Vítězslav Bryja, 2014    Attachments      #1      Schulte1  G, Bryja V1 , Rawal N, Castelo‐Branco G, Sousa KM, Arenas E (2005): Purified HA‐ Wnt5a increases differentiation of dopaminergic precursor cells. J. Neurochem 92:1550‐3.  1  equal contribution        Impact factor (2005): 4.500  Times cited (without autocitations, WoS, Feb 21st 2014): 44  Significance: Description of the first purification of the Wnt ligand, which can activate  non‐canonical Wnt signaling. This step allowed biochemical analysis of Dishevelled  in the beta‐catenin‐independent pathways.  Contibution  of  the  author/author´s  team:  Purification  of  Wnt5a  (preparation  of  conditioned  media,  chromatography)  and  functional  validation  in  dopaminergic  cells                RAPID COMMUNICATION Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation Gunnar Schulte,1 Vı´teˇzslav Bryja,1 Nina Rawal, Goncalo Castelo-Branco, Kyle M. Sousa and Ernest Arenas Laboratory of Molecular Neurobiology, Department of Medical Biochemistry and Biophysics, Karolinska Institutet, Stockholm, Sweden Abstract The Wnt family of lipoproteins regulates several aspects of the development of the nervous system. Recently, we reported that Wnt-3a enhances the proliferation of midbrain dopaminergic precursors and that Wnt-5a promotes their differentiation into dopaminergic neurones. Here we report the purification of hemagglutinin-tagged Wnt-5a using a three-step purification method similar to that previously described for Wnt-3a. Haemagglutinin-tagged Wnt-5a was biologically active and induced the differentiation of immature primary midbrain precursors into tyrosine hydroxylase-positive dopaminergic neurones. Using a substantia nigra-derived dopaminergic cell line (SN4741), we found that Wnt-5a, unlike Wnt-3a, did not promote b-catenin phosphorylation or stabilization. However, both Wnt-5a and Wnt- 3a activated dishevelled, as assessed by a phosphorylationdependent mobility shift. Moreover, the activity of Wnt-5a on dishevelled was blocked by pre-treatment with acyl protein thioesterase-1, indicating that palmitoylation of Wnt-5a is necessary for its function. Thus, our results suggest that Wnt-3a and Wnt-5a, respectively, activate canonical and non-canonical Wnt signalling pathways in ventral midbrain dopaminergic cells. Furthermore, we identify dishevelled as a key player in transducing both Wnt canonical and non-canonical signals in dopaminergic cells. Keywords: b-catenin, dopaminergic neurones, proliferation, SN4741, Wnt-3a, Wnt signalling. J. Neurochem. (2005) 92, 1550–1553. Wnts are secreted, lipid-anchored proteins (Nusse 2003) that bind and activate the Frizzled receptor family (Malbon 2004). Palmitoylation of Wnts renders these proteins highly lipophilic (Willert et al. 2003), a fact which has hampered purification of biologically active Wnts for a long time. Recently, however, Willert et al. (2003) published a purification procedure for active Wnt-3a, a method that is applicable to the purification of other members of the Wnt family. Wnt signalling via Frizzled receptors involves several different important players, including the phosphoprotein dishevelled (Dvl), axin, glycogen synthase kinase 3, b-catenin and heterotrimeric and monomeric G proteins and others (Nelson and Nusse 2004). Glycogen synthase kinase 3 constitutively phosphorylates b-catenin, targeting it for proteasome-dependent degradation. Activation of canonical Wnt signalling inhibits glycogen synthase kinase 3 and consequently leads to decreased b-catenin phosphorylation and accumulation of transcriptionally active b-catenin. Non-canonical Wnt signalling pathways include the calcium pathway, involving heterotrimeric G proteins, small GTPases, calcineurin and nuclear factor activated in T-cells (NFAT; Huelsken and Behrens 2002; Wang and Malbon 2004). In addition, non-canonical signalling via the planar cell polarity pathway involves Rho, Rock and JNK (Huelsken and Behrens 2002). However, according to recent studies, Dvl appears to be the element common to all of these pathways as it is activated by Frizzleds irrespective of the signalling pathway (Huelsken and Behrens 2002; Chen et al. 2003; Gonza´lez-Sancho et al. 2004). Although Wnt-1 and Wnt-3a primarily stimulate canonical signalling and Wnt-5a stimulates non-canonical pathways, the specific pharmacological profiles of Wnt binding to its cognate Frizzled receptors remain unknown (Hsieh 2004). In addition to Frizzleds, Wnts have been shown to bind to the low-density lipoprotein-related protein 5 and 6 coreceptor, apparently only necessary for canonical signalling (Gonza´lez-Sancho et al. 2004). The development of dopaminergic neurones in the ventral midbrain requires a complex programme of epigenetic and genetic signals in order to acquire a mature neuronal phenotype (Riddle and Pollock 2003). We have previously shown that the Wnt family plays an important role in the regulation of midbrain dopaminergic development. While Wnt-3a increased the proliferation of precursor cells, Wnt-5a increased the maturation of dopaminergic precursors into dopaminergic neurones (Castelo-Branco et al. 2003). Here we describe the purification of hemagglutinin (HA)-tagged Wnt- 5a and its effects on the differentiation of E14.5 rat ventral midbrain dopaminergic precursor cells. We report that Wnt-5a activates Dvl in the substantia nigra-derived cell line SN4741 (Son et al. 1999). Furthermore, we show that post-translational lipid modification of Wnt-5a is necessary for its activity. Materials and methods Materials The chromatography columns, polyvinylidene difluoride membrane and enhanced chemiluminescence system (ECLplus) were from Amersham Received November 10, 2004; revised manuscript received December 1, 2004; accepted December 2, 2004. Address correspondence and reprint requests to Ernest Arenas, Laboratory of Molecular Neurobiology, Department of Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden. E-mail: ernest.arenas@mbb.ki.se 1 These authors contributed equally to this article. Abbreviations used: Dvl, dishevelled; HA, haemagglutinin; PBS, phosphatebuffered saline. Journal of Neurochemistry, 2005, 92, 1550–1553 doi:10.1111/j.1471-4159.2004.03022.x 1550 Ó 2005 International Society for Neurochemistry, J. Neurochem. (2005) 92, 1550–1553 Biosciences (Uppsala, Sweden). CHAPS was from Amresco (Solon, OH, USA). N-glycosidase F was from Roche (Mannheim, Germany). Cell culture media and foetal calf serum were from Invitrogen (Carlsbad, CA, USA). Recombinant Wnt-3a was purchased from R & D Systems (Minneapolis, MN, USA) and Wnt-5a was a gift from R & D Systems. Primary antibodies were: mouse anti-HA (Covance Princeton, NJ, USA); mouse anti-Wnt-5a (R & D Systems); mouse anti-Dvl3 (Santa Cruz Biotechnology, Santa Cruz, CA, USA); mouse anti-(Ser37/Thr41 dephospho) active b-catenin (Upstate Biotechnology, Lake Placid, NY, USA); mouse anti-b-catenin (BD Biosciences, San Jose, CA, USA) and rabbit anti-Ser33/37/Thr41 phospho-b-catenin (Cell Signaling Technology, Beverly, MA, USA). APT-1 was kindly provided by Dr A. Gilman (University of Texas, Dallas, TX, USA). Purification of haemagglutinin-tagged Wnt-5a Purification of C-terminally-tagged HA-Wnt-5a was performed according to Willert et al. (2003) with minor changes. Briefly, 4 L of conditioned medium from HA-Wnt-5a expressing rat B1a fibroblasts (Shimizu et al. 1997) were harvested and filtered (0.45 lm filter). Before Blue Sepharose affinity chromatography (Blue Sepharose 6 Fast Flow), 1% CHAPS was added and the medium was filtered again (0.2 lm filter). The Blue Sepharose column was equilibrated with 20 mM Tris, 150 mM KCl, 1% CHAPS, pH 7.5 (Buffer A), loaded with the HA-Wnt- 5a-containing medium and washed extensively in Buffer A. Fractions eluted with high salt buffer (20 mM Tris, 1.5 M KCl, 1% CHAPS, pH 7.5) were analysed by immunoblotting for the presence of HA tag (mouse anti-HA) and Wnt-5a protein (anti-Wnt-5a). HA-Wnt-5a-positive fractions were pooled and concentrated in Centricon Plus-80 filters (Millipore Corporation, Bedford, MA, USA) for loading onto a HiLoad 26/60 Superdex 200 gel filtration column using phosphate-buffered saline (PBS), 1% CHAPS, pH 7.5. Eluted fractions were analysed for the presence of HA-Wnt-5a by immunoblotting and further purified with a HiTrap Heparin HP column using PBS, 1% CHAPS, pH 7.5 as loading buffer and PBS, 1 M NaCl, 1% CHAPS, pH 7.5 as elution buffer. The presence of HA-Wnt-5a was confirmed by immunoblotting for HA tag and Wnt-5a protein. Deglycosylation of HA-Wnt-5a was performed in PBS, 1 M NaCl, 1% CHAPS using 3 U N-glycosidase F at 37°C for 90 min. Removal of lipid modification was performed by pre-treatment with APT-1 at 30°C overnight [reaction mix: 100 ng Wnt, 1 lg bovine serum albumin, 19 ng APT-1 in total 15 lL PBS, 1% CHAPS]. Primary cell culture and culture of SN4741 cells Rat ventral midbrain E14.5 precursor cultures were prepared as described previously (Castelo-Branco et al. 2003). Experiments were performed with the approval of the ethical committee (Stockholms Norra Djurfo¨rso¨ksetiska Na¨mnd). Cells were treated with HA-Wnt-5a (100 ng/mL) for 3 days in vitro before fixation. The biological effects of HA-Wnt-5a on precursor cultures were quantified in three independent experiments performed in duplicate and analysed by one-sample Student’s t-test. Data were normalized by setting the control-stimulated values to 100% while experimental values were compared with normalized values and found to be significantly different. SN4741 cells were generously provided by Dr J.H. Son (Son et al. 1999) and grown in Dulbecco’s modified Eagle’s medium, 10% foetal calf serum, L-glutamine (2 mM), penicillin (50 U/mL), streptomycin (50 U/mL), glucose (0.6%). For analysis of intracellular signalling, 100 000 cells/well were seeded in 12-well plates, grown overnight in the absence of serum and stimulated for 2 h in the same medium according to the figure legends. Control stimulations were performed with equivalent volumes of PBS, 1 M NaCl, 1% CHAPS, hereafter referred to as control stimulation. Cells were washed twice in PBS and lysed by addition of 200 lL of 1% sodium dodecyl sulphate, 10% glycerol and 0.1 M Tris-Cl (pH 7.4). Gel electrophoresis and immunoblotting Fractions of the HA-Wnt-5a purification and SN4741 cell lysates denatured with Laemmli buffer were analysed by sodium dodecyl sulphate–polyacrylamide gel electrophoresis. Silver staining was performed according to the Blum silver staining protocol. For immunoblotting, proteins were transferred to polyvinylidene difluoride membranes (Immobilon P). Detection of proteins was performed with the following primary antibodies: anti-HA; anti-Wnt-5a; anti-Dvl3, active b-catenin; b-catenin, phospho-b-catenin and appropriate horseradish peroxidase-conjugated secondary antibodies. Signals were detected with the ECLplus system. Results and Discussion We have previously reported that partially purified conditioned medium from HA-Wnt-5a-expressing fibroblasts promotes the differentiation of dopaminergic precursors into tyrosine hydroxylase-positive neurones (Castelo-Branco et al. 2003). To ascertain the effects of pure Wnt-5a and investigate the mechanism by which Wnt-5a promotes the differentiation of ventral midbrain precursors into dopaminergic neurones, we set out to purify HA-Wnt-5a. The purification protocol was adapted from the recently published report by the group of Roel Nusse (Willert et al. 2003), with minor modifications. Subsequent analysis of HA-Wnt-5aconditioned media and the fractions collected after each purification step revealed an increasing concentration and purity of HA-Wnt-5a (final concentration $10 lg/mL), as detected by immunoblotting with anti-HA tag and anti-Wnt-5a antibodies (Fig. 1a). MALDI-TOF analysis of an in gel trypsin-digested sample from a Coomassie-stained gel containing the HA-positive bands confirmed the identity of Wnt-5a (data not shown). Post-heparin fractions only contained minor trace contaminants. HAWnt-5a migrated in 10% sodium dodecyl sulphate–polyacrylamide gel electrophoresis at a size of $48 kDa. This slow migration was probably attributable to heavy glycosylation as shedding of the glycosylation with N-glycosidase F resulted in a shift in mobility to $40 kDa, reflecting the expected molecular weight of the HA-tagged Wnt-5a (Fig. 1b). The concentration of the final HA-Wnt-5a fractions eluted from the heparin column was determined by comparing the immunoblot signal obtained from anti-Wnt-5a antibody staining with dilution series of mouse nontagged Wnt-5a. The three-step purification process resulted in an HAWnt-5a fraction that was tested for biological activity. Treatment of rat ventral midbrain E14.5 precursor cultures for 3 days with either control buffer (elution buffer from the heparin column; PBS, 1 M NaCl, 1% CHAPS) or the final HA-Wnt-5a fraction increased the number of tyrosine hydroxylase-positive cells by about 2.8-fold compared with control treated cells (Fig. 1d). Thus, our results confirm that the effects of partially purified conditioned media from HA-Wnt-5a-expressing fibroblasts (Castelo-Branco et al. 2003) are indeed due to the presence of biologically active HA-Wnt-5a. These results are in agreement with those obtained by others (Shimizu et al. 1997; Jo¨nsson and Andersson 2001; Mericskay et al. 2004), indicating that HA-Wnt-5a is biologically active in different applications and assays. We next decided to investigate the intracellular signalling pathways activated by HA-Wnt-5a in dopaminergic neurones. For this purpose we utilized the SN4741 immortalized cell line that is derived from an E13.5 mouse tyrosine hydroxylase-positive substantia nigra dopaminergic neurone (Son et al. 1999). This cell line is a suitable model for the investigation of biochemical changes induced by Wnts in immature dopaminergic cells as it expresses both lipoprotein-related protein 5/6 Wnt-5a effects on dopaminergic precursors 1551 Ó 2005 International Society for Neurochemistry, J. Neurochem. (2005) 92, 1550–1553 and several of the frizzled receptors (Fz 1, 2 and 4–9), as assessed by RT-PCR (unpublished results). To investigate intracellular signalling upon Wnt stimulation, we examined the stabilization of b-catenin as well as the phosphorylation and dephosphorylation status of b-catenin using a pan-b-catenin, a phospho-specific b-catenin and the dephospho active b-catenin antibodies, respectively. Wnt-3a treatment, which is known to activate canonical signalling (Willert et al. 2003), led to a strong dephosphorylation of b-catenin and a slight increase in total b-catenin levels in the SN4741 dopaminergic cell line. In contrast, treatment with Wnt-5a did not modify the levels of b-catenin or its phosphorylation state (Fig. 2a). Compared with b-catenin dephosphorylation, the phospho-specific b-catenin antibody revealed an inverted pattern, showing a decrease in labelling upon Wnt-3a stimulation, while Wnt-5a treatment did not affect b-catenin phosphorylation. Wnt-5a treatment did not affect the basal expression and basal phosphorylation pattern of b-catenin (Fig. 2a). Thus, in our system, unlike in the report of Topol et al. (2003), Wnt-5a cannot block canonical signalling. Recent studies show that Dvl becomes hyperphosphorylated upon activation of either the canonical or non-canonical pathway (a) (b) (c) (d) HA-Wnt-5a 160 kDa 105 kDa 75 kDa 50 kDa 35 kDa - + N-glycosidase F HA-Wnt-5a TH / Hoechst conditionedmedium Bluesepharose Heparin Superdex anti-HA anti-Wnt-5a silver stain control Fig. 1 Purification of hemagglutinin (HA)-tagged Wnt-5a and its effects on maturation of dopaminergic cells. (a) Summary of the three purification steps of HA-Wnt-5a. Immunoblots show the presence of HA-Wnt-5a in the fractions collected after each step. Purification was monitored by either anti-HA antibody, anti-Wnt-5a antibody or silver stain (a). HA-Wnt-5a migrated at a molecular weight of $48 kDa. Deglycosylation with N-glycosidase F resulted in a mobility shift giving a signal at the expected molecular weight of $40 kDa (b). Treatment of rat E14.5 VM precursor cultures with HA-Wnt-5a (100 ng/mL) led to an increase in the number of tyrosine hydroxylase (TH)-positive cells after 3 days of treatment as assessed by immunocytochemistry (c and d). (c) HA-Wnt-5a and control treated rat E14.5 VM precursor cultures immunostained with anti-TH and the nuclear counterstain Hoechst-33342. (d) Three independent experiments performed in duplicate. Values were normalized to number of TH-positive cells/Hoechst-33342 in control treated cells (*p < 0.05). Error bars give SEM. P-β-catenin Dvl3 dephospho β-catenin control Wnt-3a HA-Wnt-5a β-catenin Dvl3 control HA-Wnt-5a,4°C HA-Wnt-5a,-80°C Wnt-5a Wnt-5a Wnt-3a APT-1 - - + - + + - + + APT-1 - - + Dvl3 Dvl3 (a) (b) (c) Fig. 2 (a) Wnt signalling in SN4741 cells. Dephosphorylation, phosphorylation and stabilization of b-catenin and phosphorylation of dishevelled (Dvl)3 upon Wnt (100 ng/mL) stimulation were analysed by immunoblotting. (b) Removal of the lipid modification(s) with APT-1 abolished the capability of Wnt-5a and Wnt-3a to activate signalling. (c) Comparison of haemagglutinin (HA)-tagged Wnt-5a stored at 4 and )80°C and the non-tagged Wnt-5a with regard to their capability to phosphorylate Dvl3. For each panel, one experiment out of at least three is shown. 