, suggesting that another factor(s) may act as an adaptor between CtBP and the Arm bound to TCF (Fang, 2006).

Arm has transcriptional activation activity in both the N- and C-terminal portions of the protein. CtBP overexpression or RNAi depletion greatly effects the activity of the N-terminal half of Arm but has no effect on the C-terminal portion. Consistent with this, the N-terminal portion can recruit CtBP to a reporter gene, but not the C-terminus. Other factors that have been linked to the N-terminal portion of Arm include Lgs and Pygo and the ATPases Pontin and Reptin. It may be that CtBP acts in concert with one or more of these factors (Fang, 2006).

CtBPs have strong sequence similarity with D2-hydroxyacid dehydrogenases. hCtBP1 is a functional dehydrogenase and point mutations blocking CtBP1 dehydrogenase activity inhibit its ability to interact with binding partners and act as a transcriptional corepressor. However, another group found that similar mutations had no effect on the ability of CtBP to repress transcription. In this report, mutation of two residues (D290A and H312T) predicted to be essential for catalytic activity had no effect on the ability of fly CtBP to potentiate Gal4-Arm transcriptional activation. Further complicating the issue is data from experiments expressing the fly CtBP fused to Gal4DBD in mammalian cells. In some cells, Gal4-CtBP activated a UAS reporter, while the same reporter was repressed in other cell lines. Interestingly, conversion of CtBP's catalytic histidine to glutamine abolished transcriptional activation, but not repression. The heterologous nature of these experiments and the differences in the assays employed may explain the discrepancy between these studies, and further experiments will be needed on endogenous targets to determine how much dehydrogenase activity of CtBP contributes to repression and activation of Wnt targets (Fang, 2006).

Although CtBP is required for maximal activation of CG6234 expression and a Gal4-Arm-dependent reporter gene, Wg activation of nkd does not appear to require CtBP. The basis for this gene-specific requirement for CtBP is not clear. CtBP is recruited to the nkd WRE in a Wg-dependent manner, similar to what was observed for CG6234. It may be that CtBP is required for nkd activation, but this is masked by its role in repressing nkd expression. This hypothesis could be tested if were possible to separate CtBP's activator and repressor activities (Fang, 2006).

The requirement for CtBP in Wnt transcriptional activation may have been previously overlooked due to its well-characterized role as a co-repressor. For example, mouse embryos that lack CtBP2 have axial truncations and reduced Brachyury (T) expression that is reminiscent of Wnt3a mutants. These results suggest that the activating role for CtBP in Wnt signaling that was identified is evolutionarily conserved (Fang, 2006).

SoxNeuro acts with Tcf to control Wg/Wnt signaling activity

Wnt signaling specifies cell fates in many tissues during vertebrate and invertebrate embryogenesis. To understand better how Wnt signaling is regulated during development, genetic screens were performed to isolate mutations that suppress or enhance mutations in the fly Wnt homolog, wingless (wg). This study finds that loss-of-function mutations in the neural determinant SoxNeuro (also known as Sox-neuro, SoxN) partially suppress wg mutant pattern defects. SoxN encodes a HMG-box-containing protein related to the vertebrate Sox1, Sox2 and Sox3 proteins, which have been implicated in patterning events in the early mouse embryo. In Drosophila, SoxN has been shown to specify neural progenitors in the embryonic central nervous system. This study shows that SoxN negatively regulates Wg pathway activity in the embryonic epidermis. Loss of SoxN function hyperactivates the Wg pathway, whereas its overexpression represses pathway activity. Epistasis analysis with other components of the Wg pathway places SoxN at the level of the transcription factor Pan (also known as Lef, Tcf) in regulating target gene expression. In human cell culture assays, SoxN represses Tcf-responsive reporter expression, indicating that the fly gene product can interact with mammalian Wnt pathway components. In both flies and in human cells, SoxN repression is potentiated by adding ectopic Tcf, suggesting that SoxN interacts with the repressor form of Tcf to influence Wg/Wnt target gene transcription (Chao, 2007).

SoxN downregulates the Wg/Wnt pathway to reduce target gene expression. Downregulation is a crucial process because it sensitizes the signal response to allow rapid on/off switching and also keeps the system off in cells that are not actively responding to signal. Many genes have been shown to negatively regulate Wg/Wnt pathway activity through the destabilization of Arm/beta-catenin. Far fewer are known to exert negative regulatory effects downstream of Arm. The vertebrate Sox proteins -- Sox9, XSox3, XSox17α and XSox17ß -- as well as Chibby, a conserved nuclear factor, antagonize Wg/Wnt signaling by binding to Arm/beta-catenin and preventing it from partnering with Tcf to activate target gene expression. SoxN, however, does not bind beta-catenin in cell-culture assays, and does not share strong homology with the C-terminal sequences through which vertebrate Sox proteins bind this protein. Furthermore, SoxN function is not influenced by Arm levels. No difference was observed in SoxN-mediated TOPflash repression when cells were induced by co-transfection with a constitutively stabilized beta-catenin versus with Wnt-induced medium. Instead, both TOPflash and genetic experiments indicate that SoxN function depends on Tcf and Gro, its co-repressor (Chao, 2007).

One way to explain these observations is that SoxN contributes to the assembly or stability of the Tcf repressor complex on DNA. The consensus-sequence recognition for HMG domains in the Sox and Tcf families is reported to be similar, although XSox3 and XSox17ß fail to bind a consensus Tcf DNA sequence. It is shown that SoxN does not compete for Tcf-binding sites as a means of repressing target gene transcription, but the data support a model in which SoxN might bind DNA elsewhere or might bind Tcf sites transiently to initiate or stabilize the assembly of a repressor complex (Chao, 2007).

A similar model may explain the results from Xenopus that showed that XSox3-mediated repression does not require interaction between XSox3 and beta-catenin. XSox3 strongly interferes with dorsal fate specification in Xenopus embryos and represses TOPflash-reporter activity in vitro. HMG-domain mutations render XSox3 inactive in embryos without affecting its interaction with beta-catenin or its repression in TOPflash assays. Thus, it is the DNA-binding domain, not the beta-catenin-interacting C-terminus, that is relevant to its in vivo function in dorsal determination in Xenopus. XSox3 represses the expression of the dorsal-specific Nodal-related gene Xnr5 through optimal core binding sequences adjacent to and partially overlapping with Tcf sites in the Xnr5 promoter (Zhang, 2003). By contrast, the fly SoxN shows no discrepancy between its behavior in TOPflash assays and its in vivo effects. This suggests that the synthetic Tcf-binding sites arranged in the TOPflash-reporter plasmid are sufficient to support SoxN repressor function (Chao, 2007).