1552 G. Schulte et al. Ó 2005 International Society for Neurochemistry, J. Neurochem. (2005) 92, 1550–1553 (Gonza´lez-Sancho et al. 2004). Indeed, both Wnt-3a and HA-Wnt-5a induced a shift of fast-migrating unphosphorylated Dvl to slower migrating phosphorylated Dvl as detected by immunoblotting with an anti-Dvl3 antibody (Fig. 2a). A remaining open question is whether Dvl is differentially phosphorylated by Wnt-3a and Wnt-5a. It has been shown recently that differential phosphorylation of Dvl by casein kinase e might either lead to stabilization of b-catenin or activation of jun-Nterminal kinase (Cong et al. 2004). However, our initial screen for possible downstream targets of non-canonical Wnt signalling in SN4741 cells using diverse phospho-specific antibodies (P-c-jun, P-JNK, P-PKC, P-ERK1/2, P-CaMKII, P-PKB/Akt, P-p38 and P-glycogen synthase kinase 3) has not given positive results (data not shown) but restricted the number of possible downstream signalling components. Thus, our results show that HA-Wnt-5a is biologically active and capable of phosphorylating Dvl but does not regulate b-catenin levels or phosphorylation. These results suggest that Wnt-5a works on dopaminergic cells by activating non-canonical pathways rather than canonical Wnt signalling. In order to investigate whether the biological activity of Wnt-5a is dependent on post-translational lipid modification, we treated Wnt-5a with APT-1, an enzyme that removes lipid moieties in proteins and has been previously reported to result in the loss of biological activity of Wnt-3a (Willert et al. 2003). We found that the shedding of the lipid modification of Wnt-3a or Wnt-5a abolished the capacity of both Wnts to induce Dvl phosphorylation in SN4741 cells (Fig. 2b), suggesting that the palmitolate is essential for the biological activity of Wnts in dopaminergic cells. Finally, we examined the stability of the HA-Wnt-5a protein and compared its ability to induce Dvl3 phosphorylation with non-HA-tagged Wnt-5a. Interestingly, we found that the presence of the HA tag or storage at 4 or )80°C did not reduce the activity of Wnt-5a (Fig. 2c). In summary, we report for the first time the purification of biologically active and stable HA-Wnt-5a from fibroblast-conditioned media. The purified protein induced the differentiation of dopaminergic precursors and activated intracellular signalling pathways in dopaminergic cells by a mechanism requiring the integrity of the lipid modification of Wnt-5a. Moreover, our data indicate that Wnt-5a, unlike Wnt-3a, does not activate the Wnt canonical signalling pathway, suggesting that the effects of Wnt-5a on dopaminergic neurones are mediated by non-canonical pathways. In agreement with this, it has been reported that Wnt-5a signals in other systems via non-canonical pathways (Ku¨hl et al. 2000; Ishitani et al. 2003). Our results identify dopaminergic neurones as a useful model for studying and understanding cell signalling by Wnts. Future studies will aim to identify the receptors and pathway(s) activated by Wnt-5a in dopaminergic neurones and the mechanisms by which Dvl distinguishes between and mediates both canonical and non-canonical signalling pathways. Acknowledgements We wish to thank Drs Jawed Shafqat, Hans Jo¨rnvall, Carina Palmberg and Tomas Bergman for fruitful discussions, advice and help in the purification and sequencing of HA-Wnt-5a; Dr J.H. Son for providing SN4741 cells; Patricia Kelly (R & D Systems) for providing Wnt-5a; Dr Anita Hall for critical reading of the manuscript and Lena Amaloo and Claudia Tello for additional assistance. Financial support was obtained from the Swedish Foundation for Strategic Research, Swedish Royal Academy of Sciences, Knut and Alice Wallenberg Foundation, European Commission, Swedish MRC, Karolinska Institutet and Lars Hiertas Minnesfond. GS was supported by a post-doctoral fellowship from the Swedish Society for Medical Research. GC-B was supported by the Praxis XXI programme of the Portuguese Fundac¸a˜o para a Cieˆncia e Tecnologia/European Social Fund, the Karolinska Institute and Calouste Gulbenkian Foundation. References Castelo-Branco G., Wagner J., Rodriguez F. J., Kele J., Sousa K., Rawal N., Pasolli H. A., Fuchs E., Kitajewski J. and Arenas E. (2003) Differential regulation of midbrain dopaminergic neuron development by Wnt-1, Wnt-3a, and Wnt- 5a. Proc. Natl Acad. Sci. USA 100, 12 747–12 752. Chen W., ten Berge D., Brown J., Ahn S., Hu L. A., Miller W. E., Caron M. G., Barak L. S., Nusse R. and Lefkowitz R. J. (2003) Dishevelled 2 recruits b-arrestin 2 to mediate Wnt5A-stimulated endocytosis of Frizzled 4. Science 301, 1391–1394. Cong F., Schweizer L. and Varmus H. (2004) Casein kinase I e modulates the signaling specificities of dishevelled. Mol. Cell Biol. 24, 2000–2011. Gonza´lez-Sancho J. M., Brennan K. R., Castelo-Soccio L. A. and Brown A. M. (2004) Wnt proteins induce dishevelled phosphorylation via an LRP5/6independent mechanism, irrespective of their ability to stabilize beta-catenin. Mol. Cell Biol. 24, 4757–4768. Hsieh J. C. (2004) Specificity of WNT–receptor interactions. Front. Biosci. 9, 1333–1338. Huelsken J. and Behrens J. (2002) The Wnt signalling pathway. J. Cell Sci. 115, 3977–3978. Ishitani T., Kishida S., Hyodo-Miura J., Ueno N., Yasuda J., Waterman M., Shibuya H., Moon R. T., Ninomiya-Tsuji J. and Matsumoto K. (2003) The TAK1-NLK mitogen-activated protein kinase cascade functions in the Wnt- 5a/Ca2+ pathway to antagonize Wnt/b-catenin signaling. Mol. Cell Biol. 23, 131–139. Jo¨nsson M. and Andersson T. (2001) Repression of Wnt-5a impairs DDR1 phosphorylation and modifies adhesion and migration of mammary cells. J. Cell Sci. 114, 2043–2053. Ku¨hl M., Sheldahl L. C., Malbon C. C. and Moon R. T. (2000) Ca(2+)/calmodulindependent protein kinase II is stimulated by Wnt and Frizzled homologs and promotes ventral cell fates in Xenopus. J. Biol. Chem. 275, 12 701– 12 711. Malbon C. C. (2004) Frizzleds: new members of the superfamily of G-proteincoupled receptors. Front. Biosci. 9, 1048–1058. Mericskay M., Kitajewski J. and Sassoon D. (2004) Wnt5a is required for proper epithelial–mesenchymal interactions in the uterus. Development 131, 2061– 2072. Nelson W. J. and Nusse R. (2004) Convergence of Wnt, beta-catenin, and cadherin pathways. Science 303, 1483–1487. Nusse R. (2003) Wnts and Hedgehogs: lipid-modified proteins and similarities in signaling mechanisms at the cell surface. Development 130, 5297–5305. Riddle R. and Pollock J. D. (2003) Making connections: the development of mesencephalic dopaminergic neurons. Brain Res. Dev. Brain Res. 147, 3–21. Shimizu H., Julius M. A., Giarre´ M., Zheng Z., Brown A. M. and Kitajewski J. (1997) Transformation by Wnt family proteins correlates with regulation of b-catenin. Cell Growth Differ. 8, 1349–1358. Son J. H., Chun H. S., Joh T. H., Cho S., Conti B. and Lee J. W. (1999) Neuroprotection and neuronal differentiation studies using substantia nigra dopaminergic cells derived from transgenic mouse embryos. J. Neurosci. 19, 10–20. Topol L., Jiang X., Choi H., Garret-Beal L., Carolan P. J. and Yang Y. (2003) Wnt- 5a inhibits the canonical Wnt pathway by promoting GSK-3-independent b-catenin degradation. J. Cell Biol. 162, 899–908. Wang H. Y. and Malbon C. C. (2004) Wnt-frizzled signaling to G-protein-coupled effectors. Cell Mol. Life Sci. 61, 69–75. Willert K., Brown J. D., Danenberg E., Duncan A. W., Weissman I. L., Reya T., Yates J. R. and Nusse R. (2003) Wnt proteins are lipid-modified and can act as stem cell growth factors. Nature 423, 448–452. Wnt-5a effects on dopaminergic precursors 1553 Ó 2005 International Society for Neurochemistry, J. Neurochem. (2005) 92, 1550–1553 Vítězslav Bryja, 2014    Attachments      #2      Bryja  V1 ,  Schulte  G1 ,  Arenas  E  (2007):  Wnt‐3a  utilizes  a  novel  low  dose  and  rapid  pathway  that  does  not  require  casein  kinase  1‐mediated  phosphorylation  of  Dvl  to  activate ‐catenin. Cell. Signal. 19: 610‐616.   1  equal contribution      Impact factor (2007): 4.305  Times cited (without autocitations, WoS, Feb 21st 2014): 20  Significance: Identification of an alternative Wnt‐3a‐induced pathway activating beta‐ catenin. This pathway (restricted to some cell types) is rapid, sensitive and does  not lead to an apparent activation of Dishevelled phosphorylation shift.  Contibution of the author/author´s team: Together with Gunnar Schulte full design,  performance and interpretation of the experiments.                Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate β-catenin Vítězslav Bryja 1 , Gunnar Schulte 1,2 , Ernest Arenas ⁎ Karolinska Institutet, Department Medical Biochemistry and Biophysics, Laboratory Molecular Neurobiology, S-171 77 Stockholm, Sweden Received 18 May 2006; received in revised form 21 August 2006; accepted 21 August 2006 Available online 30 August 2006 Abstract The current view of canonical Wnt signalling is that following Wnt binding to its receptors (Frizzled-Lrp5/6), dishevelled (Dvl) becomes hyperphosphorylated, and the signal is transduced to the APC-GSK3β-axin-β-catenin multiprotein complex, which subsequently dissociates. Asa result β-catenin is not phosphorylated, escapes proteosomal degradation and activates its target genes after translocation to the nucleus. Here, we analyzed the importance of the Wnt-3a-induced phosphorylation and shift in electrophoretic migration of Dvl (PS-Dvl) for the activation of β-catenin. Analysis of Wnt-3a time- and dose-responses in a dopaminergic cell line showed that β-catenin is activated rapidly (within minutes) and at a low dose of Wnt-3a (1 ng/ml). Surprisingly, PS-Dvl appeared only after 30 min and at greater doses (≥20 ng/ml) of Wnt-3a. Moreover, we found that a casein kinase 1 inhibitor (D4476) or siRNA for casein kinase 1 δ/ε (CK1δ/ε) blocked the Wnt-3a-induced PS-Dvl. Interestingly, CK1 inhibition or siRNA for CK1δ/ε did not ablate the activation of β-catenin by Wnt-3a, indicating that there is a PS-Dvl-independent path to activate β-catenin. The increase in β-catenin activation by Wnt-3a (PS-Dvl-dependent or -independent) were blocked by Dickkopf1 (Dkk1), suggesting that the effect of Wnt-3a is in both cases mediated by Lrp5/6 receptors. Thus, our results show that Wnt-3a rapidly induce a partial activation of β-catenin in the absence of PS-Dvl at low doses, while at high doses induce a full activation of β-catenin in a PS-Dvl-dependent manner. © 2006 Elsevier Inc. All rights reserved. Keywords: Wnt-3a; Dishevelled; Dopaminergic cells; Casein kinase 1; Activation of β-catenin; Canonical Wnt signalling 1. Introduction The Wnt signaling pathway is one of the most evolutionary conserved biochemical signalling pathways and is involved in a vast array of processes including embryonic development, adult tissue homeostasis and cancer (for review see [1–3]). The Wnts are secreted glyco-lipoproteins that bind to their membrane receptors (Frizzleds) and their co-receptors (Lrp5/6). The primary signalling components that transduce the Wnt signal are cytoplasmic proteins such as Dishevelled (Dvl), and the degradation complex, formed by Axin, APC, and GSK3β. Upon Wnt stimulation, Dvl is phosphorylated and the degradation complex is inhibited leading to the stabilization of β-catenin. β-catenin can then act in the cytosol, or in the nucleus, by binding to TCF/ LEF and regulating transcription (for additional information see Wnt homepage [4]). The current view of Wnt signalling is that canonical Wnt binds to Frizzled-Lrp5/6 containing receptor complexes, Dvl becomes hyperphosphorylated, and then the signal is transduced to the GSK3β-axin-β-catenin complex, which dissociates and consequently β-catenin is not phosphorylated and becomes activated. Intriguingly, some crucial details of the molecular pathways induced by Wnts in mammalian cells are still not fully understood. One of the most elusive parts of the pathway is the link between phosphorylated Dvl and GSK3β inhibition resulting in the activation and stabilization of β-catenin. It was shown previously in mammalian cells that canonical Wnts are able to induce the phosphorylation of Dvl and the activation of β-catenin [5]. However, the mutual link between these two signalling events induced by canonical Wnts was not investigated in detail. We previously found that dopaminergic neurons respond to Wnt signalling [6,7] and we now take advantage of the SN4741 dopaminergic neuronal cell line [8] as a model system. SN4741 Cellular Signalling 19 (2007) 610–616 www.elsevier.com/locate/cellsig ⁎ Corresponding author. Tel.: +46 8 52487663; fax: +46 8 34 1960. E-mail address: ernest.arenas@ki.se (E. Arenas). 1 both authors contributed equally to this work. 2 present address: Karolinska Institutet, Dept. Physiology and Pharmacology, Receptor Biology and Signaling, S-17177 Stockholm, Sweden. 0898-6568/$ - see front matter © 2006 Elsevier Inc. All rights reserved. doi:10.1016/j.cellsig.2006.08.011 cells, as midbrain dopaminergic neurons [9,10], express a complex battery of Wnt signaling components (N. Rawal and G. Schulte, unpublished) and constitute a good model to study molecular mechanisms of the Wnt signalling [9,10]. We hereby report that Wnt-3a induces an unexpectedly rapid activation of β-catenin that requires Lrp5/6 and low doses of Wnt-3a, but does not require the casein kinase-1-mediated Dvl phosphorylation. However, at late time points and at high doses, Wnt-3a potentiated the initial activation of β-catenin by a Dvl phosphorylation-dependent mechanism. Thus our findings suggest that β-catenin can be activated by two subsequent mechanisms that are independent or dependent of casein kinase 1 activity and PS-Dvl. 2. Materials and methods 2.1. Cell culture and treatments SN4741 cells were generously provided by Dr. J. H. Son [8] and grown in DMEM, 10% FCS, L-glutamine (2 mM), penicillin (50 U/ml), streptomycin (50 U/ml), glucose (0.6%) (all from Gibco Life Sciences). For analysis of intracellular signaling, 40,000 cells/well were seeded in 24 well plates, grown overnight in the absence of serum and in the same medium stimulated with recombinant mouse Wnt-3a (R&D Systems) for 2 h according to the figure legends. Control stimulations were done with equivalent volumes of PBS, 1% CHAPS. Treatment of SN4741 cells with poorly cell-permeable D4476 was performed in the presence of 1 μl of FuGENE 6 Transfection Reagent (Roche) per 200 μl of culture media. 2.2. Western blotting For Western blot analysis, cell samples were prepared as follows: Cells on culture dish were washed with PBS (pH 7.4) and lysed in 1× Laemmli buffer. Equal amounts of cell lysate were subjected to 7.5% SDS PAGE, electrotransferred onto Hybond-P membrane, immunodetected using appropriate primary and secondary antibodies, and visualized by ECL+Plus reagent according to manufacturer′s instructions (Amersham). When required, membranes were stripped in 62.5 mM Tris/HCl (pH 6.8), 2% SDS, and 100 mM β-mercaptoethanol, washed, and reblotted with another antibody. When required, the intensities of signals were assessed by densitometry using Scion Image Beta 4.02 software (Scion Corporation). After immunodetection, each membrane was stained by amidoblack to confirm equal protein loading. The antibodies used were as follows: anti-Dvl1 (sc-7397), anti-Dvl2 (sc-13974), anti-Dvl3 (sc-8027), anti-Myc (sc-40) and antiCK1ε (sc-6471); anti-β-catenin was from BD Bioscences; anti-active-β-catenin (anti-ABC) was from Upstate Biotechnology; anti-β-actin (Ab6276) was from Abcam; anti-phospho-β-catenin (Ser33/37/Thr41) and anti-phospho-β-catenin (Thr41/Ser45) were from Cell Signaling Technology and anti-phospho-β-catenin (Ser45) was from Biosource. 2.3. RNA interference and cell transfections SN4741 cells were transfected with siRNA using neofection (Ambion). siRNAs against individual isoforms of mouse CK1 were purchased from AmbionCK1δ (#88298) and CK1ε (II #188528). Silencer® Negative Control siRNA (#4635, Ambion) was used as a negative control. For transfection into SN4741 cell, siRNAs (0.75 μl of 20 μM) were mixed with Lipofectamine 2000 (2 μl; Invitrogene) and OptiMEM (47.25 μl; Gibco).incubated for 20 min at RT, and the mixture (50 μl) was added to a freshly trypsinized suspension of 25 000 cells in 500 μl of complete media (in 24-well plates). The final concentration of the siRNA in each well was 30 nM. The culture media was exchanged after 5 h and 48 h post transfection the cells were treated with Wnt-3a and collected for further analyses. The efficiency of the silencing was assessed by Western blotting. For plasmid transfection 40000–60000 cells/well were seeded (in 24 well plates) and grown overnight. SN4741 cells were transfected with 1 μg of DNA and 2 μl of Lipofectamine 2000 (Invitrogen) per well under serum-free conditions, according to manufacturer′s recommendations. Medium was changed 4 h post transfection and cells were grown in complete culture medium for another 24 h for further analysis by Western Blotting. 3. Results 3.1. Wnt-3a induces activation of β-catenin and phosphorylation of Dvl In order to characterize the pathways activated by Wnt-3a in dopaminergic cells we treated SN4741 cells with Wnt-3a (100 ng/ml) for 2 h. To monitor the activation of canonical Wnt-signaling we have used antibodies recognizing different phosphorylation states of β-catenin: a) active β-catenin (ABC) antibody [11] specific to forms of β-catenin dephosphorylated on GSK3β target sites (Ser37 or Thr41); and b) phospho-βcatenin (p-β-catenin) antibody specific to the forms of β-catenin targeted for degradation by GSK3β-mediated phosphorylation at Ser33, Ser37, or Thr41. As shown in Fig. 1A, 2 h treatment with Wnt-3a induced an increase in active β-catenin (ABC) and a decrease in phosphorylated β-catenin. However, activation of β-catenin did not lead to any detectable change in the level of total β-catenin after 2 h. Treatment with Wnt-3a also lead to the phosphorylation of Dvl, which is detected as a mobility shift of the protein on SDS-PAGE [5,12]. As other phosphorylation events may not result in a mobility shift of Dvl, in the rest of the manuscript we will refer to the slowly migrating form of Dvl as PS-Dvl (phosphorylated and shifted Dvl). As shown in Fig. 1B all three Dvl isoforms (Dvl1, Dvl2, Dvl3) become phosphorylated following Wnt-3a treatment. However, although all Dvl1, Dvl2 and Dvl3 are expressed in SN4741 cells, the amount of Dvl1 is just above Western blotting detection limit and is thus difficult to detect. In further work we therefore analyzed Dvl1 only in the experiments where information about all three Dvl isoforms was needed for the conclusions. In order to confirm that non-phosphorylated β-catenin also results in the stabilization of β-catenin, we analyzed later time points up to 24 h post stimulation. A significant increase in total β-catenin levels was detected after 6 h, the time at which the levels of active β-catenin also peaked (Fig. 1C). β-catenin levels remained high for up to 24 h, demonstrating that Wnt-3a induces not only increased levels of dephosphorylated β-catenin but also leads to the stabilization of β-catenin. 3.2. The activation of β-catenin precedes the phosphorylationdependent Dvl mobility shift We next examined in more detail the temporal dynamics of the Dvl phosphorylation mobility shift and the β-catenin activation by 100 ng/ml Wnt-3a. As we show in Fig. 2A, Wnt-3a was able to induce active β-catenin within few minutes and the amount of active β-catenin gradually increased during the 2 h of treatment (Fig. 2B). Importantly, all three Dvl isoforms showed the same characteristic phosphorylationdependent mobility shift, which was first detected at 30 min and peaked by 1–2 h of treatment (Fig. 2B). These data are 611V. Bryja et al. / Cellular Signalling 19 (2007) 610–616 compatible with the current Wnt signaling model in which Dvl phosphorylation mediates β-catenin activation, since the Dvl mobility shift and the second increase in active β-catenin coincide in time. However, our findings also showed that Wnt- 3a could also induce a rapid activation β-catenin at an earlier time-point by a mechanism that did not involve PS-Dvl. 3.3. The activation of β-catenin occurs at lower doses of Wnt- 3a than the phoshorylation-dependent Dvl mobility shift As a second parameter we examined whether the stimulation of SN4741 cells for 2 h resulted in a dose dependent activation of β-catenin and/or phosphorylation-dependent Dvl mobility shift. As shown in Fig. 2C, Wnt-3a induced non-phosphorylated active β-catenin (ABC) at very low doses (0.3–1 ng/ml). Activation of β-catenin reached a plateau at a dose of 1 ng/ml that was maintained up to a dose of 10 ng/ml. However, the activation of β-catenin only reached maximal levels by increasing the dose to at least 30 ng/ml of Wnt-3a. Importantly only this higher dose of Wnt-3a leads to a phosphorylationdependent Dvl mobility shift of all three Dvl isoforms (Dvl1, Dvl2, Dvl3). Quantification of active β-catenin in the dose response experiment showed a clear two step activation pattern (Fig. 2D). These experiments, together with the time-course analysis suggested that Wnt-3a could perhaps activate β-catenin by two separate mechanisms: A) rapid (1–15 min) and/or low dose (0.3–3 ng/ml) Wnt-3a that does not involve the phosphorylation-dependent Dvl mobility, and B) a second one that requires a longer activation (≥30 min) and/or higher Wnt- 3a dose (N10 ng/ml) and involves PS-Dvl. We therefore decided to examine whether Wnt-3a-induced β-catenin activation requires PS-Dvl. 3.4. CK1δ/ε is required for Wnt-3a-induced PS-Dvl It has been previously shown that endogenous casein kinase 1 (CK1) mediates the phosphorylation of Dvl in response to canonical Wnts [13–15]. We thus wanted to identify whether CK1 was also responsible for the Wnt-3a-induced PS-Dvl in SN4741 cells. Using an RNAi approach we blocked CK1δ and CK1ε, the two isoforms of CK1 that phosphorylate Dvl and positively regulate canonical Wnt signalling [13,16,17]. As we show in Fig. 3, RNAi mediated knockdown of CK1δ/ε significantly reduced the level of PS-Dvl2 in response to Wnt-3a, suggesting that CK1δ/ε is necessary for PS-Dvl induction in response to Wnt-3a. Similar results were observed for Wnt-5a (VB, GS and EA, unpublished). Importantly, although CK1δ/ε siRNAs clearly attenuated the induction of PS-Dvl2 by Wnt-3a, there was no effect on the activation of βcatenin, as measured by ABC antibody (Fig. 3). This result suggested that PS-Dvl and CK1δ/ε might be dispensable for the activation of β-catenin by Wnt-3a. However, since PS-Dvl was only partially blocked by CK1δ/ε siRNA we used pharmacological inhibitors to block the function of CK1 and determine the function PS-Dvl. 3.5. D4476, a casein kinase 1 (CK1) inhibitor, blocks the phosphorylation-dependent mobility shift of endogenous Dvl (PS-Dvl) but allows the activation of β-catenin by Wnt-3a Since Dvl has been previously identified as an upstream regulator of β-catenin, we sought to determine the relative contribution of PS-Dvl to the activation of β-catenin. We therefore examined whether a known CK1δ inhibitor, D4476 (4-[4-(2,3- dihydro-benzo[1,4]dioxin-6-yl)-5-pyridin-2-yl-1H-imidazol-2yl]benzamide) [18], also inhibited the effects of CK1δ/ε on PSDvl. We utilized an experimental system based on the overexpression of CK1ε and Dvl2-Myc. We found that CK1 induced a phosphorylation-dependent mobility shift of Dvl2-Myc (Fig. 4A, compare lanes 1 and 3). When D4476 is added (lanes 2 and 4) the effect of CK1ε on Dvl2-Myc is efficiently inhibited, demonstrating that D4476 is able to block CK1ε-mediated phosphorylation of Dvl. These data demonstrate that D4476 is an efficient inhibitor of CK1δ/ε activity and can thus serve as useful and specific tool for the analysis of the role of PS-Dvl in β-catenin activation. We therefore blocked Wnt-3a-induced PS-Dvl formation Fig. 1. Wnt-3a activates β-catenin and Dvl in SN4741 cells. (A, B) SN4741 cells were treated with Wnt-3a (100 ng/ml) or with control solution, and were lysed after 2 h of stimulation. The activation of β-catenin (A) was determined by Western blotting using antibodies against active (Ser33/37-dephosphorylated) (ABC), phosphorylated (phospho Ser33/37/Thr41) (p-β-catenin) and total β-catenin. (B) The phosphorylation of Dvl isoforms was detected as phosphorylation-dependent mobility shift of total Dvl1, Dvl2 and Dvl3 using Dvl isoform-specific antibodies. The position of dephosphorylated (full triangle) or phosphorylated and shifted Dvl (PS-Dvl, open triangle) are indicated. (C) SN4741 cells were treated with Wnt-3a (100 ng/ml) and were lysed after 0, 3, 6, 12 and 24 h. The level of ABC, total β-catenin and the phosphorylation of Dvl2 were determined as in part A and B. Data in A–C are representative of at least three independent replicates. 612 V. Bryja et al. / Cellular Signalling 19 (2007) 610–616 pharmacologically using D4476. As shown in Fig. 4B, D4476 dose-dependently inhibited both basal and Wnt-3a-induced phosphorylation-dependent Dvl mobility shifts. However, under the conditions shown in Fig. 4B (50 ng/ml Wnt-3a for 2 h), the efficiency of the inhibition of Dvl-phosphorylation was incomplete (see lanes 7 and 8) even at high doses of D4476. Since our findings indicated that the shift of Dvl was dose-dependent and maximal at already 30 ng/ml (Fig. 2D), we lowered the dose of Wnt-3a to 20 ng/ml for 2 h and found that the Wnt-3a-induced Dvl2 and Dvl3 phosphorylation-dependent mobility shift was now efficiently blocked by 100 μM D4476 (Fig. 4C). CK1 was previously shown to be involved not only in Dvl phosphorylation but also in the phosphorylation of β-catenin. Importantly, CK1α was found to be a kinase for β-catenin that phosphorylates its Ser45 residue thereby priming β-catenin for subsequent phosphorylation by GSK3β [19–21] and degradation. Importantly, we found that under the conditions described above (20 ng/ml Wnt-3a for 2 h), D4476 completely blocked both basal as well as Wnt-3a-induced phosphorylation of all three Dvl isoforms — Dvl 1, Dvl2, and Dvl3, but still allowed the activation of β-catenin by Wnt-3a (Fig. 4D). This was shown by the high levels of ABC (non-phosphorylated Ser37/ Thr41β-catenin) and the low signal by anti-phospho-Ser33/37/ Thr41-β-catenin antibodies after Wnt-3a stimulation. As we further show by phosphospecific antibodies against phosphoSer45-and phospho-Thr41/Ser45-β-catenin, the phosphorylation at the CK1α target site of β-catenin (Ser45) is significantly decreased following D4476 treatment suggesting that CK1α is efficiently inhibited by D4476 (Fig. 4D). Importantly the level of CK1α-dependent phosphorylation of β-catenin at Thr41/ Ser45 is not changed by Wnt-3a in the presence of D4476, indicating that the blocking is not at the level of β-catenin. The Fig. 2. Time-course and dose-response of the activation of β-catenin and Dvl by Wnt-3a. (A) SN4741 cells were treated with Wnt-3a (100 ng/ml) and were lysed after 0, 1, 5, 10, 15, 30, 60 and 120 min. (B) Quantification of ABC signal from the time response experiment. The signal was quantified using densitometry and the intensity of the signal at 120 min was normalized to 1. (C) SN4741 cells were treated with increasing doses of Wnt-3a, and were lysed after 2 h of stimulation. (D) Quantification of ABC signal from dose response experiments (n=3). The intensity of the signal in the sample treated with 300 ng/ml Wnt-3a was normalized to 1. In A, C the activation of β-catenin was determined by Western blotting using antibodies against active (Ser33/37-dephosphorylated) β-catenin (ABC). The phosphorylation of Dvl1, Dvl2 and Dvl3 was detected as phosphorylation-dependent mobility shift with anti-Dvl1-, anti-Dvl2- and antiDvl3-specific antibodies. The position of dephosphorylated (full triangle) and phosphorylated (PS-Dvl, open triangle) Dvl is indicated. Data are representative of at least three independent replicates. Fig. 3. CK1δ/ε induces PS-Dvl in response to Wnt-3a. SN4741 cells were transfected with control siRNA and a mixture of siRNAs against CK1δ and CK1ε and treated with control, 10 and 20 ng/ml of Wnt-3a. The phosphorylation of Dvl isoforms was detected as phosphorylation-dependent mobility shift of total Dvl2. The position of non-phosphorylated (PS-Dvl, full triangle) and phosphorylated Dvl (open triangle) is indicated. The activation of β-catenin was determined by Western blotting using antibodies against active (Ser33/37dephosphorylated) β-catenin (ABC). The level of CK1ε and β-actin were determined to confirm the efficiency of the knockdown and protein loading, respectively. The experiment was performed three times with comparable results. 613V. Bryja et al. / Cellular Signalling 19 (2007) 610–616 expression level of CK1ε was also not modulated by Wnt-3a and D4476 treatment. In summary, these results indicate that the two events, the Dvl phosphorylation shift and activation of βcatenin can occur independently even at higher Wnt-3a doses. Thus, our results suggest that Wnt-3a can activate β-catenin in the absence or the presence of PS-Dvl. We therefore set to determine whether the activation of β-catenin in the absence or the presence of PS-Dvl required the participation of Lrp5/6 and was therefore blocked by Dkk1. 3.6. Dkk1 blocks the activation of β-catenin by Wnt-3a in the absence or the presence of PS-Dvl In order to further characterize the mechanism by which Wnt- 3a activates β-catenin, we treated SN4741 cells with Dkk1 — an inhibitor of canonical Wnt signaling [22], which blocks the signaling activity of Lrp5 and Lrp6 co-receptors [23,24]. Dkk1 efficiently reduced the activation of β-catenin, leaving the phosphorylation-dependent mobility shift of Dvl by Wnt-3a intact (Fig. 5A). These data suggested that Dkk1 can inhibit the activation of β-catenin by Wnt-3a in the presence of PS-Dvl. We next examined whether Dkk1 can block active β-catenin by in the absence of the Dvl mobility shift, in the presence of the CK1 inhibitor, D4476 (100 μM). Interestingly, we found that Dkk1 dose-dependently blocked the activation of β-catenin by 20 ng/ml Wnt-3a (Fig. 5B). These results indicated that the regulation of active β-catenin depends on signalling via the Lrp5/6 coreceptors but occurred independently PS-Dvl. Finally, we examined whether the Dkk1-sensitive and CK1-dependent Wnt-3a-induced pathways cooperate in the activation of β-catenin. Treatment of SN4741 cells with either Dkk1 or D4476 induced a modest reduction in the activation of β-catenin by 50 ng/ml Wnt-3a (Fig. 5C and D). However, their effects were additive and combined treatment with Dkk1 and D4476 induced a greater Fig. 4. CK1 inhibition efficiently blocks Wnt-3a-induced phosphorylation of Dvl. (A) CK1 inhibitor, D4476 (100 μM) blocks the phosphorylation dependent shift of Dvl2-Myc induced by overexpression of CK1ε in SN4741 cells. The position of non-phosphorylated (full triangle) and phosphorylated (open triangle) Dvl2 is indicated. (B) SN4741 cells were treated with control or Wnt-3a (50 ng/ml) for 2 h in the presence or absence of increasing doses of D4476. (C) SN4741 cells were treated with control solution, 100 ng/ml of Wnt-3a or 20 ng/ml of Wnt-3a for 2 h in the presence or absence of D4476 (100 μM). In B, C cells were lysed and the phosphorylation of Dvl2 and Dvl3 was detected as phosphorylation-dependent mobility shift. (D) SN4741 cells were treated with control solution or Wnt-3a (20 ng/ ml) for 2 h in the presence or absence of D4476 (100 μM). Cells were lysed and phosphorylation of β-catenin was determined by Western blotting using specific antibodies against active (Ser33/37- dephosphorylated) β-catenin (ABC), phosphoSer33/37/Thr41, phosphoSer45, phosphoThr41/Ser45 and total β-catenin. The phosphorylation of Dvl1, Dvl2 and Dvl3 was detected as phosphorylation dependent mobility shift. Presence of CK1ε was confirmed using a CK1ε specific antibody. The position of dephosphorylated (full triangle) and phosphorylated (PS-Dvl, open triangle) Dvl is indicated. Data in A–D are representative of at least three independent replicates. 