Because adding Tcf-site competitor DNA does not diminish the repressive capacity of limiting amounts of SoxN, the role of SoxN in repression does not appear to be stoichiometric. Therefore, the idea is favored that Sox proteins may act in a catalytic fashion during repressor-complex assembly at Wnt target gene promoters, rather than forming a structural part of the repressor complex itself. It was not possible to detect direct binding of SoxN with either Tcf, Gro or Arm, raising the possibility that SoxN interacts with some as yet unidentified protein that chaperones assembly of the repressor complex. A SoxN-binding cofactor, SNCF, has been identified in Drosophila (Bonneaud, 2003), but this gene is expressed only in pre-gastrulation embryos. Because Wg signaling occurs exclusively post-gastrulation, and specification of naked cuticle begins more than 4 hours after gastrulation, it is not thought that SNCF is a likely candidate for mediating this aspect of SoxN function. Rather, it is likely to play a role in the neuronal specification events promoted by SoxN at earlier stages of embryogenesis (Chao, 2007).

It is curious that uniformly overexpressed SoxN represses Wg signal transduction in dorsal epidermal cells more severely than in ventral cells. This effect is evident in both cuticle pattern elements and in en expression, and is reminiscent of defects observed in the 'transport-defective' class of wg mutant alleles, which includes wgNE2. These mutations restrict Wg-ligand movement ventrally to promote only local signaling response while simultaneously abolishing all dorsal signaling, suggesting a fundamental difference in ventral and dorsal cell response. Perhaps it is not a coincidence that the NC14 mutation was isolated in the wgNE2 mutant background. Further analysis of SoxN function may help to determine the molecular basis for dorsoventral differences in Wg signal transduction (Chao, 2007).

The role of Pygopus in the differentiation of intracardiac valves in Drosophila

Cardiac valves serve an important function; they support unidirectional blood flow and prevent blood regurgitation. Wnt signaling plays an important role in the formation of mouse cardiac valves and cardiac valve proliferation in Zebrafish, but identification of the specific signaling components involved has not been addressed systematically. Of the components involved in Wnt signal transduction, pygopus (pygo), first identified as a core component of Wnt signaling in Drosophila, has not been investigated with respect to valve development and differentiation. This study took advantage of the Drosophila heart model to study the role of pygo in formation of valves between the cardiac chambers. Cardiac-specific pygo knockdown in the Drosophila heart was found to cause dilation in the region of these cardiac valves, and their characteristic dense mesh of myofibrils does not form and resembles that of neighboring cardiomyocytes. In contrast, heart-specific knockdown of the transcription factors, arm/beta-Cat, lgs/BCL9, or pan/TCF, which mediate canonical Wnt signal transduction, shows a much weaker valve differentiation defect. Double-heterozygous combinations of mutants for pygo and the Wnt-signaling components have no additional effect on heart function compared with pygo heterozygotes alone. These results are consistent with the idea that pygo functions independently of canonical Wnt signaling in the differentiation of the adult interchamber cardiac valves (Tang, 2014).


REFERENCES

Adam, R. C., Yang, H., Ge, Y., Lien, W. H., Wang, P., Zhao, Y., Polak, L., Levorse, J., Baksh, S. C., Zheng, D. and Fuchs, E. (2018). Temporal layering of signaling effectors drives chromatin remodeling during hair follicle stem cell lineage progression. Cell Stem Cell 22(3): 398-413. PubMed ID: 29337183

Ahmed, Y., et al. (1998). Regulation of armadillo by a Drosophila APC inhibits neuronal apoptosis during retinal development. Cell 93(7): 1171-1182. PubMed Citation: 9657150

Alvarez-Medina, R., Le Dreau, G., Ros, M. and Martí, E. (2009). Hedgehog activation is required upstream of Wnt signalling to control neural progenitor proliferation. Development 136(19): 3301-9. PubMed Citation: 19736325

Andoniadou, C. L., et al. (2011). HESX1- and TCF3-mediated repression of Wnt/?-catenin targets is required for normal development of the anterior forebrain. Development 138(22): 4931-42. PubMed Citation: 22007134

Angerer, L. M., Newman, L. A. and Angerer, R. C. (2005). SoxB1 downregulation in vegetal lineages of sea urchin embryos is achieved by both transcriptional repression and selective protein turnover. Development 132: 999-1008. 15689377

Aoki, M., et al. (1999). Nuclear endpoint of wnt signaling: neoplastic transformation induced by transactivating lymphoid-enhancing factor 1. Proc. Natl. Acad. Sci. 96(1): 139-44. PubMed Citation: 9874785

Archbold, H. C., Broussard, C., Chang, M. V. and Cadigan, K. M. (2014). Bipartite recognition of DNA by TCF/Pangolin is remarkably flexible and contributes to transcriptional responsiveness and tissue specificity of Wingless signaling. PLoS Genet 10: e1004591. PubMed ID: 25188465

Arnold, S. J., et al. (2000). Brachyury is a target gene of the Wnt/beta-catenin signaling pathway. Mech. Dev. 91: 249-258. PubMed Citation: 10704849

Bagga, R. and Emerson, B. M. (1997). An HMG I/Y-containing repressor complex and supercoiled DNA topology are critical for long-range enhancer-dependent transcription in vitro. Genes Dev. 11: 629-639. PubMed Citation: 9119227

Balakirev, M. Y., Tcherniuk, S. O., Jaquinod, M. and Chroboczek, J. (2003). Otubains: A new family of cysteine proteases in the ubiquitin pathway. EMBO Rep. 4: 517-522. PubMed Citation: 12704427

Barrio, L. and Milan, M. (2020). Regulation of anisotropic tissue growth by two orthogonal signaling centers. Dev Cell 52(5): 659-672. PubMed ID: 32084357

Benchabane, H., et al. (2011). Jerky/Earthbound facilitates cell-specific Wnt/Wingless signalling by modulating beta-catenin-TCF activity. EMBO J. 30: 1444-1458. PubMed Citation: 21399610

Barker, N., et al. (2001). The chromatin remodeling factor Brg-1 interacts with ß-catenin to promote target gene activation. EMBO J. 20: 4935-4943. 11532957

Batlle, E., et al. (2002). ß-Catenin and TCF mediate cell positioning in the intestinal epithelium by controlling the expression of EphB/EphrinB. Cell 111: 251-263. 12408869

Behrens, J., et al. (1996). Functional interaction of beta-catenin with the transcription factor LEF-1. Nature 382: 638-642. PubMed Citation: 8757136