614 V. Bryja et al. / Cellular Signalling 19 (2007) 610–616 decrease in active β-catenin. These results suggest that CK1mediated Dvl phosphorylation and Dkk1-mediated pathways cooperate to regulate the activation of β-catenin by Wnt-3a. 4. Discussion Canonical Wnts (e.g. Wnt-3a) transduce their signal into the cell by inducing the phosphorylation of Dvl, which in turn results in inhibition of GSK3β preventing thus the phosphorylation of β-catenin. Although the precise mechanism connecting Dvl phosphorylation with GSK3β inhibition is not known (for review see e.g. [25,26]), these events are considered to be sequential. Here we show using siRNA and pharmacological inhibitors that Wnt-3a-induced PS-Dvl is dependent on CK1δ/ε. Using these approaches we analyzed the role the role of PS-Dvl in β-catenin activation. Our findings suggest that Wnt-3a induces the activation of β-catenin in a PS-Dvl-dependent manner. However, we also found that β-catenin can also be activated by Wnt-3a by a mechanism that is independent from the CK1δ/ ε-mediated phosphorylation of Dvl and from PS-Dvl. This pathway is activated very rapidly (in 1 minute) at high doses of Wnt-3a (100 ng/ml) or by constant exposure to low doses of Wnt-3a (0.3–3 ng/ml for 2 h), two modalities of Wnt signalling that are likely to be very relevant in vivo. Interestingly, this rapid/low Wnt-3a dose PS-Dvl-independent pathway, is potentiated by a mechanism involving PS-Dvl that is activated with a slow time course and with high Wnt-3a doses. We also found that both PS-Dvl dependent and independent pathways are sensitive to Lrp5/6 inhibitor Dkk1 [22], a finding that is consistent with the idea that canonical Wnts (e.g. Wnt-1 or Wnt-3a) bind Lrp5/6 [27,28]. Moreover, Dvlindependent pathways have been previously found to be dependent on Lrp5/6 containing receptor complexes [5,29–31]. As we demonstrate for Wnt-3a, canonical Wnts have been found to activate β-catenin directly with high potency but low efficacy, by a PS-Dvl-independent mechanism that is Dkk1 sensitive and most likely requires Lrp5/6 and Axin [32,33]. On the other hand canonical Wnts act through Frizzled receptors in an Lrp5/6 independent manner [5, present data] to activate CK1δ/ε [15], which in turn results in PS-Dvl (present results). Our data show that high doses of Wnt-3a lead to the activation of CK1/Dvl, which then potentiates the Lrp6-dependent activation of β-catenin. Thus, a high level of canonical Wnt might be necessary to bring together Lrp5/6 and Frizzled [27], mediate their oligomerization at the membrane [31] and potentiate βcatenin signalling (present results). Indeed, phosphorylation of Dvl by CK1ε specifically enhances the interaction of Lrp5/6 with some binding proteins e.g. Frat-1 [34,35] that can then Fig. 5. Effects of Dkk1 on Wnt-3a-induced signalling. (A) SN4741 cells were treated with Wnt-3a (20 ng/ml) for 2 h in the presence or absence of 500 ng/ml of rmDkk1 (B) SN4741 cells were treated with control solution or with indicated combinations of D4476 (100 μM) and Wnt-3a (20 ng/ml) in presence of increasing doses of Dkk1 (100, 250, 500 and 1000 ng/ml). (C) SN4741 cells were pretreated with control solution, Dkk1 (500 ng/ml), D4476 (100 μM) or Dkk1+D4476, and then treated with Wnt-3a (50 ng/ml) or left untreated for 2 h. (A–C) After lysis the activation of β-catenin was determined by Western blotting using an antibody against active (Ser33/37-dephosphorylated) β-catenin (ABC). The phosphorylation of Dvl2 and/or Dvl3 was detected as phosphorylation-dependent mobility shift with the anti-Dvl2-specific antibody. The position of dephosphorylated (full triangle) and phosphorylated (PS-Dvl, open triangle) Dvl2 is indicated. (D) Quantification of active β-catenin signal from the experiment described in C (n=3). The signal was quantified using densitometry and the intensity of Wnt-3a only treated sample was normalized to 1. Statistical differences in Wnt-3a treated samples were tested by Student's t-test and the significant differences from control are indicated by asterisks (⁎pb0.05, ⁎⁎pb0.01). 615V. Bryja et al. / Cellular Signalling 19 (2007) 610–616 either directly or indirectly (e.g. via Axin) stabilize the Wntinduced Lrp5/6-Frizzled dimer intracellularly. In summary, our data are compatible with the idea that the reinforcement of βcatenin activation by high Wnt-3a doses could result from the oligomerization of Lrp5/6 and Frizzled and the recruitment of phosphorylated Dvl to the molecular machinery transducing the signal to β-catenin [31]. Our results also show that depending on the level of canonical Wnt, the cells may utilize PS-Dvl for signaling or not. The PS-Dvl independent, Dkk1 sensitive and Lrp5/6-mediated pathway is utilized mainly in the presence of low doses of canonical Wnt and also immediately after stimulation. Although molecular details of this pathway remain to be identified, the dynamics and other properties resemble recently reported G protein-dependent pathway leading to GSK3β inhibition in response to Wnt-3a [36]. In contrast, at higher doses of Wnt-3a and more than 30 min after stimulation, Dvl becomes concomitantly activated and potentiates β-catenin signalling. A high dose of canonical Wnt has been found to be required to mediate the recruitment of Lrp5/6 and Frizzled [31]. This type of mechanism is thought to result in the involvement of phosphorylated Dvl and the activation of β-catenin. One interesting question raised by this model is whether PS-Dvl-dependent or PS-Dvl-independent canonical signalling induced by high or low Wnt concentration, could differ in the activation of βcatenin not only quantitatively but also qualitatively, by regulating distinct functions. This possibility remains to be tested. In conclusion, our analysis of Wnt-3a action in SN4741 cells shows that Wnt-3a activates canonical β-catenin signalling by at least two mechanisms: by a CK1δ/ε and PS-Dvl-independent pathway that is rapid and can be activated by low doses of Wnt, and a PS-Dvl-dependent pathway that is slow and requires high doses of Wnt. Acknowledgments Financial support was obtained from the Swedish Foundation for Strategic Research, Swedish Royal Academy of Sciences, Knut and Alice Wallenberg Foundation, European Commission, Swedish MRC, Karolinska Institutet and Lars Hiertas Minnesfond. G.S. was supported by a post-doctoral fellowship from the Swedish Society for Medical Research (SSMF). We would like to thank Claudia Tello, Johny Söderlund and Annika Käller for technical and secretarial assistance and the members of Arenas lab for stimulating discussions. References [1] J. Huelsken, W. Birchmeier, Curr. Opin. Genet. Dev. 11 (2001) 547. [2] T.P. Yamaguchi, Curr. Biol. 11 (2001) R713. [3] A. Patapoutian, L.F. Reichardt, Curr. Opin. Neurobiol. 10 (2000) 392. [4] R. Nusse, (1 December 2004, revision date). [5] J.M. Gonzalez-Sancho, K.R. Brennan, L.A. Castelo-Soccio, A.M. Brown, Mol. Cell Biol. 24 (2004) 4757. [6] G. Castelo-Branco, J. Wagner, F.J. Rodriguez, J. Kele, K. Sousa, N. Rawal, H.A. Pasolli, E. Fuchs, J. Kitajewski, E. Arenas, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 12747. [7] G. Castelo-Branco, N. Rawal, E. Arenas, J. Cell Sci. 117 (2004) 5731. [8] J.H. Son, H.S. Chun, T.H. Joh, S. Cho, B. Conti, J.W. Lee, J. Neurosci. 19 (1999) 10. [9] G. Schulte, V. Bryja, N. Rawal, G. Castelo-Branco, K.M. Sousa, E. Arenas, J. Neurochem. 92 (2005) 1550. [10] N. Rawal, G. Castelo-Branco, K.M. Sousa, J. Kele, K. Kobayashi, H. Okano, E. Arenas, Exp. Cell Res. (2006). [11] M. van Noort, J. Meeldijk, R. van der Zee, O. Destree, H. Clevers, J. Biol. Chem. 277 (2002) 17901. [12] J.S. Lee, A. Ishimoto, S. Yanagawa, J. Biol. Chem. 274 (1999) 21464. [13] J.M. Peters, R.M. McKay, J.P. McKay, J.M. Graff, Nature 401 (1999) 345. [14] R.M. McKay, J.M. Peters, J.M. Graff, Dev. Biol. 235 (2001) 388. [15] W. Swiatek, I.C. Tsai, L. Klimowski, A. Pepler, J. Barnette, H.J. Yost, D.M. Virshup, J. Biol. Chem. 279 (2004) 13011. [16] J. Liu, A.G. Bang, C. Kintner, A.P. Orth, S.K. Chanda, S. Ding, P.G. Schultz, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 1927. [17] M. Kishida, S. Hino, T. Michiue, H. Yamamoto, S. Kishida, A. Fukui, M. Asashima, A. Kikuchi, J. Biol. Chem. 276 (2001) 33147. [18] G. Rena, J. Bain, M. Elliott, P. Cohen, EMBO Rep. 5 (2004) 60. [19] C. Liu, Y. Li, M. Semenov, C. Han, G.H. Baeg, Y. Tan, Z. Zhang, X. Lin, X. He, Cell 108 (2002) 837. [20] S. Amit, A. Hatzubai, Y. Birman, J.S. Andersen, E. Ben-Shushan, M. Mann, Y. Ben-Neriah, I. Alkalay, Genes Dev. 16 (2002) 1066. [21] H. Matsubayashi, S. Sese, J.S. Lee, T. Shirakawa, T. Iwatsubo, T. Tomita, S. Yanagawa, Mol. Cell Biol. 24 (2004) 2012. [22] A. Glinka, W. Wu, H. Delius, A.P. Monaghan, C. Blumenstock, C. Niehrs, Nature 391 (1998) 357. [23] B. Mao, W. Wu, Y. Li, D. Hoppe, P. Stannek, A. Glinka, C. Niehrs, Nature 411 (2001) 321. [24] A. Bafico, G. Liu, A. Yaniv, A. Gazit, S.A. Aaronson, Nat. Cell Biol. 3 (2001) 683. [25] C.Y. Logan, R. Nusse, Annu. Rev. Cell Dev. Biol. 20 (2004) 781. [26] J. Huelsken, J. Behrens, J. Cell Sci. 115 (2002) 3977. [27] K. Tamai, M. Semenov, Y. Kato, R. Spokony, C. Liu, Y. Katsuyama, F. Hess, J.P. Saint-Jeannet, X. He, Nature 407 (2000) 530. [28] G. Liu, A. Bafico, V.K. Harris, S.A. Aaronson, Mol. Cell Biol. 23 (2003) 5825. [29] K. Brennan, J.M. Gonzalez-Sancho, L.A. Castelo-Soccio, L.R. Howe, A.M. Brown, Oncogene 23 (2004) 4873. [30] L. Li, J. Mao, L. Sun, W. Liu, D. Wu, J. Biol. Chem. 277 (2002) 5977. [31] F. Cong, L. Schweizer, H. Varmus, Development 131 (2004) 5103. [32] K. Tamai, X. Zeng, C. Liu, X. Zhang, Y. Harada, Z. Chang, X. He, Mol. Cell 13 (2004) 149. [33] N.S. Tolwinski, M. Wehrli, A. Rives, N. Erdeniz, S. DiNardo, E. Wieschaus, Dev. Cell 4 (2003) 407. [34] E. Hay, C. Faucheu, I. Suc-Royer, R. Touitou, V. Stiot, B. Vayssiere, R. Baron, S. Roman-Roman, G. Rawadi, J. Biol. Chem. 280 (2005) 13616. [35] S. Hino, T. Michiue, M. Asashima, A. Kikuchi, J. Biol. Chem. 278 (2003) 14066. [36] X. Liu, J.S. Rubin, A.R. Kimmel, Curr. Biol. 15 (2005) 1989. 616 V. Bryja et al. / Cellular Signalling 19 (2007) 610–616       Vítězslav Bryja, 2014    Attachments      #3      Bryja  V1 ,  Schulte  G1 ,  Rawal  N,  Grahn  A,  Arenas  E  (2007):  Wnt‐5a  induces  Dishevelled  phosphorylation  and  dopaminergic  differentiation  via  a  CK1‐dependent  mechanism.  J.  Cell  Sci. 120: 586‐595.   1  equal contribution      Impact factor (2007): 6.247  Times cited (without autocitations, WoS, Feb 21st 2014): 40  Significance:  This  work  for  the  first  time  identified  a  kinase  –  casein  kinase  1  delta/epsilon – which is responsible for phosphorylation of Dishevelled in the non‐ canonical Wnt pathway. Till today, CK1‐controlled Dvl shift represent one of the  most  commonly  used  readouts  of  the  non‐canonical  Wnt  pathway.  Nowadays  commonly  used  term  –  PS‐Dvl  (phosphorylated  and  shifted  Dvl)  has  been  introduced in this study.  Contibution of the author/author´s team:  Together with Gunnar Schulte identification  of CK1 as the kinase acting downstream of Wnt5a. Biochemical and functional  validation of the significance.                586 Research Article Introduction The Wnt signalling pathway is a highly conserved biochemical pathway that is involved in a vast array of processes in both embryonic development and adult tissue homeostasis (for reviews, see Huelsken and Birchmeier, 2001; Patapoutian and Reichardt, 2000; Yamaguchi, 2001). Moreover, key molecular players of the Wnt pathway have been found to regulate midbrain development (McMahon and Bradley, 1990; Pinson et al., 2000; Thomas and Capecchi, 1990; Wang et al., 2002) and various aspects of dopaminergic neuron (DN) development (McMahon and Bradley, 1990; Pinson et al., 2000; Thomas and Capecchi, 1990; Wang et al., 2002; Arenas, 2005; CasteloBranco et al., 2003; Prakash et al., 2006). We have shown that Wnt-5a – a non-canonical Wnt, classified by its inability to activate ␤-catenin (Shimizu et al., 1997) – plays a pivotal role in the ventral midbrain DN differentiation (Castelo-Branco et al., 2006; Castelo-Branco et al., 2003). To date, the underlying molecular mechanism of action of Wnt-5a and the signalling pathways activated in DN as well as in other mammalian cells is still largely unknown (Veeman et al., 2003). Several molecular players have been implicated, including the Wnt receptors of the Frizzled family and the downstream signalling phosphoprotein Dishevelled (Dvl) (Gonzalez-Sancho et al., 2004; Hsieh, 2004; Wallingford and Habas, 2005; Wharton, Jr, 2003). Here, we examined the mechanism through which Wnt-5a activates Dvl in dopaminergic cells. We report the identification of casein kinase 1 (CK1) ␦ and CK1⑀ (hereafter referred to as CK1␦/⑀) as kinases phosphorylating Dvl2 and Dvl3 in response to Wnt-5a. We show that Wnt-5a-induced CK1⑀-mediated phosphorylation of Dvl2 results in changes of the cytoplasmic distribution of Dvl2. Finally, we report that activity of endogeneous CK1 is crucial for the prodifferentiation function of Wnt-5a in dopaminergic precursors. Thus, we hereby identify CK1 as a positive regulator of DN development. Results Wnt-5a phosphorylates Dvl in a dopaminergic neuronal cell line To characterise the pathways activated by Wnt-5a in dopaminergic cells, we treated SN4741 cells with Wnt-5a and used Wnt-3a (a canonical Wnt) for comparison. Treatment with either Wnt form (at 100 ng/ml) lead to the phosphorylation of Dvl2 and Dvl3, as shown previously by a mobility shift of the protein on SDS-PAGE (Gonzalez-Sancho et al., 2004; Lee et al., 1999; Schulte et al., 2005; Bryja et al., 2007). Both Dvl2 and Dvl3 showed the first visible signs of phosphorylation at 30 minutes of treatment and a clear phosphorylation shift after 1 hour (Fig. 1A). Maximal effects were detected after 2 hours and, hence, this time point was used subsequently unless otherwise specified. Whereas both Wnt-3a and Wnt-5a induced Dvl phosphorylation (Fig. 1A), only Wnt-3a induced ␤-catenin activation (Fig. 1B), as assessed by an antibody recognizing Previously, we have shown that Wnt-5a strongly regulates dopaminergic neuron differentiation by inducing phosphorylation of Dishevelled (Dvl). Here, we identify additional components of the Wnt-5a-Dvl pathway in dopaminergic cells. Using in vitro gain-of-function and loss-of-function approaches, we reveal that casein kinase 1 (CK1) ␦ and CK1⑀ are crucial for Dvl phosphorylation by non-canonical Wnts. We show that in response to Wnt-5a, CK1⑀ binds Dvl and is subsequently phosphorylated. Moreover, in response to Wnt-5a or CK1⑀, the distribution of Dvl changed from punctate to an even appearance within the cytoplasm. The opposite effect was induced by a CK1⑀ kinase-dead mutant or by CK1 inhibitors. As expected, Wnt-5a blocked the Wnt-3a-induced activation of ␤catenin. However, both Wnt-3a and Wnt-5a activated Dvl2 by a CK1-dependent mechanism in a cooperative manner. Finally, we show that CK1 kinase activity is necessary for Wnt-5a-induced differentiation of primary dopaminergic precursors. Thus, our data identify CK1 as a component of Wnt-5a-induced signalling machinery that regulates dopaminergic differentiation, and suggest that CK1␦/⑀mediated phosphorylation of Dvl is a common step in both canonical and non-canonical Wnt signalling. Key words: Casein kinase 1␦/⑀, Dishevelled, Wnt-5a, Dopaminergic neurons, Non-canonical Wnt signalling, siRNA Summary Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism Vítezslav Bryja*,‡ , Gunnar Schulte*,§ , Nina Rawal¶ , Alexandra Grahn¶ and Ernest Arenas§ Laboratory of Molecular Neurobiology, Department of Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden *,¶ These authors contributed equally to this work ‡ Present address: Laboratory of Cytokinetics, Institute of Biophysics AS CR, 61265, Brno, Czech Republic § Present address: Section of Receptor Biology and Signaling, Department of Physiology and Pharmacology, Karolinska Institutet, S-171 77 Stockholm, Sweden § Author for correspondence (e-mail: ernest.arenas@ki.se) Accepted 1 December 2006 Journal of Cell Science 120, 586-595 Published by The Company of Biologists 2007 doi:10.1242/jcs.03368 JournalofCellScience 587CK1 mediates Wnt-5a action active ␤-catenin (ABC, the form of ␤-catenin dephosphorylated on Ser37 and Thr41) (van Noort et al., 2002). To confirm that the observed signalling was specific to Wnts, we treated SN4741 cells with the broad-spectrum inhibitor of Wnt signalling, soluble Fz8-CRD (the cysteinerich domain of Frizzled 8 that competes with Frizzled receptors for Wnt binding) (Hsieh et al., 1999). Pre-treatment with Fz8CRD blocked both basal and Wnt-induced Dvl phosphorylation and ␤-catenin activation (Fig. 1C), indicating that the effects observed were specifically induced by Wnts. CK1␦/⑀ mediate Wnt-5a-induced phosphorylation of Dvl A large number of kinases have been implicated in Wnt signalling; however, it remains unclear which kinase(s) are responsible for Dvl phosphorylation, leading to its electrophoretic mobility shift after Wnt-5a stimulation. To identify the relevant signalling pathways leading to the phosphorylation-dependent mobility shift of Dvl, we analysed a panel of small molecule compounds for their ability to block or reduce the Wnt-5a-induced mobility shift of Dvl2 that takes place after phosphorylation (Gonzalez-Sancho et al., 2004). Twenty five pharmacological compounds interfering with heterotrimeric G proteins, protein kinase C, protein kinase A, MEK1/2, PI3K, p38, JNK, CamKII, GSK3, cAMP signalling, EPAC, adenylyl cyclase, Src-like kinases, EGFR, phospholipase C, CK1 or Ser/Thr kinases were tested (see Table 1). Only D4476, an inhibitor of CK1 (Rena et al., 2004), was able to block the Wnt-5a-induced phosphorylationdependent mobility shift of Dvl2. Interestingly, D4476 reduced both basal and also Wnt-5a-induced phosphorylation of Dvl2 and Dvl3 (Fig. 2A). It should be noticed that our Fig. 1. Wnt-5a and Wnt-3a activate Dvl in dopaminergic cells. (A) Time course of Wnt-5a- and Wnt-3a-induced activation of Dvl2 and Dvl3. SN4741 cells, treated with Wnt-3a (100 ng/ml) or Wnt-5a (100 ng/ml), were lysed after 0, 1, 5, 10, 15, 30, 60 and 120 minutes. (B) Both Wnt-5a and Wnt-3a activated Dvl2 and Dvl3, but only Wnt- 3a activated ␤-catenin after 2 hours of stimulation. (C) The effects of Wnt-5a or Wnt-3a (100 ng/ml) on Dvl were blocked by Fz8-CRDconditioned medium. (A-C) Phosphorylation of Dvl isoforms were detected as phosphorylation-dependent mobility shifts of total Dvl2 and Dvl3. The position of dephosphorylated (᭣) and phosphorylated Dvl (᭠) is indicated. The activation of ␤-catenin was determined by western blotting using antibodies against active (Ser33-Thr41dephosphorylated) ␤-catenin (ABC). Data are representative of at least three independent replicates. Table 1. Screening of compounds for the ability to interfere with the Wnt-induced shift of Dvl Compound Target Concn Activity PTX Gi/o 100 ng/ml No PDBu PKC activator 1 ␮M No Wortmannin PI3K 50 nM No LY294002 PI3K 50 ␮M No PD98059 MEK1/2 10 ␮M No UO126 MEK1/2 10 ␮M No SB203580 p38 10 ␮M No JNKII inhib JNK 6 ␮M No Genistein PKC 50 ␮M No Chelerythrine PKC 10 ␮M No Ro-31 8220 PKC 1 ␮M No BIM I PKC 500 nM No KN93 CaMKII 10 ␮M No I3M GSK-3 2 ␮M No Kenpaullone GSK-3 6 ␮M No H89 PKA 10 ␮M No 8-Br-cAMP cAMP pathway activator 10 ␮M No 8CPT-2Me-cAMP EPAC activator 30 ␮M No SQ22536 Adenylyl cyclase 100 ␮M No MDL12330 Adenylyl cyclase 10 ␮M No PP2 Src-like 10 ␮M No AG1276 EGFR 10 ␮M No ET-18-OCH3 PLC 10 ␮M No D4476 Casein kinase 1 100 ␮M Yes Staurosporine Ser/Thr kinases, PKC 2 ␮M No Gi/o, heterometric G protein (inhibitory); PTX, pertussis toxin; PDBu, phorboldibutyrate; PKC, Ca2+ -dependent protein kinase; PI3K, phosphatidylinositol-3Ј-kinase; LY294002, 2-(4-morpholino)-8-phenyl-4H-1benzopyran-4-one; PD98059, 2Ј-amino-3Ј-methoxyflavone; MEK1/2, MAPK and ERK kinase1/2; UO126, 1,4-diamino-2,3-dicyano-1,4-bis(2aminophenylthio)-butadiene; SB203580, 4-[5-(4-fluorophenyl)-2-[4(methylsulfonyl)phenyl]-1H-imidazol-4-yl]pyridine; p38, stress-activated protein kinase p38; JNK, c-jun N-terminal kinase; Ro-318220, bisindolylmaleimide IX; BIM I, bisindolmaleimide I; KN93, 2-[N-(2- hydroxyethyl)]-N-(4-methoxybenzenesulfonyl)]amino-N-(4-chlorocinnamyl)N-methylbenzylamine); CaMKII, Ca2+ /calmodulin-dependent kinase II; I3M, indirubin-3-monoxime; GSK-3, glycogen synthase kinase 3; H89, N-[2-(pbromocinnamylamino)ethyl]-5-isoquinolinesulfonamide; PKA, cAMPdependent protein kinase; 8-Br-cAMP, 8-bromo-cyclic AMP; 8CPT-2MecAMP, 8-(4-chloro-phenylthio)-2Ј-O-methyladenosine-3Ј,5Ј-cyclic monophosphate; SQ22536, 9-(tetrahydro-2-furanyl)-9H-purin-6-amine; MDL12330, cis-N-(2-phenylcyclopentyl)azacyclotridec-1-en-2-amine; PP2, 4-Amino-5-(4-chlorophenyl)-7-(t-butyl)pyrazolo[3,4-d]pyrimidine; AG1478, 4-(3-chloroanillino)-6,7-dimethoxyquinazoline; EGFR, epidermal growth factor receptor; ET-18-OCH3, rac-2-methyl-1-octadecyl-glycero-(3)phosphocholine; PLC, phospholipase C; D4476, 4-(4-(2,3- dihydrobenzo[1,4]dioxin-6-yl)-5-pyridin-2-yl-1H-imidazol-2-yl)benzamide. JournalofCellScience 588 results do not exclude the possibility that other phosphorylation events exist that were not detected in the mobility shift assay. Journal of Cell Science 120 (4) CK1 has previously been shown to be involved in various steps of Wnt signal transduction (Davidson et al., 2005; Kishida et al., 2001; Price, 2006; Zeng et al., 2005). Importantly, CK1␣ was shown to phosphorylate ␤-catenin at Ser45, priming ␤-catenin for subsequent phosphorylation by GSK3␤ (Fig. 2B) (Amit et al., 2002; Liu et al., 2002; Matsubayashi et al., 2004). By contrast, CK1␦ and CK1⑀ lack the ability to phosphorylate ␤-catenin (Liu et al., 2002; Peters et al., 1999), but are known to bind and phosphorylate Dvl in the canonical Wnt signalling pathway (Cong et al., 2004; Kishida et al., 2001; McKay et al., 2001a; Peters et al., 1999; Swiatek et al., 2004). To test which CK1 subtype is inhibited by D4476, we analysed the effect of D4476 on the level of CK1␣-mediated phosphorylation of the Ser45 residue of ␤catenin. As shown by phosphorylation-specific antibodies, the phosphorylation of ␤-catenin at Ser45 was not affected by Wnt-5a but was significantly decreased following the application of D4476 (Fig. 2A), suggesting that CK1␣ can be inhibited by D4476. To elucidate which CK1 isoforms are responsible for Dvl2 phosphorylation, we used the more specific CK1 inhibitor IC261. IC261 has previously been reported to efficiently inhibit CK1␦/⑀ at low micromolar doses (in vitro IC50=1 ␮M) but not CK1␣ (in vitro IC50=16 ␮M) (Mashhoon et al., 2000). Despite being less efficient than D4476, IC261 (10 ␮M) inhibited Wnt-5a-induced Dvl2 phosphorylation (Fig. 2C). However, in contrast to D4476, IC261 did not reduce the levels of ␤-catenin phosphorylated at Ser45, indicating that CK1␦/⑀, but not CK1␣, kinase activity was inhibited. The levels of active ␤-catenin, as well as total ␤-catenin levels, were unchanged by Wnt-5a, D4476 or IC261 treatment (Fig. 2A). These experiments suggested that CK1␦/⑀, rather than CK1␣, phosphorylate Dvl2 and Dvl3 in response to Wnt-5a. To confirm that CK1⑀ phosphorylates Dvl2, gain-of- and loss-of-function experiments were performed in SN4741 cells transiently transfected with plasmids encoding Dvl2-Myc (Lee et al., 1999), CK1⑀ or the CK1⑀ (K>R) mutant (a kinase-dead form of CK1⑀) (Fish et al., 1995). We found that CK1⑀, but not CK1⑀ (K>R), phosphorylated Dvl2-Myc (Fig. 2D). Similar data were obtained with Dvl2-GFP (not shown), demonstrating that the kinase activity of CK1⑀ is required for Dvl2 phosphorylation in the overexpression system. To analyse the role of endogenous CK1 in the Wnt-5ainduced phosphorylation of Dvl2, we performed gene knockdown of CK1␣, CK1␦ and CK1⑀ using small interfering RNAs (siRNA). Three independent siRNAs, each designed against the various CK1 isoforms, were tested for their efficiency in silencing endogenous CK1 in SN4741 cells. Efficiency of gene knockdown was analysed by western blotting for CK1⑀ (Fig. 3A) and by quantitative reverse transcriptase (RT)-PCR for CK1␣ and CK1␦ (not shown), where subtype-specific antibodies failed to detect endogenous CK1␣ and CK1␦. At least one siRNA against each CK1 isoform provided a strong gene knockdown of more than 50%, as assessed by western blotting (CK1⑀, Fig. 3A) and quantitative RT-PCR for CK1␦ and CK1␣ (data not shown). The most efficient siRNAs – CK1␣III, CK1␦III and/or CK1⑀II were then transfected into SN4741 that were stimulated with increasing doses of Wnt-5a. The compound knockdown of CK1␦ and CK1⑀ resulted in a significant reduction of the Wnt- 5a-induced phosphorylation of Dvl2, whereas the CK1␣ Fig. 2. CK1 inhibition blocks Wnt-5a-induced phosphorylation of Dvl2 and Dvl3. SN4741 cells were treated with vehicle (control) or Wnt-5a (100 ng/ml) for 2 hours in the presence or absence of D4476 (100 ␮M) (A) or IC261 (10 ␮M) (C). Western blot analysis was performed as in Fig. 1. The position of dephosphorylated (᭣) and phosphorylated (᭠) Dvl is indicated. Antibodies against phosphoSer45-␤-catenin (CK1␣ target site), against active (Ser37-Thr41dephosphorylated) ␤-catenin (ABC) and total ␤-catenin were used. (B) Scheme of phosphorylation sites of ␤-catenin. Ser45 is a CK1␣ target site, which serves as a priming for sequential phosphorylation of Thr41, Ser37 and Ser33 by GSK3␤. (D) SN4741 cells were transfected with plasmids encoding Dvl2-Myc and either CK1⑀ or kinase-dead CK1⑀ (K>R) mutant. Phosphorylation of Dvl2-Myc was detected as a phosphorylation-dependent mobility shift by Myc- and Dvl2-specific antibodies. CK1⑀ levels were monitored with a CK1⑀ specific antibody. Data in A, C and D are representative of at least three independent replicates. JournalofCellScience 589CK1 mediates Wnt-5a action siRNA had no effect compared with control (Fig. 3B). Although one cannot exclude the possibility that lack of the effect of CK1␣ siRNA was due to incomplete knockdown, our data suggest that CK1␦/⑀, rather than CK1␣, are responsible for Wnt-5a-mediated phosphorylation of Dvl2. Interestingly, knockdown of CK1⑀ alone, or together with less efficient CK1 (␦I and ␦II) siRNAs, was not sufficient to reduce the effects of Wnt-5a on Dvl2 (not shown), suggesting that CK1␦ and CK1⑀ are to some extent redundant in their function. To confirm that CK1␦III does not act by an off-target mechanism but rather that a joint knockdown of CK1␦ and CK1⑀ is necessary, we treated SN4741 cells with control RNA, CK1␦III or CK1⑀II, or the combination of the siRNAs. As we show in Fig. 3C, CK1␦III or CK1⑀II alone were not able to block Wnt-5a-mediated phosphorylation of Dvl, whereas their combination blocked phosphorylation of endogenous Dvl2 very efficiently. Taken together, these experiments demonstrate that CK1␦ and CK1⑀, rather than CK1␣, phosphorylate Dvl2 in response to Wnt-5a. Wnt-5a induces the activation of endogenous CK1⑀ kinase that interacts with Dvl2 To examine whether Wnt-5a induces the kinase activity of endogenous CK1⑀, we tested and confirmed that CK1⑀ can be immunoprecipitated in SN4741 cells (Fig. 4A). Using myelin basic protein (MBP, a general substrate of Ser/Thr kinases) as a substrate in an in vitro CK1⑀ kinase assay, we found that treatment of SN4741 cells with either Wnt-5a or Wnt-3a clearly upregulates CK1⑀ activity (Fig. 4B), which is an effect previously shown only for canonical Wnts (Swiatek et al., 2004). This demonstrates that Wnt-5a, as well as Wnt-3a, activates endogenous CK1⑀ in dopaminergic SN4741 cells. Additionally, we examined whether Dvl2 and CK1⑀ physically interact in SN4741 cells, as it has been shown previously in other cellular models (Peters et al., 1999; Sakanaka et al., 1999), and found that Dvl2-Myc binds both overexpressed CK1⑀ (Fig. 4C, lane 2) and endogenous CK1⑀ (Fig. 4D). Please notice that endogenous CK1⑀ was only clearly detected when beads coupled to antibody where used to enhance the signal above background. Importantly, the lack of kinase activity in CK1⑀ (K>R) does not prevent the interaction with Dvl2 (Fig. 4C, lane 3). These results suggest that the kinase activity of CK1⑀ is not needed for binding to Dvl2, but it is required for the phosphorylation of Dvl2. CK1⑀ and Wnt-5a change the subcellular localization of Dvl2 The subcellular distribution of Dvl was examined after transfection of SN4741 cells with a Dvl2-Myc plasmid. Dvl2Myc was detected by immunocytochemistry in the cytoplasm and found either in a diffuse and even pattern, or in punctae that represent large multiprotein complexes (Schwarz-Romond et al., 2005; Smalley et al., 2005). Based on the presence of Dvl2-Myc punctae and their size, cells were sorted into four different categories (Fig. 5A): (1) even distribution (no punctae detected), (2) punctae of small size (small dot-like punctae on the background of still detectable even cytoplasmic staining), (3) punctae of intermediate size (distinct punctae strongly contrasting with the negative staining of the remaining cytoplasm) or, (4) punctae of large size (individual punctae already fused that form large doughnut-like structures). Interestingly, the distribution of Dvl2-Myc was dramatically regulated by transfection of the CK1⑀ constructs (Fig. 5B). Both CK1⑀ and CK1⑀ (K>R) showed a homogeneous cytoplasmatic distribution when transfected alone and stained with an anti-CK1⑀-specific antibody (not shown). Upon coexpression of CK1⑀ with Dvl2-Myc, the distribution of Dvl2-Myc changed from punctate 70% in the control to an even distribution in 65% of the cells (Fig. 5B,C); a finding similar to what has previously been shown in HEK293 cells (Cong et al., 2004). Conversely, CK1⑀ (K>R) promoted the punctate localization of Dvl2-Myc in 95% of the cells by predominantly increasing intermediate puncta (Fig. 5B,C). In all cases, we found that CK1⑀ (wt or K>R) colocalises with Dvl2-Myc, suggesting that CK1⑀ controls the distribution of Dvl. Fig. 3. Gene knockdown of CK1␦/⑀ reduces the phosphorylation of Dvl2 induced by Wnt-5a. (A) SN4741 cells were transfected with control siRNA and three independent siRNAs against CK1⑀ isoform. The efficiency of gene knockdown induced by individual siRNAs against CK1⑀ was quantified in western blots. A non-specific protein, resulting in a band recognised by anti-CK1⑀ antibody served as a loading control. Quantification (normalised to loading control) is shown below. (B,C) SN4741 cells were transfected with the indicated combinations of siRNAs, and treated with control, 50 and 100 ng/ml of Wnt-5a. The phosphorylation of Dvl isoforms was detected as phosphorylation-dependent mobility shift of total Dvl2. Data are representative of three (B) or two (C) independent experiments. The position of dephosphorylated (᭣) and phosphorylated Dvl (᭠) is indicated. The levels of CK1⑀ and ␤-actin were also determined to confirm the efficient gene knockout and the equal protein loading. JournalofCellScience 590 Journal of Cell Science 120 (4) When the CK1 inhibitor D4476 was added after transfection, it significantly reduced CK1⑀-mediated phosphorylation of Dvl2-Myc (Fig. 5D, lanes 6, 8 and 10) confirming that this CK1 inhibitor specifically blocks the action of CK1⑀ on Dvl2 in a dose-dependent manner. Interestingly, when the cells were transfected with Dvl2-Myc alone and treated with the CK1 inhibitor D4476, Dvl2-Myc phosphorylation was prevented (Fig. 5D, lane 4) and the localization of Dvl2-Myc changed from an even distribution in 50% of the cells to a punctate pattern in 85% of the cells (Fig. 5E). Similar results were obtained using the CK1␦/⑀specific inhibitor IC261 (Fig. 5F). To directly test whether activation of endogenous CK1␦/⑀ by Wnt-5a resulted in a relocalization of Dvl2-Myc similar to the one induced by overexpressed CK1, we treated Dvl2-Myc-overexpressing SN4741 cells with Wnt-5a. Wnt-5a treatment resulted in a statistically significant increase in the number of cells with even distribution of Dvl2-Myc at the expense of the cells with Dvl2-Myc in punctae (Fig. 5G). When using 100 ng/ml of Wnt-3a and an identical experimental setup as for Wnt-5a, we failed to detect similar changes in cytoplasmic distribution of Dvl2-Myc induced by Wnt-3a (not shown), suggesting that Wnt-induced relocalization of Dvl is an effect specific for non-canonical Wnts. In summary, these results demonstrate that Wnt-5a has an effect similar to that of CK1⑀ and suggest that endogenous CK1␦/⑀ regulate the phosphorylation and cellular localization of Dvl2 in response to Wnt-5a. Combined, our results indicate that active CK1⑀, either overexpressed or endogenous (activated by Wnt-5a), phosphorylates Dvl2 and induces a diffuse cytoplasmic distribution of phosphorylated Dvl2 that can be blocked by either CK1 inhibition or the kinase-dead CK1⑀ (K>R). Wnt-5a cooperates with Wnt-3a in the phosphorylation of Dvl, but antagonises Wnt-3a in the activation of ␤- catenin Given that D4476 is a reversible competitive inhibitor of the ATP binding site in CK1, we examined whether the inhibition of Dvl2 phosphorylation by D4476 is modulated by increasing doses of Wnt-5a. Increased amounts of Wnt-5a dose dependently overcame the D4476-mediated inhibition and lead to a phosphorylation-dependent mobility shift of Dvl2 (Fig. 6A). We next explored whether Wnt-5a and Wnt-3a phosphorylates Dvl by similar mechanisms and, if so, whether their effects are additive. SN4741 cells were pre-treated with Wnt-5a (100 ng/ml) or Wnt-3a (20 ng/ml), the lowest doses leading to the efficient phosphorylation of Dvl2 and Dvl3 (data not shown); then, increasing doses of Wnt-3a (50, 100 and 200 ng/ml) or Wnt-5a (100, 200 and 500 ng/ml) were applied. The results showed that both Wnt-3a and Wnt-5a activate Dvl phosphorylation (Fig. 5B and C, respectively). Moreover, Wnt- 3a and Wnt-5a cooperated in the phosphorylation of Dvl in the absence of D4476 (as monitored by the disappearance of the non-shifted band of Dvl2). When the additive effects of Wnts were tested in the presence of D4476, Wnt-3a rescued the D4476-mediated block of Wnt-5a-induced phosphorylation of Dvl2 and vice versa (Fig. 6C,D). Not surprisingly, active ␤catenin was induced when Wnt-3a was added to Wnt-5a pretreated cells (Fig. 6B,D). By contrast, when Wnt-5a was added to cells pre-treated with Wnt-3a, it significantly and dose dependently reduced the activation of ␤-catenin, irrespective of Fig. 4. CK1⑀ is activated by Wnt-5a and binds Dvl2 in dopaminergic cells. (A) SN4741 cell lysates were immunoprecipitated with a control IgG or a CK1⑀-specific antibody and endogenous or overexpressed CK1⑀ and detected by western blotting. (B) Wnt-3a and Wnt-5a increased the phosphorylation of MBP as detected by western blotting using a specific antibody against phosphorylated serine. Cell lysates from control (1% CHAPS), Wnt-3a or Wnt-5a treated (100 ng/ml) SN4741 cells were immunoprecipitated with a CK1⑀-specific antibody and subjected to kinase assay using MBP as a substrate. (C) Dvl2-Myc interacts with either CK1⑀ (lane 2) or a kinase-dead CK1⑀ (K>R) (lane 3) mutant in SN4741 cells, as assessed by immunoprecipitation and western blotting with Myc- and CK1⑀-specific antibodies. Co-precipiated CK1⑀ by the Myc antibody is indicated by ᭣. For all other panels: ᭠, phosphorylated Dvl2-Myc; ᭣, unphosphorylated Dvl2-Myc. (D) Dvl2-Myc interacts with endogenous CK1⑀ as assessed by immunoprecipitation of SN4741 cells transfected with Dvl2-Myc only, by using Myc