Benchabane, H., et al. (2011). Jerky/Earthbound facilitates cell-specific Wnt/Wingless signalling by modulating beta-catenin-TCF activity. EMBO J. 30: 1444-1458. PubMed Citation: 21399610

Bertrand, V. and Hobert, O. (2009). Linking asymmetric cell division to the terminal differentiation program of postmitotic neurons in C. elegans. Dev. Cell 16(4): 563-75. PubMed Citation: 19386265

Bonneaud, N., Savare, J., Berta, P. and Girard, F. (2003). SNCF, a SoxNeuro interacting protein, defines a novel protein family in Drosophila melanogaster. Gene 319: 33-41. Medline abstract: 14597169

Brannon, M., et al. (1997). A beta-catenin/XTcf-3 complex binds to the siamois promoter to regulate dorsal axis specification in Xenopus. Genes Dev. 11(18): 2359-2370. PubMed Citation: 9308964

Brannon, M., et al. (1999). XCtBP is a XTcf-3 co-repressor with roles throughout Xenopus development. Development 126: 3159-3170. PubMed Citation: 10375506

Brantjes, H., Roose, J., van De Wetering, M. and Clevers, H. (2001). All Tcf HMG box transcription factors interact with Groucho-related co-repressors. Nucleic Acids Res. 29(7): 1410-9. 11266540

Brugmann, S. A., et al. (2007). Wnt signaling mediates regional specification in the vertebrate face. Development 134(18): 3283-95. PubMed citation: 17699607

Bruhn, L., Munnerlyn, A. and Grosschedl, R. (1997). ALY, a context-dependent coactivator of LEF-1 and AML-1, is required for TCRalpha enhancer function. Genes Dev. 11: 640-653. PubMed Citation: 9119228

Brunner, E., et al. (1997). pangolin encodes a Lef-1 homologue that acts downstream of Armadillo to transduce the Wingless signal in Drosophila. Nature 385: 829-933. PubMed Citation: 9039917

Calvo, D. et al. (2001). A POP-1 repressor complex restricts inappropriate cell type-specific gene transcription during Caenorhabditis elegans embryogenesis. EMBO J. 20: 7197-7208. 11742996

Cavallo, R. A., et al. (1998). Drosophila Tcf and Groucho interact to repress Wingless signalling activity. Nature 395(6702): 604-8. PubMed Citation: 9783586

Chang, W., Lloyd, C. E. and Zarkower, D. (2005). DSH-2 regulates asymmetric cell division in the early C. elegans somatic gonad Mech. Dev. 122: 781-789. 15907376

Chao, A. T., Jones, W. M. and Bejsovec, A. (2007). The HMG-box transcription factor SoxNeuro acts with Tcf to control Wg/Wnt signaling activity. Development 134(5): 989-97. Medline abstract: 17267442

Cho, S. H. and Cepko, C. L. (2006). Wnt2b/β-catenin-mediated canonical Wnt signaling determines the peripheral fates of the chick eye. Development 133(16): 3167-77. 16854977

Chodaparambil, J. V., Pate, K. T., Hepler, M. R., Tsai, B. P., Muthurajan, U. M., Luger, K., Waterman, M. L. and Weis, W. I. (2014). Molecular functions of the TLE tetramerization domain in Wnt target gene repression. EMBO J 33(7): 719-731. PubMed ID: 24596249

Collins, R. T. and Treisman, J. E. (2000). Osa-containing Brahma chromatin remodeling complexes are required for the repression of Wingless target genes. Genes Dev. 14: 3140-3152. 11124806

Conacci-Sorrell, M. E., et al. (2002). Nr-CAM is a target gene of the ß-catenin/LEF-1 pathway in melanoma and colon cancer and its expression enhances motility and confers tumorigenesis. Genes Dev. 16: 2058-2072. 12183361

Cox, R. T., et al. (1999). Membrane-tethered Drosophila Armadillo cannot transduce Wingless signal on its own. Development 126: 1327-1335. PubMed Citation: 10021350

Daniels, D. L. and Weis, W. I. (2002). ICAT inhibits ß-Catenin binding to Tcf/Lef-family transcription factors and the general coactivator p300 using independent structural modules. Molec. Cell 10: 573-584. 12408825

DasGupta, R. and Fuchs, E. (1999). Multiple roles for activated LEF/TCF transcription complexes during hair follicle development and differentiation. Development 126: 4557-4568. PubMed Citation: 10498690

Deb, D. K., Tanaka-Matakatsu, M., Jones, L., Richardson, H. E. and Du, W. (2008). Wingless signaling directly regulates cyclin E expression in proliferating embryonic PNS precursor cells. Mech. Dev. 125(9-10): 857-64. PubMed Citation: 18606221

de la Roche, M. and Bienz, M. (2007). Wingless-independent association of Pygopus with dTCF target genes. Curr. Biol. 17(6): 556-61. Medline abstract: 17320388

Deshpande, R., Inoue, T., Priess, J. R. and Hill, R. J. (2005). lin-17/Frizzled and lin-18 regulate POP-1/TCF-1 localization and cell type specification during C. elegans vulval development. Dev. Biol. 278(1): 118-29. 15649465

Dooijes, D., et al. (1998). Genomic organization of the segment polarity gene pan in Drosophila melanogaster. Mol. Gen. Genet. 258(1-2): 45-52. PubMed Citation: 9613571

Dorsky, R. I., Moon, R. T. and Raible, D. W. (1998). Control of neural crest cell fate by the Wnt signalling pathway. Nature 396(6709): 370-3. PubMed Citation: 9845073

Dorsky, R. I., et al. (1999). Maternal and embryonic expression of zebrafish lef1. Mech. Dev. 86(1-2): 147-150. PubMed Citation: 10446273

Dorsky, R. I., Itoh, M., Moon, R. T. and Chitnis, A. J. (2003). Two tcf3 genes cooperate to pattern the zebrafish brain. Development 130: 1937-1947. 12642497

Dorsky, R. I., Sheldahl, L. C. and Moon, R. T. (2002). A transgenic Lef1/beta-catenin-dependent reporter is expressed in spatially restricted domains throughout zebrafish Development. Dev. Biol. 241(2): 229-37. 11784107

Duan, D., et al. (1999). Submucosal gland development in the airway is controlled by Lymphoid Enhancer Binding Factor 1 (LEF1). Development 126: 4441-4453. PubMed Citation: 10498680

Eresh, S., et al. (1997). A CREB-binding site as a target for decapentaplegic signalling during Drosophila endoderm induction. EMBO J. 16: 2014-22. PubMed Citation: 9155027