antibody-conjugated agarose beads (control, Gprotein-conjugated beads). Data are representative of three independent replicates. JournalofCellScience 591CK1 mediates Wnt-5a action the presence of D4476 (Fig. 6D,E). Combined, these data suggests that Wnt-3a and Wnt-5a phosphorylate Dvl in SN4741 cells by a similar or even identical mechanism involving the activation of CK1␦/⑀. Moreover, we show that, although Wnt-3a and Wnt-5a cooperate in the phosphorylation of Dvl, Wnt-5a can antagonise Wnt-3a-mediated induction of ␤-catenin. These results suggest that phosphorylated Dvl serves different functions when recruited to pathways activated by Wnt-3a or Wnt-5a. CK1 inhibitors block the biological effects of Wnt-5a on dopaminergic precursors To determine whether CK1 is also part of the signalling machinery mediating the pro-differentiation activity of Wnt- 5a on dopaminergic precursors (Castelo-Branco et al., 2003; Schulte et al., 2005), we analysed the consequences of CK1 inhibition in rat embryonic day14.5 (E14.5) primary ventral midbrain precursor cultures. Cells were treated with Wnt-5a (100 ng/ml) with or without increasing concentrations of D4476. D4476 had no effect on the total cell number (not shown) and, in agreement with our previous results (Schulte et al., 2005), the number of tyrosine-hydroxylase-positive (TH+ ) cells per field increased after Wnt-5a treatment. Importantly, we found that this effect was reduced in a dosedependent manner upon addition of D4476 (Fig. 7). Thus, our results suggest that CK1 activity is necessary for the biological effects of Wnt-5a on primary dopaminergic cells. Discussion We have previously shown that Wnt-5a induces the differentiation of dopaminergic progenitors (Castelo-Branco et al., 2003; Schulte et al., 2005). This present study identifies another component mediating the function of Wnt-5a in DA neurons (DN) development. After testing a panel of smallmolecule drugs and performing loss-of-function (CK1␦/⑀specific inhibitors and siRNA) as well as gain-of-function experiments, we identified CK1␦/⑀ as the relevant kinases hyperphosphorylating Dvl in response to Wnt-5a. Our findings place CK1␦/⑀ in a signalling pathway activated by a noncanonical Wnt and show for the first time that CK1 activity is Fig. 5. CK1⑀ and Wnt-5a mediate the changes in phosphorylation and cytoplasmic distribution of Dvl2. (A) The subcellular distribution of Dvl2-Myc in transfected SN4741 cells was assessed through an antiMyc antibody by confocal microscopy. Four patterns were detected: (1) even distribution, and punctae of (2) small, (3) intermediate or, (4), large size. For a detailed description see the results section. (B) Co-transfection of Dvl2-Myc and CK1⑀ or the kinase-dead CK1⑀ (K>R) mutant resulted in a predominant even or punctate distribution of Dvl2-Myc, respectively. The arrow points to a cell transfected with Dvl2-Myc (Cy3, red) with no or low levels of exogenous CK1⑀ (Cy2, green), serving as an internal control of the experiment. (C) Quantification of changes in the intracellular localization of Dvl2-Myc upon coexpression of CK1⑀ or CK1⑀ (K>R) as described in A and B. (D) D4476 (100 ␮M) blocks the phosphorylation dependent shift of Dvl2-Myc induced by different amounts (ng/well) of transfected CK1⑀ in SN4741 cells. The position of dephosphorylated (᭣) and phosphorylated (᭠) Dvl2 is indicated. Data are representative of at least three independent replicates. (E-G) SN4741 cells were transfected with Dvl2-Myc and treated as indicated: (E) The general CK1 inhibitor D4476 (100 ␮M) and (F) the CK1␦/⑀-specific inhibitor IC261 (20 ␮M) changed the subcellular localization of Dvl2-Myc to a predominant punctate distribution. This is in contrast with Wnt-5a treatment (100 ng/ml) (G), which leads to a more even distribution of Dvl2-Myc than in control. Data in C,E,F,G (nу3) were statistically evaluated, Data represent the mean ± s.d., one-way ANOVA with Bonferroni post-tests (*P<0.05, **P<0.01, ***P<0.001). JournalofCellScience 592 required for a biological process induced by a non-canonical Wnt. CK1␦/⑀ have previously been reported to be a DvlJournal of Cell Science 120 (4) phosphorylating kinase acting in the ␤-catenin pathway (Gao et al., 2002; Kishida et al., 2001; McKay et al., 2001a; Peters et al., 1999; Swiatek et al., 2004). A recent report (Cong et al., 2004) describes that overexpression of CK1⑀ potentiates canonical Wnt signalling and diminishes JNK activation induced by Dvl1 overexpression. This finding led to the suggestion that CK1⑀ modulates the signalling specificity of Dvl towards ␤-catenin (Cong et al., 2004). Here, we report that CK1␦/⑀ also mediate non-canonical signalling, suggesting that canonical or non-canonical specificities are not determined by CK1⑀ but rather by the ligand. Moreover, the findings reported by Cong et al. could be alternatively explained by the fact that CK1⑀-mediated phosphorylation diminishes the activity of axin in the MEKK1-JNK pathway at the expense of its function in canonical Wnt signalling (Zhang et al., 2002). Although the involvement of CK1␦/⑀ in the signal transduction of a noncanonical Wnt has not been demonstrated to date, a role of CK1␦/⑀ in other biological processes driven by non-canonical Wnts has been described. These include convergent extension movements in Xenopus (McKay et al., 2001b) and functions regulated by planar cell polarity pathway in Drosophila (Klein et al., 2006; Strutt et al., 2006). Thus, our data, together with published reports, support the notion that CK1⑀ mediates noncanonical Wnt signalling. Interestingly, a recent report by Fig. 6. Wnt-5a cooperates with Wnt-3a in the phosphorylation of Dvl2, but inhibits Wnt-3a-induced activation of ␤-catenin. (A) Increasing doses of Wnt-5a (50, 100, 300, 500, 1000 ng/ml) increase Dvl2 phosphorylation in vehicle (DMSO)-treated SN4741 cells and compete in the blockade of CK1 by D4476. (B-E) SN4741 cells were pretreated with 100 ng/ml of Wnt-5a (B,D) or 20 ng/ml of Wnt-3a (C,E) in the presence (D,E) or absence (B,C) of D4476 (100 ␮M) for 5 minutes. Subsequently, Wnt-3a (50, 100 and 200 ng/ml) or Wnt-5a (100, 200 and 500 ng/ml), was added for 2 hours. In A-E the phosphorylation of Dvl2 was detected by western blotting as phosphorylation-dependent mobility shift (dephosphorylated, ᭣ ; phosphorylated, ᭠). In B-E the activation of ␤-catenin (ABC) was determined using an antibody against Ser33/37-dephosphorylated ␤catenin. Results shown in B-E were quantified using densitometry and either normalised to untreated control for ABC or shown as a ratio of phosphorylated:dephosphorylated for Dvl2. Duplicate experiments showed comparable results. Fig. 7. CK1 inhibitors block the effects of Wnt-5a on dopaminergic differentiation. (A) Wnt-5a (100 ng/ml) increased the number of tyrosine-hydroxylase-positive (TH+ ) neurons in rat E14.5 ventral midbrain precursor cultures. However, addition of increasing doses of the chemical inhibitor of CK1 (D4476) reduced the numbers of TH+ neurons. (B) Double TH-Hoechst33258 immunostaining shows an increase in the number of TH immunoreactive neurons after treatment with Wnt-5a (100 ng/ml). Addition of 25 ␮M D4476 to Wnt-5a-treated cultures decreases the number of TH+ neurons after 3 days. Bar, 25 ␮m. JournalofCellScience 593CK1 mediates Wnt-5a action Takada et al. suggests that, in Drosophila cells another CK1 isoform, CK1␣ is playing a similar role to the one described here for CK1␦/⑀ in Wnt-5a-driven phosphorylation of Dvl (Takada et al., 2005). Thus, it remains to be investigated whether the involvement of individual CK1 isoforms in Dvl phosphorylation differs among species. Our results clearly show that both canonical Wnt-3a and non-canonical Wnt-5a induce the phosphorylation of Dvl by a common mechanism, involving the activation of CK1␦/⑀. This conclusion is based on the following lines of evidence: (1) the position of hyperphosphorylated Dvl bands in Wnt-3a- and Wnt-5a-treated samples is indistinguishable; (2) the time course of Dvl phosphorylation is identical when induced by either Wnt-3a or Wnt-5a; (3) the phosphorylation of Dvl by both Wnt-5a (this study) and Wnt-3a (Bryja et al., 2007) can be blocked by CK1␦/⑀ siRNAs; (4) both Wnt-3a- and Wnt-5ainduced phosphorylation of Dvl is CK1 inhibition sensitive; (5) the block of Wnt-3a-induced phosphorylation of Dvl by CK1 inhibitors can be rescued by Wnt-5a and vice versa; and (6) both Wnt-5a and Wnt-3a directly induce activation of CK1⑀ kinase. Thus, our results comply with the possibility that Wnt- 3a- and Wnt-5a-induced Dvl phosphorylation are mediated by activation of similar or identical signalling complex(es) including CK1␦/⑀. The CK1␦/⑀-mediated phosphorylation of Dvl is necessary for Dvl to interact with other pathway specific components – as demonstrated for the interaction of Dvl with Frat-1 in the canonical Wnt signalling (Hino et al., 2003). This view is supported by our findings, demonstrating the effects of Wnt-5a and CK1⑀ on the localization of Dvl2. On the basis of previous studies (Schwarz-Romond et al., 2005; Smalley et al., 2005) one can expect that Dvl2 puncta are formed predominantly by Dvl multimers. The ability of Wnt-5a and CK1⑀ to promote a more even localization or, in other words, to dissolve the puncta may then reflect a decrease in affinity of Dvl-Dvl interaction (Angers et al., 2006) following CK1␦/⑀mediated phosphorylation of Dvl. Such release of monomeric Dvl from Dvl aggregates might be a necessary step for the interaction of phosphorylated Dvl with other downstream components of Wnt pathway(s). It is important to notice that, although Wnt-3a and Wnt-5a cooperate in Dvl phosphorylation, Wnt-5a diminished Wnt-3ainduced activation of ␤-catenin. Previously, Wnt-5a has been shown to antagonise canonical signalling and different mechanisms were implicated in this process (Maye et al., 2004; Topol et al., 2003; Weidinger and Moon, 2003; Westfall et al., 2003). Our data support this concept but leave the question open of how the signal is redirected from Dvl to the final targets of canonical and non-canonical Wnt signalling pathways. It has been suggested that specific co-receptors play a role in directing the signals into different pathways and our data, demonstrating common signalling unit of canonical and non-canonical Wnt signalling, are well compatible with the crucial role of co-receptors (schematised in Fig. 8). In addition to their common cognate receptors of the Frizzled family, canonical Wnts bind to low-density lipoprotein receptor (LDLR)-related protein 5/6 (Lrp5/6) (Liu et al., 2003; Tamai et al., 2000), whereas non-canonical Wnts interact with the atypical receptor kinase Ror2 (Hikasa et al., 2002; Oishi et al., 2003) or membrane proteoglycan Knypek (Topczewski et al., 2001). A very recent report, showing that a Wnt-5a-Dkk2 CRD fusion can bind Lrp5/6 and activate canonical signalling (Liu et al., 2005), strongly supports a key role of co-receptors in directing canonical versus non-canonical Wnt-signalling. Our results in primary precursor cultures further confirm the importance of CK1 activity for the biological effects of noncanonical Wnts. We demonstrate that CK1 activity is required for the effect of Wnt-5a on the differentiation of dopaminergic precursors into DNs. Thus, our findings argue that CK1 is an essential component of the Wnt-5a-induced signalling pathway not only in a suitable cell line but also in a more complex and biologically relevant system. This finding might also have implications in other areas of biology, such as tumour biology. Wnt-5a is known as a factor promoting cell migration, epithelial mesenchymal transition and increased cancer invasiveness (Taki et al., 2003; Weeraratna et al., 2002). Our findings, linking Wnt-5a to activation of CK1␦/⑀ and showing that CK1␦/⑀ mediate the effects of Wnt-5a, correlate well with the emerging role of CK1␦/⑀ in tumour development (for a review, see Knippschild et al., 2005). The generation and analysis of CK1␦- and/or CK1⑀-deficient mice will certainly help to define how widespread the involvement of CK1␦/⑀ is in vertebrate non-canonical Wnt-signalling. Materials and Methods Cell culture and transfection SN4741 cells were generously provided by J. H. Son (Son et al., 1999) and grown in DMEM, 10% FCS, L-glutamine (2 mM), penicillin (50 U/ml), streptomycin (50 U/ml), glucose (0.6%) (all purchased from Invitrogen). For transfections, 40,000- 60,000 cells/well were seeded in 24-well plates and grown overnight. Cells were transfected under serum-free conditions using Lipofectamine 2000 (Invitrogen) Wnt-5aWnt-3a Dvl CK1δ/ε Lrp5/6 ? P P P DvlCK1δ/ε P P P non-canonical signalling 3a 5a β-catenin canonical signaling Fig. 8. Schematic model illustrating the role of CK1␦/⑀ in the activation of Dvl by Wnt-3a and Wnt-5a. Both Wnt-3a and Wnt-5a bind Frizzled that by unknown mechanism activates CK1␦/⑀, which in turn phosphorylates Dvl. The interaction of Dvl with additional pathway-specific signalling components (including co-receptors), would direct downstream signalling. In case of Wnt-3a the coreceptor might be Lrp5/6, whereas Ror2 or Knypek might be the coreceptors for Wnt-5a. This model predicts that, when both Wnt-3a and Wnt-5a are present, Frizzled and phosphorylated Dvl will compete for binding and activation of additional signalling components. The prevalent signalling direction would thus be determined by the recruitment of the additional signalling units by phosphorylated Dvl. JournalofCellScience 594 according to manufacturer’s instructions. In total, 1 ␮g of DNA encoding Dvl2Myc, CK1⑀ or CK1⑀ (K>R), or their combinations, and 2 ␮l of Lipofectamine 2000 were used per well. Medium was changed 4 hours post transfection and cells were grown in complete culture medium for another 24 hours prior to analysis by western blot or immunocytochemistry. Cell treatments For analysis of intracellular signalling, 40,000 cells/well were seeded in 24-well plates, grown overnight without serum and subsequently stimulated with recombinant mouse Wnt-3a or Wnt-5a (R&D Systems) for 2 hours. Control stimulations were done with equivalent volumes of 0.1% BSA-1% CHAPS-PBS. To screen for compounds that reduce Dvl-mobility, the cells were treated with the various chemical inhibitors (Table 1) for 15 minutes and subsequently stimulated with Wnt. (Note: PTX and PDBu, were added overnight as a pre-stimulus.) Appropriate solvent was used as a control. All compounds were tested in duplicate. The cells were also exposed to FuGENE 6 transfection reagent (Roche, 1 ␮l FuGENE per 200 ␮l of culture media) to enhance the penetration of poorly cellpermeable compounds (D4476 and IC261) into SN4741 cells. Control (DMSOtreated) and experimental conditions were both treated with FUGENE. PTX, PDBu, wortmannin, genistein, chelerythrin, BIM I, SQ22536, MDL12330, AG1278, ET- 18-OCH3 and staurosporine were purchased from Sigma; LY294002 and SB203580 from Tocris; PD98059 and UO126 from Cell Signaling Technology; Ro-31 8220, H89, 8-Br-cAMP, PP2, D4476 and IC261 from Calbiochem and KN93, I3M and kenpaullone from Alexis Biotechnology. 8CPT-2Me-cAMP was a kind gift from J. L. Bos (University of Utrecht, Netherlands). Precursor cultures Embryonic day 14.5 (E14.5) ventral mesencephala obtained from time-mated Sprague-Dawley rats (ethical approval for animal experimentation was granted by Stockholm Norra Djurförsöks Etiska Nämnd) were dissected, mechanically dissociated and plated on plates coated with poly-D-lysine (10 ␮M) at a final density of 1ϫ105 cells/cm2 . Serum-free N2 medium was added, consisting of a mixture of F12 and MEM with N2 supplement, 15 mM HEPES buffer, 1 mM glutamine, 5 mg/ml Albumax (all purchased from Invitrogen) and 6 mg/ml glucose (Sigma). Recombinant Wnt-5a and D4476 (Calbiochem) were added and the cells were cultured for 3 days in a 37°C, 5% CO2 incubator. Cells were fixed for immuncytochemistry in ice-cold 4% paraformaldehyde for 15-20 minutes and washed in PBS. The following primary and secondary antibodies were used: rabbit anti-tyrosine hydroxylase specific antibody (1:100 dilution, Pel-Freez Biologicals) and rhodamine-coupled goat anti-rabbit IgG (1:200; Jackson Laboratories). Cultures were subsequently incubated with Hoechst 33258 reagent for 10 minutes. Images were acquired from stained cells using a Zeiss Axioplan 100M microscope (LD Achroplan 40ϫ, 0.60 Korr PH2 8 0-2) and collected with a Hamamatsu camera C4742-95 (with QED imaging software). TH-immunoreactive cells from two independent experiments, three wells per condition, nine non-overlapping fields per well were counted independently by two researchers. The numbers of TH+ cells represent the mean values ± s.d. and are expressed as percentage change compared with control. Western blotting Western blot analysis and protein samples were prepared as described previously (Bryja et al., 2004). The antibodies used were: anti-Dvl2 (sc-13974), anti-Dvl3 (sc- 8027) and anti-c-Myc (sc-40) (Santa Cruz Biotechnology); anti-␤-catenin (BD Bioscences); anti-active-␤-catenin (anti-ABC, Upstate Biotechnology); antiphospho-␤-catenin (S45, Biosource) anti-phospho-serin (AB1603, Chemicon International) and anti-MBP (Dako). Immunoprecipitation and kinase assay For kinase assays, total protein from the cells (cultured in 10-cm dishes) was extracted and processed using Protein G Sepharose fast-flow beads (Amersham Biosciences) as described previously (Bryja et al., 2005). Rabbit polyclonal antibody against Myc (sc-789), mouse agarose-conjugated antibody against Myc (sc-40 AC) and goat polyclonal antibody against CK1⑀ (sc-6471) were used in the immunoprecipitation studies (Santa Cruz Biotechnology). CK1⑀ kinase reactions were carried out for 15 minutes at room temperature in a 40-␮l volume of kinaseassay buffer (50 mM HEPES pH 7.5, 10 mM MgCl2, 10 mM MnCl2, 20 mM ␤glycerolphosphate, 5 mM NaF) supplemented with 100 ␮g/ml MBP (M1831, Sigma) and 100 ␮M ATP. Reactions were terminated by addition of 5ϫ Laemmli sample buffer. Each reaction mix was then subjected to SDS-PAGE. RNA interference and quantitative RT-PCR SN4741 cells were transfected with siRNA using neofection according to manufacturer’s instructions (Ambion). In brief, siRNAs (0.75 ␮l of 20 ␮M siRNA) were mixed with Lipofectamine 2000 (2 ␮l; Invitrogen) and OptiMEM (47.25 ␮l; Gibco) and incubated for 20 minutes at room temperature. The transfection mixture (50 ␮l) was added to the 24-well plate and mixed with a suspension of freshly trypsinised SN4741 cells (25,000 cells/well in 500 ␮l of complete media) resulting in the final concentration of 30 nM siRNA. When a combination of two different Journal of Cell Science 120 (4) siRNAs was used, each siRNA was used at 30 nM and the control siRNA at 60 nM. The transfection was terminated after 5 hours by changing culture media. At 48 hours post transfection, cells were stimulated with Wnt-5a and collected for further analyses. siRNAs against individual isoforms of mouse CK1 were purchased from Ambion; CK1␣ (I, cat. no. 176063; II, cat. no. 176062; III, cat. no. 176061), CK1␦ (I, cat. no. 88202; II, cat. no. 88309; III, cat. no. 88298) and CK1⑀ (I, cat. no.188527; II, cat. no. 188528, III , cat. no. 188529). Silencer® Negative Control siRNA (cat. no. 4635, Ambion) was used as a negative control. The efficiency of the silencing was assessed by western blotting or real time RT-PCR. Quantitative RT-PCR was performed as described previously (Castelo-Branco et al., 2006) The following primers were used (DNA Technology A/S, Aarhus, Denmark): CK1␣for 5ЈTTTGAGGAAGCTCCGGATTACAT-3Ј, CK1␣rev 5Ј-TCGTCCAATCAAACGTGTAGTCAT-3´, CK1␦for 5Ј-ACATCTATCTCGGTACGGACATTG-3Ј, CK1␦rev 5Ј- GAGGATGTTTGGTTTTGACACATTC-3Ј. Confocal imaging SN4741 cells (20,000-40,000 cells/well) were grown overnight on glass coverslips and transfected with the indicated plasmids. Treatment with chemical inhibitors or Wnt-5a was performed at 4 hours or 9 hours post transfection. 24 hours post transfection cells were fixed in 4% paraformaldehyde for 15 minutes. For immunodetection cells were washed three times in PBS and blocked with 1% BSA, 0.1% Triton X-100 in PBS for 1 hour. The primary antibodies (see western blotting section) were incubated for 3 hours at room temperature and subsequently the appropriate secondary antibodies coupled to Cy3 or Cy2 (1:500, Jackson Immunoresearch) were applied for 2 hours at room temperature. After washing, coverslips were mounted on slides using glycerol gelatine-mounting medium (Sigma-Aldrich). Fluorescent labelling was examined using a Zeiss LSM510 confocal system including a Zeiss Axioplan2 microscope equipped with filters for the detection of Cy2 and Cy3. Financial support was obtained from the Swedish Foundation for Strategic Research, Swedish Royal Academy of Sciences, Knut and Alice Wallenberg Foundation, European Commission, Swedish MRC, Karolinska Institutet, Svenska Läkaresällskapet and Lars Hiertas Minnesfond. G.S. was supported by a post-doctoral fellowship from the Swedish Society for Medical Research (SSMF). We thank J. H. Son for SN4741 cells, G. Castelo-Branco for Fz8-CRD conditioned media and J. L. Bos (University of Utrecht, Netherlands) for 8CPT- 2Me-cAMP. We also thank S. Yanagawa (Kyoto University, Japan), R. J. Lefkowitz (HHMI, Durham, NC) and J. M. Graff (University of Texas, Dallas, TX) for expression vectors encoding Dvl2-Myc, CK1⑀ and CK1⑀ (K>R), respectively; B. B. Fredholm (Karolinska Institutet, Sweden) for assistance with microscopy; Clare Parish and G. CasteloBranco, Emma Andersson, and Kyle Sousa for critical reading of the manuscript. Thank you to Claudia Tello, Johny Söderlund and Annika Käller for technical and secretarial assistance, and to the members of E.A.’s lab for stimulating discussions. References Amit, S., Hatzubai, A., Birman, Y., Andersen, J. S., Ben-Shushan, E., Mann, M., BenNeriah, Y. and Alkalay, I. (2002). Axin-mediated CKI phosphorylation of betacatenin at Ser 45, a molecular switch for the Wnt pathway. Genes Dev. 16, 1066-1076. Angers, S., Thorpe, C. J., Biechele, T. L., Goldenberg, S. J., Zheng, N., MacCoss, M. J. and Moon, R. T. (2006). The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wnt-␤-catenin pathway by targeting Dishevelled for degradation. Nat. Cell Biol. 8, 348-357. Arenas, E. (2005). Engineering a dopaminergic phenotype in stem/precursor cells: role of Nurr1, Glia-derived signals, and Wnts. Ann. N. Y. Acad. Sci. 1049, 51-66. Bryja, V., Pachernik, J., Soucek, K., Horvath, V., Dvorak, P. and Hampl, A. (2004). Increased apoptosis in differentiating p27-deficient mouse embryonic stem cells. Cell. Mol. Life Sci. 61, 1384-1400. Bryja, V., Cajanek, L., Pachernik, J., Hall, A. C., Horvath, V., Dvorak, P. and Hampl, A. (2005). Abnormal development of mouse embryoid bodies lacking p27Kip1 cell cycle regulator. Stem Cells 23, 965-974. Bryja, V., Schulte, G. and Arenas, E. (2007). Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate ␤-catenin. Cell. Signal. 19, 610-616. Castelo-Branco, G., Wagner, J., Rodriguez, F. J., Kele, J., Sousa, K., Rawal, N., Pasolli, H. A., Fuchs, E., Kitajewski, J. and Arenas, E. (2003). Differential regulation of midbrain dopaminergic neuron development by Wnt-1, Wnt-3a, and Wnt- 5a. Proc. Natl. Acad. Sci. USA 100, 12747-12752. Castelo-Branco, G., Sousa, K. M., Bryja, V., Pinto, L., Wagner, J. and Arenas, E. (2006). Ventral midbrain glia express region-specific transcription factors and regulate dopaminergic neurogenesis through Wnt-5a secretion. Mol. Cell. Neurosci. 31, 251- 262. Cong, F., Schweizer, L. and Varmus, H. (2004). Casein kinase Iepsilon modulates the signaling specificities of dishevelled. Mol. Cell. Biol. 24, 2000-2011. JournalofCellScience 595CK1 mediates Wnt-5a action Davidson, G., Wu, W., Shen, J., Bilic, J., Fenger, U., Stannek, P., Glinka, A. and Niehrs, C. (2005). Casein kinase 1 gamma couples Wnt receptor activation to cytoplasmic signal transduction. Nature 438, 867-872. Fish, K. J., Cegielska, A., Getman, M. E., Landes, G. M. and Virshup, D. M. (1995). Isolation and characterization of human casein kinase I epsilon (CKI), a novel member of the CKI gene family. J. Biol. Chem. 270, 14875-14883. Gao, Z. H., Seeling, J. M., Hill, V., Yochum, A. and Virshup, D. M. (2002). Casein kinase I phosphorylates and destabilizes the beta-catenin degradation complex. Proc. Natl. Acad. Sci. USA 99, 1182-1187. Gonzalez-Sancho, J. M., Brennan, K. R., Castelo-Soccio, L. A. and Brown, A. M. (2004). Wnt proteins induce dishevelled phosphorylation via an LRP5/6- independent mechanism, irrespective of their ability to stabilize beta-catenin. Mol. Cell. Biol. 24, 4757-4768. Hikasa, H., Shibata, M., Hiratani, I. and Taira, M. (2002). The Xenopus receptor tyrosine kinase Xror2 modulates morphogenetic movements of the axial mesoderm and neuroectoderm via Wnt signaling. Development 129, 5227-5239. Hino, S., Michiue, T., Asashima, M. and Kikuchi, A. (2003). Casein kinase I epsilon enhances the binding of Dvl-1 to Frat-1 and is essential for Wnt-3a-induced accumulation of beta-catenin. J. Biol. Chem. 278, 14066-14073. Hsieh, J. C. (2004). Specificity of WNT-receptor interactions. Front. Biosci. 9, 1333- 1338. Hsieh, J. C., Rattner, A., Smallwood, P. M. and Nathans, J. (1999). Biochemical characterization of Wnt-frizzled interactions using a soluble, biologically active vertebrate Wnt protein. Proc. Natl. Acad. Sci. USA 96, 3546-3551. Huelsken, J. and Birchmeier, W. (2001). New aspects of Wnt signaling pathways in higher vertebrates. Curr. Opin. Genet. Dev. 11, 547-553. Kishida, M., Hino, S., Michiue, T., Yamamoto, H., Kishida, S., Fukui, A., Asashima, M. and Kikuchi, A. (2001). Synergistic activation of the Wnt signaling pathway by Dvl and casein kinase Iepsilon. J. Biol. Chem. 276, 33147-33155. Klein, T. J., Jenny, A., Djiane, A. and Mlodzik, M. (2006). CKIvarepsilon/discs overgrown promotes both Wnt-Fz/beta-catenin and Fz/PCP signaling in Drosophila. Curr. Biol. 16, 1337-1343. Knippschild, U., Gocht, A., Wolff, S., Huber, N., Lohler, J. and Stoter, M. (2005). The casein kinase 1 family: participation in multiple cellular processes in eukaryotes. Cell. Signal. 17, 675-689. Lee, J. S., Ishimoto, A. and Yanagawa, S. (1999). Characterization of mouse dishevelled (Dvl) proteins in Wnt/Wingless signaling pathway. J. Biol. Chem. 274, 21464-21470. Liu, C., Li, Y., Semenov, M., Han, C., Baeg, G. H., Tan, Y., Zhang, Z., Lin, X. and He, X. (2002). Control of beta-catenin phosphorylation/degradation by a dual-kinase mechanism. Cell 108, 837-847. Liu, G., Bafico, A., Harris, V. K. and Aaronson, S. A. (2003). A novel mechanism for Wnt activation of canonical signaling through the LRP6 receptor. Mol. Cell. Biol. 23, 5825-5835. Liu, G., Bafico, A. and Aaronson, S. A. (2005). The mechanism of endogenous receptor activation functionally distinguishes prototype canonical and noncanonical wnts. Mol. Cell. Biol. 25, 3475-3482. Mashhoon, N., DeMaggio, A. J., Tereshko, V., Bergmeier, S. C., Egli, M., Hoekstra, M. F. and Kuret, J. (2000). Crystal structure of a conformation-selective casein kinase- 1 inhibitor. J. Biol. Chem. 275, 20052-20060. Matsubayashi, H., Sese, S., Lee, J. S., Shirakawa, T., Iwatsubo, T., Tomita, T. and Yanagawa, S. (2004). Biochemical characterization of the Drosophila wingless signaling pathway based on RNA interference. Mol. Cell. Biol. 24, 2012-2024. Maye, P., Zheng, J., Li, L. and Wu, D. (2004). Multiple mechanisms for Wnt11mediated repression of the canonical Wnt signaling pathway. J. Biol. Chem. 279, 24659-24665. McKay, R. M., Peters, J. M. and Graff, J. M. (2001a). The casein kinase I family in Wnt signaling. Dev. Biol. 235, 388-396. McKay, R. M., Peters, J. M. and Graff, J. M. (2001b). The casein kinase I family: roles in morphogenesis. Dev. Biol. 235, 378-387. McMahon, A. P. and Bradley, A. (1990). The Wnt-1 (int-1) proto-oncogene is required for development of a large region of the mouse brain. Cell 62, 1073-1085. Oishi, I., Suzuki, H., Onishi, N., Takada, R., Kani, S., Ohkawara, B., Koshida, I., Suzuki, K., Yamada, G., Schwabe, G. C. et al. (2003). The receptor tyrosine kinase Ror2 is involved in non-canonical Wnt5a/JNK signalling pathway. Genes Cells 8, 645- 654. Patapoutian, A. and Reichardt, L. F. (2000). Roles of Wnt proteins in neural development and maintenance. Curr. Opin. Neurobiol. 10, 392-399. Peters, J. M., McKay, R. M., McKay, J. P. and Graff, J. M. (1999). Casein kinase I transduces Wnt signals. Nature 401, 345-350. Pinson, K. I., Brennan, J., Monkley, S., Avery, B. J. and Skarnes, W. C. (2000). An LDL-receptor-related protein mediates Wnt signalling in mice. Nature 407, 535- 538. Prakash, N., Brodski, C., Naserke, T., Puelles, E., Gogoi, R., Hall, A., Panhuysen, M., Echevarria, D., Sussel, L., Weisenhorn, D. M. et al. (2006). A Wnt1-regulated genetic network controls the identity and fate of midbrain-dopaminergic progenitors in vivo. Development 133, 89-98. Price, M. A. (2006). CKI, there’s more than one: casein kinase I family members in Wnt and Hedgehog signaling. Genes Dev. 20, 399-410. Rena, G., Bain, J., Elliott, M. and Cohen, P. (2004). D4476, a cell-permeant inhibitor of CK1, suppresses the site-specific phosphorylation and nuclear exclusion of FOXO1a. EMBO Rep. 5, 60-65. Sakanaka, C., Leong, P., Xu, L., Harrison, S. D. and Williams, L. T. (1999). Casein kinase I epsilon in the wnt pathway: regulation of beta-catenin function. Proc. Natl. Acad. Sci. USA 96, 12548-12552. Schulte, G., Bryja, V., Rawal, N., Castelo-Branco, G., Sousa, K. M. and Arenas, E. (2005). Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation. J. Neurochem. 92, 1550-1553. Schwarz-Romond, T., Merrifield, C., Nichols, B. J. and Bienz, M. (2005). The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J. Cell Sci. 118, 5269-5277. Shimizu, H., Julius, M. A., Giarre, M., Zheng, Z., Brown, A. M. and Kitajewski, J. (1997). Transformation by Wnt family proteins correlates with regulation of betacatenin. Cell Growth Differ. 8, 1349-1358. Smalley, M. J., Signoret, N., Robertson, D., Tilley, A., Hann, A., Ewan, K., Ding, Y., Paterson, H. and Dale, T. C. (2005). Dishevelled (Dvl-2) activates canonical Wnt signalling in the absence of cytoplasmic puncta. J. Cell Sci. 118, 5279-5289. Son, J. H., Chun, H. S., Joh, T. H., Cho, S., Conti, B. and Lee, J. W. (1999). Neuroprotection and neuronal differentiation studies using substantia nigra dopaminergic cells derived from transgenic mouse embryos. J. Neurosci. 19, 10-20. Strutt, H., Price, M. A. and Strutt, D. (2006). Planar polarity is positively regulated by casein kinase ivarepsilon in Drosophila. Curr. Biol. 16, 1329-1336. Swiatek, W., Tsai, I. C., Klimowski, L., Pepler, A., Barnette, J., Yost, H. J. and Virshup, D. M. (2004). Regulation of casein kinase I epsilon activity by Wnt signaling. J. Biol. Chem. 279, 13011-13017. Takada, R., Hijikata, H., Kondoh, H. and Takada, S. (2005). Analysis of combinatorial effects of Wnts and Frizzleds on beta-catenin/armadillo stabilization and Dishevelled phosphorylation. Genes Cells 10, 919-928. Taki, M., Kamata, N., Yokoyama, K., Fujimoto, R., Tsutsumi, S. and Nagayama, M. (2003). Down-regulation of Wnt-4 and up-regulation of Wnt-5a expression by epithelial-mesenchymal transition in human squamous carcinoma cells. Cancer Sci. 94, 593-597. Tamai, K., Semenov, M., Kato, Y., Spokony, R., Liu, C., Katsuyama, Y., Hess, F., Saint-Jeannet, J. P. and He, X. (2000). LDL-receptor-related proteins in Wnt signal transduction. Nature 407, 530-535. Thomas, K. R. and Capecchi, M. R. (1990). Targeted disruption of the murine int-1 proto-oncogene resulting in severe abnormalities in midbrain and cerebellar development. Nature 346, 847-850. Topczewski, J., Sepich, D. S., Myers, D. C., Walker, C., Amores, A., Lele, Z., Hammerschmidt, M., Postlethwait, J. and Solnica-Krezel, L. (2001). The zebrafish glypican knypek controls cell polarity during gastrulation movements of convergent extension. Dev. Cell 1, 251-264. Topol, L., Jiang, X., Choi, H., Garrett-Beal, L., Carolan, P. J. and Yang, Y. (2003). Wnt-5a inhibits the canonical Wnt pathway by promoting GSK-3-independent betacatenin degradation. J. Cell Biol. 162, 899-908. van Noort, M., Meeldijk, J., van der Zee, R., Destree, O. and Clevers, H. (2002). Wnt signaling controls the phosphorylation status of beta-catenin. J. Biol. Chem. 277, 17901-17905. Veeman, M. T., Axelrod, J. D. and Moon, R. T. (2003). A second canon. Functions and mechanisms of beta-catenin-independent Wnt signaling. Dev. Cell 5, 367-377. Wallingford, J. B. and Habas, R. (2005). The developmental biology of Dishevelled: an enigmatic protein governing cell fate and cell polarity. Development 132, 4421-4436. Wang, Y., Thekdi, N., Smallwood, P. M., Macke, J. P. and Nathans, J. (2002). Frizzled- 3 is required for the development of major fiber tracts in the rostral CNS. J. Neurosci. 22, 8563-8573. Weeraratna, A. T., Jiang, Y., Hostetter, G., Rosenblatt, K., Duray, P., Bittner, M. and Trent, J. M. (2002). Wnt5a signaling directly affects cell motility and invasion of metastatic melanoma. Cancer Cell 1, 279-288. Weidinger, G. and Moon, R. T. (2003). When Wnts antagonize Wnts. J. Cell Biol. 162, 753-755. Westfall, T. A., Brimeyer, R., Twedt, J., Gladon, J., Olberding, A., Furutani-Seiki, M. and Slusarski, D. C. (2003). Wnt-5/pipetail functions in vertebrate axis formation as a negative regulator of Wnt/beta-catenin activity. J. Cell Biol. 162, 889-898. Wharton, K. A., Jr (2003). Runnin’ with the Dvl: proteins that associate with Dsh/Dvl and their significance to Wnt signal transduction. Dev. Biol. 253, 1-17. Yamaguchi, T. P. (2001). Heads or tails: Wnts and anterior-posterior patterning. Curr. Biol. 11, R713-R724. Zeng, X., Tamai, K., Doble, B., Li, S., Huang, H., Habas, R., Okamura, H., Woodgett, J. and He, X. (2005). A dual-kinase mechanism for Wnt co-receptor phosphorylation and activation. Nature 438, 873-877. Zhang, Y., Qiu, W. J., Chan, S. C., Han, J., He, X. and Lin, S. C. (2002). Casein kinase I and casein kinase II differentially regulate axin function in Wnt and JNK pathways. J. Biol. Chem. 277, 17706-17712. JournalofCellScience Vítězslav Bryja, 2014    Attachments      #4      Bryja V, Čajánek L, Grahn A, Schulte G (2007): Inhibition of endocytosis blocks Wnt  signalling to ‐catenin by promoting dishevelled degradation. Acta Physiol.