Estella, C., McKay, D. J. and Mann, R. S. (2008). Molecular integration of Wingless, Decapentaplegic, and autoregulatory inputs into Distalless during Drosophila leg development. Dev. Cell 14: 86-96. PubMed Citation: 18194655

Fagotto, F., Gluck, U. and Gumbiner, B. M. (1998). Nuclear localization signal-independent and importin/karyopherin-independent nuclear import of beta-catenin. Curr. Biol. 8(4): 181-190. PubMed Citation: 9501980

Fancy, S. P., et al. (2009). Dysregulation of the Wnt pathway inhibits timely myelination and remyelination in the mammalian CNS. Genes Dev. 23(13): 1571-85. PubMed Citation: 19515974

Fang, M., Li, J., Blauwkamp, T., Bhambhani, C., Campbell, N. and Cadigan, K. M. (2006). C-terminal-binding protein directly activates and represses Wnt transcriptional targets in Drosophila. EMBO J. 25(12): 2735-45. 16710294

Feng, Y., et al. (2007). Drosophila split ends homologue SHARP functions as a positive regulator of Wnt/β-Catenin/T-cell factor signaling in neoplastic transformation, Cancer Res. 67: 485-491. PubMed Citation: 17234755

Fiedler, M., Sanchez-Barrena, M. J., Nekrasov, M., Mieszczanek, J., Rybin, V., Muller, J., Evans, P. and Bienz, M. (2008). Decoding of methylated histone H3 tail by the Pygo-BCL9 Wnt signaling complex. Mol Cell 30(4): 507-518. PubMed ID: 18498752

Fiedler, M., Graeb, M., Mieszczanek, J., Rutherford, T. J., Johnson, C. M. and Bienz, M. (2015). An ancient Pygo-dependent Wnt enhanceosome integrated by Chip/LDB-SSDP. Elife 4. PubMed ID: 26312500

Galceran, J., et al. (1999). Wnt3a-/--like phenotype and limb deficiency in Lef1(-/-)Tcf1(-/-) mice. Genes Dev. 13(6): 709-17. PubMed Citation: 10090727

Galceran, J., et al. (2000). Hippocampus development and generation of dentate gyrus granule cells is regulated by LEF1. Development 127: 469-482. PubMed Citation: 10631168

Galceran, J., Hsu, S.-C. and Grosschedl, R. (2001). Rescue of a Wnt mutation by an activated form of LEF-1: Regulation of maintenance but not initiation of Brachyury expression. Proc. Natl. Acad. Sci. 98: 8668-8673. 11447280

Galceran, J., Sustmann, C., Hsu, S. C., Folberth, S. and Grosschedl, R. (2004). LEF1-mediated regulation of Delta-like1 links Wnt and Notch signaling in somitogenesis. Genes Dev. 18(22): 2718-23. 15545629

Gao, Y., Toska, E., Denmon, D., Roberts, S. G. and Medler, K. F. (2014). WT1 regulates the development of the posterior taste field. Development 141: 2271-2278. PubMed ID: 24803588

Gat, U., et al. (1998). De Novo hair follicle morphogenesis and hair tumors in mice expressing a truncated beta-catenin in skin. Cell 95(5): 605-14

Giese, K., et al. (1995). Assembly and function of a TCR alpha enhancer complex is dependent on LEF-1-induced DNA bending and multiple protein-protein interactions. Genes Dev. 9(8): 995-1008

Giese, K., Pagel, J. and Grosschedl, R. (1997). Functional analysis of DNA bending and unwinding by the high mobility group domain of LEF-1. Proc. Natl. Acad. Sci. 94(24): 12845-12850

Graham, T. A., et al. (2002). The crystal structure of the ß-Catenin/ICAT complex reveals the inhibitory mechanism of ICAT Molec. Cell 10: 563-571. 12408824

Gribble, S. L., Kim, H. S., Bonner, J., Wang, X. and Dorsky, R. I. (2009). Tcf3 inhibits spinal cord neurogenesis by regulating sox4a expression. Development 136(5): 781-9. PubMed Citation: 19176587

Halfon, M. S., et al. (2000). Ras pathway specificity is determined by the integration of multiple signal-activated and tissue-restricted transcription factors. Cell 103: 63-74.

Hamilton, F. S., Wheeler, G. N. and Hoppler, S. (2001). Difference in XTcf-3 dependency accounts for change in response to ß-catenin-mediated Wnt signalling in Xenopus blastula. Development 128: 2063-2073

Hanson, A. J., et al. (2012). XIAP monoubiquitylates Groucho/TLE to promote canonical Wnt signaling. Mol. Cell 45(5): 619-28. PubMed Citation: 22304967

Haremaki, T., et al. (2003). Integration of multiple signal transducing pathways on Fgf response elements of the Xenopus caudal homologue Xcad3. Development 130: 4907-4917. 12930781

Haussmann I., White, K. and Soller, M. (2008). Erect wing regulates synaptic growth in Drosophila by integration of multiple signaling pathways. Genome Biol. 9: R73. PubMed Citation: 18419806

He, T.-C., et al. (1998). Identification of c-MYC as a target of the APC pathway. Science 281(5382): 1509-1512

Herman, M. A. (2001). C. elegans POP-1/TCF functions in a canonical Wnt pathway that controls cell migration and in a noncanonical Wnt pathway that controls cell polarity. Development 128: 581-590. 11171341

Hepker, J., Blackman, R. K. and Holmgren, R. (1999). Cubitus interruptus is necessary but not sufficient for direct activation of a wing-specific decapentaplegic enhancer. Development 126: 3669-3677

Hikasa, H. and Sokol, S. Y. (2004). The involvement of Frodo in TCF-dependent signaling and neural tissue development. Development 131: 4725-4734. 15329348

Hilton, E., Rex, M. and Old, R. (2003). VegT activation of the early zygotic gene Xnr5 requires lifting of Tcf-mediated repression in the Xenopus blastula. Mech. Dev. 120: 1127-1138. 14568102

Hobmayer, B., et al. (2000). WNT signaling molecules act in axis formation in the diploblastic metazoan Hydra. Nature 407: 186-189.

Hofmann, M., et al. (2004). WNT signaling, in synergy with T/TBX6, controls Notch signaling by regulating Dll1 expression in the presomitic mesoderm of mouse embryos. Genes Dev. 18(22): 2712-7. 15545628

Houston, D. W., et al. (2002). Repression of organizer genes in dorsal and ventral Xenopus cells mediated by maternal XTcf3. Development 129: 4015-4025. 12163405

Hsu, S. C., Galceran, J., Grosschedl, R. (1998). Modulation of transcriptional regulation by LEF-1 in response to wnt-1 signaling and association with beta-catenin. Mol. Cell. Biol. 18(8): 4807-4818

Hu, M. C. and Rosenblum, N. D. (2005). Smad1, ß-catenin and Tcf4 associate in a molecular complex with the Myc promoter in dysplastic renal tissue and cooperate to control Myc transcription. Development 132: 215-225. 15576399

Huang, S., Shetty, P., Robertson, S. M. and Lin, R. (2007). Binary cell fate specification during C. elegans embryogenesis driven by reiterated reciprocal asymmetry of TCF POP-1 and its coactivator β-catenin SYS-1. Development 134(14): 2685-95. PubMed citation; Online text

Huber, O., et al. (1996). Nuclear localization of ß-catenin by interaction with transcription factor LEF-1. Mech. Dev. 56: 3-10

Jiang, L. I. and Sternberg, P. W. (1999). An HMG1-like protein facilitates Wnt signaling in Caenorhabditis elegans. Genes Dev. 13(7): 877-889

Johnson, A. N., Bergman, C. M., Kreitman, M. and Newfeld, S. J. (2003). Embryonic enhancers in the dpp disk region regulate a second round of Dpp signaling from the dorsal ectoderm to the mesoderm that represses Zfh-1 expression in a subset of pericardial cells. Dev. Biol. 262: 137-151. 14512024

Kadoch, C. and Crabtree, G. R. (2015). Mammalian SWI/SNF chromatin remodeling complexes and cancer: Mechanistic insights gained from human genomics. Sci Adv 1(5): e1500447. PubMed ID: 26601204

Kardon, G., Harfe, B. D. and Tabin, C. J. (2003). A Tcf4-positive mesodermal population provides a prepattern for vertebrate limb muscle patterning. Dev. Cell 5: 937-944. 14667415

Kengaku, M., et al. (1998). Distinct WNT pathways regulating AER formation and dorsoventral polarity in the chick limb bud. Science 280(5367): 1274-1277

Kiger, J. A., Natzle, J. E. and Green, M. M. (2001). Hemocytes are essential for wing maturation in Drosophila melanogaster. Proc. Natl. Acad. Sci. 98(18): 10190-5. 11504926

Kioussi, C., et al. (2002). Identification of a Wnt/Dvl/ß-Catenin --> Pitx2 pathway mediating cell-type-specific proliferation during development. Cell 111: 673-685. 12464179

Klingel, S., Morath, I., Strietz, J., Menzel, K., Holstein, T. W. and Gradl, D. (2012). Subfunctionalization and neofunctionalization of vertebrate Lef/Tcf transcription factors. Dev Biol 368: 44-53. Pubmed: 22641013

Knirr, S. and Frasch, M. (2001). Molecular integration of inductive and mesoderm-intrinsic inputs governs even-skipped enhancer activity in a subset of pericardial and dorsal muscle progenitors. Dev. Bio. 238: 13-26

Korinek, V., et al. (1997). Constitutive transcriptional activation by a ß-Catenin-Tcf complex in APC-/- colon carcinoma. Science 275: 1784-87

Korinek, V., et al. (1998). Two members of the Tcf family implicated in Wnt/beta-catenin signaling during embryogenesis in the mouse. Mol. Cell. Biol. 18(3): 1248-1256

Korswagen, H. C., (2000). Distinct beta-catenins mediate adhesion and signaling functions in C. elegans. Nature 406: 527-532.

Krieghoff, E., Behrens, J. and Mayr, B. (2006). Nucleo-cytoplasmic distribution of β-catenin is regulated by retention. J. Cell Sci. 119(Pt 7): 1453-63. 16554443

Kubota, K., Goto, S. and Hayashi, S. (2003). The role of Wg signaling in the patterning of embryonic leg primordium in Drosophila. Dev. Biol. 257: 117-126. 12710961

Kuhl, M., et al. (2001). Antagonistic regulation of convergent extension movements in Xenopus by Wnt/ß-catenin and Wnt/Ca2+ signaling. Mech. Dev. 106: 61-76. 11472835

Kunz, M., et al. (2004). Autoregulation of canonical Wnt signaling controls midbrain development. Dev. Biol. 273: 390-401. 15328021

Kuwahara, A., et al. (2010). Wnt signaling and its downstream target N-myc regulate basal progenitors in the developing neocortex. Development. 2010 Apr;137(7):1035-44. PubMed Citation: 20215343

Labbe, E., Letamendia, A. and Attisano, L. (2000). Association of smads with lymphoid enhancer binding factor 1/T cell-specific factor mediates cooperative signaling by the transforming growth factor-beta and wnt pathways. Proc. Natl. Acad. Sci. 97(15): 8358-63.

Lam, N., Chesney, M. A. and Kimble, J. (2006). Wnt signaling and CEH-22/tinman/Nkx2.5 specify a stem cell niche in C. elegans. Curr. Biol. 16: 287-295. 16461282

Laurent, M. N., et al. (1997). The Xenopus homeobox gene twin mediates Wnt induction of goosecoid in establishment of Spemann's organizer. Development 124(23): 4905-4916

Lawrence, N., et al. (2001). Notch signaling targets the Wingless responsiveness of a Ubx visceral mesoderm enhancer in Drosophila. Curr. Biol. 11: 375-385. 11301248

Lee, H.-H. and Frasch, M. (2000). Wingless effects mesoderm patterning and ectoderm segmentation events via induction of its downstream target sloppy paired. Development 127: 5497-5508

Lee, W., Swarup, S., Chen, J., Ishitani, T. and Verheyen, E. M. (2009) Homeodomain-interacting protein kinases (Hipks) promote Wnt/Wg signaling through stabilization of β-catenin/Arm and stimulation of target gene expression. Development 136(2): 241-51. PubMed Citation: 19088090

Levanon, D., et al. (1998). Transcriptional repression by AML1 and LEF-1 is mediated by the TLE/Groucho corepressors. Proc. Natl. Acad. Sci. 95(20): 11590-5. PubMed Citation: 9751710

Li, B., et al. (2002). The LEF1/beta-catenin complex activates movo1, a mouse homolog of Drosophila ovo required for epidermal appendage differentiation. Proc. Natl Acad. Sci. 99: 6064-6069. PubMed Citation: 11983900

Li, J., Wang, C. Y. (2008). TBL1-TBLR1 and beta-catenin recruit each other to Wnt target-gene promoter for transcription activation and oncogenesis. Nat Cell Biol 10: 160-169. PubMed Citation: 18193033

Lickert, H., et al. (2000). Wnt/beta-catenin signaling regulates the expression of the homeobox gene Cdx1 in embryonic intestine. Development 127: 3805-3813. PubMed Citation: 10934025

Lin, R., Hill, R. J. and Priess, J. R. (1998). POP-1 and anterior-posterior fate decisions in C. elegans embryos. Cell 92: 229-239. PubMed Citation: 9458047

Lo, M.-C., et al. (2004). Phosphorylation by the ß-Catenin/MAPK complex promotes 14-3-3-mediated nuclear export of TCF/POP-1 in signal-responsive cells in C. elegans. Cell 117: 95-106. 15066285

Lowry, W. E., et al. (2005). Defining the impact of ß-catenin/Tcf transactivation on epithelial stem cells. Genes Dev. 19: 1596-1611. 15961525

Maduro, M. F., Lin, R. and Rothman, J. H. (2002). Dynamics of a developmental switch: Recursive intracellular and intranuclear redistribution of Caenorhabditis elegans POP-1 parallels Wnt-inhibited transcriptional repression. Dev. Bio. 248: 128-142. 12142026

Major, R. J. and Irvine, K. D. (2005). Influence of Notch on dorsoventral compartmentalization and actin organization in the Drosophila wing. Development 132(17): 3823-33. 16049109

Mayall, T. P., et al. (1997). Distinct roles for P-CREB and LEF-1 in TCRalpha enhancer assembly and activation on chromatin templates in vitro Genes Dev. 11: 887-899

McCartney, B. M., et al. (1999). Drosophila APC2 is a cytoskeletally-associated protein that regulates wingless signaling in the embryonic epidermis. J. Cell Biol. 146(6): 1303-18

McGraw, H. F., (2011). Lef1 is required for progenitor cell identity in the zebrafish lateral line primordium. Development 138(18): 3921-30. PubMed Citation: 21862556

McKendry, R., et al. (1997). LEF-1/TCF proteins mediate wnt-inducible transcription from the Xenopus nodal-related 3 promoter. Dev. Biol. 192(2): 420-431

Merrill, B. J., et al. (2001). Tcf3 and Lef1 regulate lineage differentiation of multipotent stem cells in skin. Genes Dev. 15: 1688-1705. 11445543

Merrill, B. J., et al. (2003). Tcf3: a transcriptional regulator of axis induction in the early embryo. Development 131: 263-274. 14668413

Nguyen, D. X., et al. (2009). WNT/TCF signaling through LEF1 and HOXB9 mediates lung adenocarcinoma metastasis. Cell 138(1): 51-62. PubMed Citation: 19576624

Mizumoto, K. and Sawa, H. (2007). Cortical beta-catenin and APC regulate asymmetric nuclear beta-catenin localization during asymmetric cell division in C. elegans. Dev. Cell 12(2): 287-99. PubMed Citation: 17276345

Molenaar, M., et al. (1996). XTcf-3 transcription factor mediates ß-catenin-induced axis formation in Xenopus embryos. Cell 86: 391-399. PubMed Citation: 9739131

Moon, R. T., Brown, J. D. and Torres, M. (1997). WNTs modulate cell fate and behavior during vertebrate development. Trends Genet. 13: 157-162. PubMed Citation: 9097727

Novak, A., et al. (1998). Cell adhesion and the integrin-linked kinase regulate the LEF-1 and beta-catenin signaling pathways. Proc. Natl. Acad. Sci. 95(8): 4374-4379. PubMed Citation: 9539744

Nyholm, M. K., Wu, S. F., Dorsky, R. I. and Grinblat, Y. (2007). The zebrafish zic2a-zic5 gene pair acts downstream of canonical Wnt signaling to control cell proliferation in the developing tectum. Development 134(4): 735-46. Medline abstract: 17215296

Olson, L. E., et al. (2006). Homeodomain-mediated beta-catenin-dependent switching events dictate cell-lineage determination. Cell 125(3): 593-605. Medline abstract: 16678101

Onai, T., et al. (2004). Xenopus XsalF anterior neuroectodermal specification by attenuating cellular responsiveness to Wnt signaling. Dev. Cell 7: 95-106. 15239957

Orsulic, S., et al. (1999). E-cadherin binding prevents beta-catenin nuclear localization and beta-catenin/LEF-1-mediated transactivation. J. Cell Sci. 112(Pt 8): 1237-1245

Park, F. D. and Priess, J. R. (2003). Establishment of POP-1 asymmetry in early C. elegans embryos. Development 130: 3547-3556. 12810601

Park, J. I., et al. (2005). Kaiso/p120-catenin and TCF/beta-catenin complexes coordinately regulate canonical Wnt gene targets. Dev. Cell 8(6): 843-54. 15935774

Persad, S., et al. (2001). Tumor suppressor PTEN inhibits nuclear accumulation of ß-Catenin and T cell/Lymphoid enhancer factor 1-mediated transcriptional activation. J. Cell Bio. 153: 1161-1174. 11402061

Piepenburg, O., Vorbruggen, G., Jackle, H. (2000). Drosophila segment borders result from unilateral repression of hedgehog activity by wingless signaling. Mol. Cell 6(1): 203-9. 10949042

Ren, H. and Zhang, H. (2010). Wnt signaling controls temporal identities of seam cells in Caenorhabditis elegans. Dev. Biol. 345: 144-155. PubMed Citation: 20624379

Riese, J., et al. (1997). LEF-1, a nuclear factor coordinating signaling inputs from wingless and decapentaplegic. Cell 88: 777-787. PubMed Citation: 9118221

Pilon, N., Oh, K., Sylvestre, J. R., Savory, J. G. and Lohnes, D. (2007). Wnt signaling is a key mediator of Cdx1 expression in vivo. Development 134(12): 2315-23. Medline abstract: 17537796

Rocheleau, C. E., et al. (1997). Wnt signaling and an APC-related gene specify endoderm in early C. elegans embryos. Cell 90(4): 707-716. PubMed Citation: 9288750

Rocheleau, C. E., et al. (1999). WRM-1 activates the LIT-1 protein kinase to transduce anterior/posterior polarity signals in C. elegans. Cell 97(6): 717-26. PubMed Citation: 10380924

Roe, G., et al. (2002). Lef-1 and Tcf-3 transcription factors mediate tissue-specific Wnt signaling during Xenopus development. Curr. Biol. 12: 1941-1945. 12445388

Rong, X., Zhou, Y., Liu, Y., Zhao, B., Wang, B., Wang, C., Gong, X., Tang, P., Lu, L., Li, Y., Zhao, C. and Zhou, J. (2017). Glutathione peroxidase 4 inhibits Wnt/beta-catenin signaling and regulates dorsal organizer formation in zebrafish embryos. Development 144(9):1687-1697. PubMed ID: 28302747

Roose, J., et al. (1998). The Xenopus Wnt effector XTcf-3 interacts with Groucho-related transcriptional repressors. Nature 395(6702): 608-12. PubMed Citation: 9783587

Ruzov, A., et al. (2009). The interaction of xKaiso with xTcf3: a revised model for integration of epigenetic and Wnt signalling pathways. Development 136(5): 723-7. PubMed Citation: 19158184

Sadot, E., et al. (1998). Inhibition of beta-catenin-mediated transactivation by cadherin derivatives. Proc. Natl. Acad. Sci. 95(26): 15339-44. PubMed Citation: 9860970

Saller, E., Kelley, A. and Bienz, M. (2002). The transcriptional repressor Brinker antagonizes Wingless signaling. Genes Dev. 16: 1828-1838. 12130542

Sampson, E. M., et al. (2001). Negative regulation of the Wnt-ß-catenin pathway by the transcriptional repressor HBP1. EMBO J. 20: 4500-4511. 11500377

Sasaki, T., et al. (2005). LEF1 is a critical epithelial survival factor during tooth morphogenesis. Dev. Biol. 278(1): 130-43. 15649466

Schmidt-Ott, K. M., et al. (2007). beta-catenin/TCF/Lef controls a differentiation-associated transcriptional program in renal epithelial progenitors. Development 134(17): 3177-90. PubMed citation; Online text

Schroeder, D. F. and McGhee, J. D. (1998). Anterior-posterior patterning within the Caenorhabditis elegans endoderm. Development 125(24): 4877-87. PubMed Citation: 9811572

Schweizer, L. and Basler, K. (1998). Drosophila ciD encodes a hybrid Pangolin/Cubitus interruptus protein that diverts the Wingless into the Hedgehog signaling pathway. Mech. Dev. 78(1-2): 141-151.

Sheridan, P. L., et al. (1995). Activation of the HIV-1 enhancer by the LEF-1 HMG protein on nucleosome-assembled DNA in vitro. Genes Dev. 9: 2090-2104

Shin, T. H., et al. (1999). MOM-4, a MAP kinase kinase kinase-related protein, activates WRM-1/LIT-1 kinase to transduce anterior/posterior polarity signals in C. elegans. Mol. Cell 4: 275-280

Shtutman, M., et al. (1999). The cyclin D1 gene is a target of the beta-catenin/LEF-1 pathway. Proc. Natl. Acad. Sci. 96(10): 5522-7

Siegfried, K. R. and Kimble, J. (2002). POP-1 controls axis formation during early gonadogenesis in C. elegans. Development 129: 443-453. 11807036

Sierra, J., et al. (2006). The APC tumor suppressor counteracts beta-catenin activation and H3K4 methylation at Wnt target genes. Genes Dev. 20: 586-600. PubMed Citation: 16510874

Song, H., et al. (2010). Coop functions as a corepressor of Pangolin and antagonizes Wingless signaling. Genes Dev. 24(9): 881-6. PubMed Citation: 20439429

Song, Y., Chung, S. and Kunes, S. (2000). Combgap relays Wingless signal reception to the determination of cortical cell fate in the Drosophila visual system. Mol. Cell 6: 1143-1154. PubMed Citation: 11106753

Stadeli, R. and Basler, K. (2005). Dissecting nuclear Wingless signalling: Recruitment of the transcriptional co-activator Pygopus by a chain of adaptor proteins. Mech. Dev. 122(11): 1171-82. 16169192

Streit, A., et al. (2002). Conserved regulation of the Caenorhabditis elegans labial/Hox1 gene ceh-13. Dev. Biol. 242: 96-108. 11820809

Szeto, D. P. and Kimelman, D. (2004). Combinatorial gene regulation by Bmp and Wnt in zebrafish posterior mesoderm formation. Development 131: 3751-3760. 15240553

Takaesu, N. T., Bulanin, D. S., Johnson, A. N., Orenic, T. V. and Newfeld, S. J. (2008). A combinatorial enhancer recognized by Mad, TCF and Brinker first activates then represses dpp expression in the posterior spiracles of Drosophila. Dev. Biol. 313(2): 829-43. PubMed Citation: 18068697

Takemaru, K.-I., et al. (2003). Chibby, a nuclear ß-catenin associated antagonist of the Wnt/Wingless pathway. Nature 422(6934): 905-9. 12712206

Takeshita, H. and Sawa, H. (2005). Asymmetric cortical and nuclear localizations of WRM-1/ß-catenin during asymmetric cell division in C. elegans. Genes Dev. 19: 1743-1748. 16077003

Tang, M., Yuan, W., Bodmer, R., Wu, X. and Ocorr, K. (2014). The role of Pygopus in the differentiation of intracardiac valves in Drosophila. Genesis 52: 19-28. PubMed ID: 24265259

Tetsu, O. and McCormick F. (1999). Beta-catenin regulates expression of cyclin D1 in colon carcinoma cells. Nature 398(6726): 422-6. PubMed ID: 10201372

Tevosian, S. G., et al. (1997). HBP1: a HMG box transcriptional repressor that is targeted by the retinoblastoma family. Genes Dev. 11:383-396. PubMed ID: 9030690

Theodosiou, N. A., and Tabin, C. J. (2003). Wnt signaling during development of the gastrointestinal tract. Dev. Biol. 259: 258-271. 12871700

Thesleff, I. and Sharpe, P. (1997). Signalling networks regulating dental development. Mech. Dev. 67(2): 111-123. PubMed Citation: 9392510

Thompson, B. J., et al. (2004). A complex of Armadillo, Legless, and Pygopus coactivates dTCF to activate Wingless target genes. Curr. Biol. 14: 458-466. 15043810

Thorpe, C. J. and Moon, R. T. (2004). nemo-like kinase is an essential co-activator of Wnt signaling during early zebrafish development. Development 131: 2899-2909. 15151990

Tilmann, C. and Kimble, J. (2005). Cyclin D regulation of a sexually dimorphic asymmetric cell division. Dev. Cell 9: 489-499. 16198291

Tolwinski, N. S. and Wieschaus, E. (2001). Armadillo nuclear import is regulated by cytoplasmic anchor Axin and nuclear anchor dTCF/Pan. Development 128: 2107-2117. PubMed Citation: 11493532

Tran, H., Hamada, F., Schwarz-Romond, T. and Bienz, M. (2008). Trabid, a new positive regulator of Wnt-induced transcription with preference for binding and cleaving K63-linked ubiquitin chains. Genes Dev. 22(4): 528-42. PubMed Citation: 18281465

Tutter, A. V., Fryer, C. J. and Jones, K. A. (2001). Chromatin-specific regulation of LEF-1-ß-catenin transcription activation and inhibition in vitro. Genes Dev. 15: 3342-3354. 11751639

Vacik, T., Stubbs, J. L. and Lemke, G. (2011). A novel mechanism for the transcriptional regulation of Wnt signaling in development. Genes Dev. 25(17): 1783-95. PubMed Citation: 21856776

Valdivia, L. E., (2011). Lef1-dependent Wnt/β-catenin signalling drives the proliferative engine that maintains tissue homeostasis during lateral line development. Development 138(18): 3931-41. PubMed Citation: 21862557

van de Wetering, M., et al. (1997). Armadillo coactivates transcription driven by the product of the Drosophila segment polarity gene dTCF. Cell 88: 789-799. PubMed Citation: 9118222

van de Wetering, M., et al. (2002). The ß-Catenin/TCF-4 complex imposes a crypt progenitor phenotype on colorectal cancer cells. Cell 111: 241-250. 12408868

van Tienen, L. M., Mieszczanek, J., Fiedler, M., Rutherford, T. J. and Bienz, M. (2017). Constitutive scaffolding of multiple Wnt enhanceosome components by Legless/BCL9. Elife 6: e20882. PubMed ID: 28296634

Verheyen, E. M., et al. (2001). The tissue polarity gene nemo carries out multiple roles in patterning during Drosophila development. Mech Dev. 101: 119-32. 11231065

Vleminckx, K., Kemler, R. and Hecht, A. (1999). The C-terminal transactivation domain of beta-catenin is necessary and sufficient for signaling by the LEF-1/beta-catenin complex in Xenopus laevis. Mech. Dev. 81(1-2): 65-74. PubMed Citation: 10330485

Vonica, A., et al. (2000). TCF Is the nuclear effector of the beta-Catenin signal that patterns the sea urchin animal-vegetal axis. Dev. Biol. 217: 230-243. PubMed Citation: 10625549

Von Ohlen, T. and Hooper, J. E. (1999). The ciD mutation encodes a chimeric protein whose activity is regulated by Wingless signaling. Dev. Biol. 208(1): 147-56. PubMed Citation: 10075848

Waltzer, L. and Bienz, M. (1998). Drosophila CBP represses the transcription factor TCF to antagonize Wingless signalling. Nature 395(6701): 521-5. PubMed Citation: 9774110

Wang, H., Lei, Q., Oosterveen, T., Ericson, J. and Matise, M. P. (2011). Tcf/Lef repressors differentially regulate Shh-Gli target gene activation thresholds to generate progenitor patterning in the developing CNS. Development 138(17): 3711-21. PubMed Citation: 21775418

Yang, X. D., et al. (2011). Distinct and mutually inhibitory binding by two divergent β-catenins coordinates TCF levels and activity in C. elegans. Development 138(19): 4255-65. PubMed Citation: 21852394

Ward, E. J. and Berg, C. A. (2005). Juxtaposition between two cell types is necessary for dorsal appendage tube formation. Mech. Dev. 122(2): 241-55. 15652711

Ward, E. J., et al. (2006). Border of Notch activity establishes a boundary between the two dorsal appendage tube cell types. Dev. Biol. 297(2): 461-70. 16828735

Weiske, J. and Huber, O. (2005). The histidine triad protein Hint1 interacts with Pontin and Reptin and inhibits TCF-beta-catenin-mediated transcription. J. Cell. Sci. 118(Pt 14): 3117-29. Medline abstract: 16014379

Xin, N., Benchabane, H., Tian, A., Nguyen, K., Klofas, L. and Ahmed, Y. (2011). Erect Wing facilitates context-dependent Wnt/Wingless signaling by recruiting the cell-specific Armadillo-TCF adaptor Earthbound to chromatin. Development 138(22): 4955-67. PubMed Citation: 22028028

Yan, S.-J., et al. (2004). Multiple signaling pathways and a selector protein sequentially regulate Drosophila wing development. Development 131: 285-298. 14701680

Yang, J., et al. (2002). Repression through a distal TCF-3 binding site restricts Xenopus myf-5 expression in gastrula mesoderm. Mech. Dev. 115: 79-89. 12049769

Yang, X., et al. (2000). decapentaplegic is a direct target of dTcf repression in the Drosophila visceral mesoderm. Development 127: 3695-3702. PubMed Citation: 10934014

Yasumoto, K.-i., et al. (2002). Microphthalmia-associated transcription factor interacts with LEF-1, a mediator of Wnt signaling. EMBO J. 21: 2703-2714. 12032083

Yoon, Y., Klomp, J., Martin-Martin, I., Criscione, F., Calvo, E., Ribeiro, J. and Schmidt-Ott, U. (2019). Embryo polarity in moth flies and mosquitoes relies on distinct old genes with localized transcript isoforms. Elife 8. PubMed ID: 31591963

Young, C. S., et al. (1998). Wnt-1 induces growth, cytosolic beta-catenin, and Tcf/Lef transcriptional activation in Rat-1 fibroblasts. Mol. Cell. Biol. 18(5): 2474-2485. PubMed Citation: 9566868

Zhang, C., Basta, T., Jensen, E. D. and Klymkowsky, M. W. (2003). The beta-catenin/VegT-regulated early zygotic gene Xnr5 is a direct target of SOX3 regulation. Development 130: 5609-5624. Medline abstract: 14522872

Zhang, C. U., Blauwkamp, T. A., Burby, P. E. and Cadigan, K. M. (2014). Wnt-mediated repression via bipartite DNA recognition by TCF in the Drosophila hematopoietic system. PLoS Genet 10: e1004509. PubMed ID: 25144371

Zeng, Y. A. and Verheyen, E. M. (2004). Nemo is an inducible antagonist of Wingless signaling during Drosophila wing development. Development 131: 2911-2920. 15169756

Zhao, X. Sawa, H. and Herman, M. A. (2003). Tcl-2 encodes a novel protein that acts synergistically with Wnt signaling pathways in C. elegans. Dev. Biol. 256: 276-289. 12679102


pangolin: Biological Overview | Evolutionary Homologs | Regulation | Protein Interactions | Developmental Biology | Effects of Mutation

date revised: 15 August 2020 

Home page: The Interactive Fly © 1997 Thomas B. Brody, Ph.D.

The Interactive Fly resides on the
Society for Developmental Biology's Web server.