(Oxford) 190  (1): 53‐59      Impact factor (2007): 2.455  Times cited (without autocitations, WoS, Feb 21st 2014): 13  Significance: Here we described that block of endocytosis in several cell types leads to  complete  depletion  of  endogenous  Dvl  in  cells.  This  observation  demonstrated  that Dvl has a rapid turnover and that its physiological levels has to be actively  maintained.  Contibution of the author/author´s team: Identification and biochemical description of  the role of endocytosis in the regulation of Dvl levels.                Inhibition of endocytosis blocks Wnt signalling to b-catenin by promoting dishevelled degradation V. Bryja,1,2 L. Cˇ aja´nek,1 A. Grahn,1 and G. Schulte3 1 Department of Medical Biochemistry & Biophysics, Molecular Neurobiology, Karolinska Institutet, Stockholm, Sweden 2 Institute of Experimental Biology, Faculty of Science, Masaryk University Brno and Department of Cytokinetics, Academy of Sciences of the Czech Republic, Kralovopolska, Czech Republic 3 Department of Physiology & Pharmacology, Karolinska Institutet, Sec. Receptor Biology & Signalling, Stockholm, Sweden Received 29 September 2006, revision requested 23 October 2006, revision received 1 November 2006, accepted 3 November 2006 Correspondence: G. Schulte, PhD, Department of Physiology & Pharmacology, Karolinska Institutet, Sec. Receptor Biology & Signalling, Nanna Svartz va¨g 2, S-17177 Stockholm, Sweden. E-mail: gunnar.schulte@ki.se Abstract Aim: The Wnt/Frizzled signalling pathway is highly conserved through evolution. Frizzled, the receptors for Wnts, have the topology of seven transmembrane spanning domain receptors. An important means of regulation of these receptors is internalization and desensitization through clathrinmediated endocytosis. Therefore, we investigated the effects of endocytosis inhibition on Frizzled4-green fluorescent protein (FZD4-GFP) localization, dishevelled levels and Wnt-3a signalling to b-catenin. Methods: Experiments were performed in the mouse neuronal cell line SN4741 that has previously proven to be valuable for the investigation of Wnt/Frizzled signalling. FZD4-GFP distribution has been examined using confocal laser scanning microscopy. Dishevelled protein expression levels and the activation of b-catenin upon treatment with endocytosis inhibitors (hyperosmolaric sucrose and K+ depletion), kinase inhibitors and Wnt-3a were analysed by immunoblotting. Results: Hyperosmotic sucrose and K+ depletion increased the membrane localization of FZD4-GFP, and in parallel triggered fast (1–2 h) and almost complete (approx. 95%) degradation of endogenous dishevelled, which was independent of Wnt-induced, CK1-mediated phosphorylation of dishevelled. In addition, dishevelled depletion induced by endocytosis inhibition completely prevented canonical signalling by Wnt-3a to b-catenin even when osmotic conditions and endocytosis were reverted to normal. Conclusions: The data provide evidence for a molecular mechanism that could be a basis for a novel negative feedback loop within the Wnt/Frizzled pathway depending on dishevelled degradation. The identification of molecular details of regulatory mechanisms for the Wnt/Frizzled signalling pathway increases our understanding of pathway regulation, which might be of special physiological significance for embryonic development, cancer and neurological disorders. Keywords b-catenin, clathrin-mediated endocytosis, desensitization, dishevelled, Dvl, frizzled, sucrose, Wnt signalling. Cellular communication via the Wnt/Frizzled (FZD) signalling pathway is highly conserved through evolution (Huelsken & Birchmeier 2001). This pathway appears to be of crucial importance during embryonic development, for the regulation of appropriate cell proliferation, the biology of cancer, neurological Acta Physiol 2007, 190, 55–61 Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x 55 disorders and other diseases (Logan & Nusse 2004, Nusse 2005). The Wnts are lipoglycoproteins (Nusse 2003) that bind to their cognate seven transmembranespanning receptors, the Frizzled family of receptors (Malbon 2004), which belong to the class of G proteincoupled receptors (GPCRs) (Foord et al. 2005). Historically, Wnt signalling was divided into canonical, i.e. b-catenin-dependent signalling and non-canonical signalling. Non-canonical pathways are independent of b-catenin, e.g. the polar cell polarity pathway or the Ca2+ pathway (Logan & Nusse 2004). Even though molecular details in the Wnt signalling pathway are still waiting for a complete understanding, basic concepts are established. Wnt binding to Frizzleds and co-receptors, e.g. the low density lipoprotein-related protein (LRP) family, namely LRP5 and 6, induces the activation of the important signalling relay dishevelled (Dvl; Malbon & Wang 2006). Activation of Dvl in the b-catenin pathway results in inhibition of a constitutive destruction complex for b-catenin consisting of Adenomatous Polyposis Coli (APC), axin and glycogen synthase kinase 3b (GSK3b). This in turn prevents proteasomal degradation of b-catenin, leading to cytoplasmatic stabilization of the protein and a final translocation to the nucleus. There it will exert gene transcriptional modulation in cooperation with TCF/ Lef family transcription factors (for more information see also the Wnt homepage http://www.stanford.edu/ $rnusse/wntwindow.html). It has been shown that endocytosis and especially clathrin-mediated endocytosis is of importance for the Wnt/Frizzled pathway: the b-arrestin-dependent receptor internalization through clathrin-coated pits is a hallmark of classical GPCRs. It has been shown to occur in the Wnt-5a-induced activation of Frizzled4green fluorescent protein (FZD4-GFP) (Chen et al. 2003). That study presented the phosphoprotein Dvl as an important link between FZD and the endocytotic machinery, especially b-arrestin2. Furthermore, Wnt internalization was reported to be important for creating Wnt gradients for proper embryonic development (Seto & Bellen 2006) and very recently, Blitzer & Nusse (2006) described canonical b-catenin signalling being dependent on Wnt-3a internalization through clathrin-coated pits. Furthermore, caveolin-dependent endocytosis of the Wnt co-receptor LRP6 was shown to be necessary for Wnt-3a induced b-catenin signalling (Yamamoto et al. 2006). It appears that endocytosis is an important feature of Wnt signalling and the regulation of this essential way of cellular communication (Kikuchi & Yamamoto, 2007). However, the importance of endocytosis and the consequences of inhibition of endocytosis have not been investigated yet on the level of the Dvl. A signalling relay station such as Dvl, which serves as a central crossroads of Wnt/Frizzled signalling to canonical and non-canonical pathways needs to be tightly regulated to allow fine-tuning of the signalling system (Malbon & Wang 2006). Modulation of Dvl is achieved by posttranslational modifications such as activating phosphorylation through a variety of kinases, e.g. casein kinases 1 and 2 and protein kinase C (for a summary, see e.g. Wharton 2003). Furthermore, Dvl ubiquitination mediates proteosomal degradation (Miyazaki et al. 2004, Simons et al. 2005, Angers et al. 2006). Here, we show that inhibition of endocytosis by hyperosmolar sucrose or K+ depletion resulted in the redistribution of overexpressed FZD4-GFP from predominantly cytoplasmic localization to the cell membrane. In parallel, the block of clathrin-mediated endocytosis induced a rapid degradation of Dvl, leading to a nearly complete depletion of Dvl from the neuronal dopaminergic cell line SN4741. In addition, we show that the lack of Dvl in these cells after inhibition of endocytosis blocked Wnt-3a-induced signalling to b-catenin. Thus, we identify a novel molecular mechanism in the Wnt/FZD signalling pathway with all the properties of a negative feedback mechanism based on the degradation of the central signalling module Dvl. Material and methods Cell culture, treatments and transfection The mouse neuron-like, dopaminergic cell line SN4741 was provided by Dr J.H. Son (Son et al. 1999) and grown in Dulbecco’s modified Eagle’s medium (DMEM), 10% FCS, l-glutamine (2 mm), penicillin (50 U mL)1 ), streptomycin (50 U mL)1 ), glucose (0.6%) (all purchased from Invitrogen, San Diego, CA, USA). For inhibition of endocytosis, cells were incubated either in growth medium containing 0.45 m sucrose for 2 h or depleted of K+ according to a protocol described previously (Larkin et al. 1983): cells were pretreated with hypotonic medium (DMEM : water 1 : 1) for 5 min and then incubated in isotonic K+ -free buffer [50 mm N-2-hydroxyethylpiperazine-N¢-2-ethane sulphonic acid (HEPES), 100 mm NaCl] for the indicated times. Casein kinase 1 inhibitors (from Calbiochem, Nottingham, UK) D4476 (100 lm) and IC261 (50 lm) were applied 2 h before sucrose treatment, control stimulation was performed with DMSO. Cells were stimulated with Wnt-3a (from R&D Systems, Abingdon, UK) at 100 ng mL)1 . For imaging, cells were seeded on sterile coverslips in 24 well plates, grown overnight and transfected under serum-free conditions using Lipofectamine 2000 (Invitrogen) according to manufacturer’s instructions. Medium was changed 4 h post-transfection and cells were grown in complete culture medium for another 24 h before fixation with 56 Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x Rapid Dvl degradation by endocytosis block ÆV Bryja et al. Acta Physiol 2007, 190, 55–61 4% paraformaldehyde. For imaging of GFP expression cells were washed three times in phosphate-buffered saline and coverslips were mounted on slides using glycerol gelatine-mounting medium (Sigma-Aldrich, Stockholm, Sweden). Fluorescence was examined using a Zeiss LSM510 confocal system including a Zeiss Axioplan2 (Zeiss, Jena, Germany) microscope equipped with filters for the detection of GFP. Immunoblotting Immunoblot analysis and protein samples were prepared as described previously (Bryja et al. 2004). In brief, samples were subjected to polyacrylamide gel electrophoresis (PAGE), transferred to a polyvinylidene difluoride (PVDF) membrane and the following antibodies were used for protein detection: anti-Dvl2 (sc-13974), anti-Dvl3 (sc-8027) (from Santa Cruz Biotechnology, Santa Cruz, CA, USA). Anti-b-catenin (BD Biosciences, Stockholm, Sweden); anti-active-bcatenin (anti-ABC, Upstate Biotechnology, Cambridge, UK); anti-actin from (Abcam, Hampshire, UK). Results We were interested in the effects of inhibition of endocytosis on the regulation of FZDs and the central Wnt signalling module Dvl in the dopaminergic neuronal cell line SN4741 that we have characterized extensively with regard to Wnt signalling (Schulte et al. 2005, Bryja et al. 2007a,b). To prevent clathrin-coated pits formation and thereby the budding and internalization of clathrin-coated vesicles, we applied hyperosmolaric (0.45 m) sucrose (Heuser & Anderson 1989) or K+ depletion (Larkin et al. 1983, 1986) for indicated times. We were using overexpression of green-fluorescent protein-tagged FZD4 (FZD4-GFP) to monitor the effect of endocytosis inhibition on the receptor localization. In the dopaminergic cell line SN4741 (Son et al. 1999) FZD4-GFP overexpression resulted in a predominantly cytoplasmatic labelling with very little FZD4GFP in the plasma membrane (Fig. 1a). This implicates a continuous receptor uptake into the cell. The inhibition of endocytosis by 0.45 m sucrose or potassium depletion altered the picture and we detected mainly membranous FZD4-GFP signal (Fig. 1b). This strengthened the assumption that cytoplasmatic receptors represent receptors that have been inserted in the membrane and subsequently incorporated by endocytosis. Furthermore, we conclude that FZD4-GFP embedding in the membrane was not impaired by, e.g. overexpressing the tagged receptor. We suggest that constitutive internalization might, e.g. be due to receptor overexpression or autocrine stimulation of the receptor due to endogenously expressed Wnts (G. Schulte, unpublished data). These data indicate that the treatments with endocytosis blockers were suitable for the purpose of the study as they promoted membranous localization of presumably internalized FZD4-GFP. In parallel, inhibition of endocytosis for 2 h led to a strong downregulation of endogenous Dvl2 and 3 (Fig. 1c), the main isoforms of the Dvl family expressed in SN4741 cells (Bryja et al. 2007a). Other cell types were used to test whether the rapid and efficient degradation of Dvl in response to inhibition of endocytosis is a more general phenomenon and takes places in non-neuronal cells as well. We treated mouse embryonic fibroblast (MEF) and human embryonic kidney cells (Hek293) with 0.45 m sucrose and obtained similar results (Fig. 1d). This suggests that No endocytosisctrl FZD4-GFP 10 µm cFZD4-GFP Dvl2 Sucrose K+ depletion Actin Dvl3 – + – + Dvl2 Sucrose – + MEFs Hek293 Sucrose – + Dvl2 ERK1 (a) (b) (c) (d) Figure 1 Confocal images showing SN4741 cells overexpressing FZD4-GFP. Cells were untreated (ctrl, a) or treated (b) with 0.45 m sucrose (insert in b: K+ -depletion) to prevent clathrin-coated pit-mediated endocytosis. Bar ¼ 10 lm. Dishevelled levels are reduced upon blockade of endocytosis by hyperosmolaric sucrose or K+ -depletion. Fig. 1c shows immunoblotting detection of Dvl2 and 3 isoforms as well as actin as a loading control in SN4741 after 2 h of either sucrose (0.45 m) treatment or K+ depletion. Fig. 1d shows the detection of Dvl2 (filled arrowheads) in mouse embryonic fibroblasts (MEFs) and human embryonic kidney cells (Hek293) in the presence or absence of 0.45 m sucrose. In MEFs the Dvl-2 antibody shows an unspecific band at lower molecular weight, serving as an internal loading control. Detection of ERK1 was shown for Hek293 cells. Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x 57 Acta Physiol 2007, 190, 55–61 V Bryja et al. ÆRapid Dvl degradation by endocytosis block this mechanism is general rather than being a cell-line specific phenomenon. In addition, we noted-especially after sucrose treatment-that Dvl degradation was preceded by the formation of a higher molecular weight Dvl resulting in a rapid change in electrophoretic mobility at 10 min (Fig. 2). The changes in electrophoretic mobility of Dvl2 and Dvl3 indicate the existence of post-translational modification of Dvl preceding its degradation. This modification is most likely ubiquitination, which was described to mediate Dvl degradation in other cellular contexts (Miyazaki et al. 2004, Simons et al. 2005, Angers et al. 2006). Inhibition of endocytosis by K+ depletion clearly reduced levels of Dvl (Fig. 1D) but did not show the same fast and dramatic kinetics as sucrose, although higher molecular forms of Dvl were also detectable (Fig. 2). Previously, we have shown that canonical and noncanonical Wnts induce the phosphorylation-dependent electrophoretic mobility shift of Dvl (PS-Dvl) in a casein kinase 1d/e-dependent manner through similar if not identical mechanisms (Bryja et al. 2007a,b). To examine whether the CK1-dependent phosphorylation of Dvl precedes the degradation we used the casein kinase 1 inhibitors D4476 (Rena et al. 2004) and IC261 (Mashhoon et al. 2000). As shown in Figure 3a, these two different CK1 inhibitors, both prevented the formation of the PS-Dvl. Interestingly, in combination with hyperosmolaric sucrose, neither inhibition of CK1 with D4476 nor with IC261 could block the rapid Dvl degradation (Fig. 3b). Therefore, we conclude that Dvl is degraded independently of its activity status and that PS-Dvl formation is not necessary for Dvl to be prone to a rapid degradation induced by inhibition of endo- cytosis. Next, we investigated whether endocytosis inhibition associated with rather complete depletion of Dvl in SN4741 might affect Wnt-3a-induced canonical signalling. Pre-treatment with either hyperosmolaric sucrose (2 h; Fig. 4a) or K+ -depletion (2 h; Fig. 4b) led to a strong downregulation of Dvl2 expression as assessed by immunoblotting. Basal b-catenin phosphorylation was not affected by endocytosis inhibition as measured with the ABC antibody, recognizing the active, dephosphorylated form of b-catenin. However, the strong increase in ABC observed under control conditions after 2 h treatment with Wnt-3a (100 ng mL)1 ) was completely blocked when endocytosis was inhibited, indicating that canonical signalling was not functional when (i) endocytosis was blocked, and/or (ii) the cells are depleted of Dvl. The experimental set-up of inhibition of endocytosis leading to the downregulation of Dvl did not allow us to distinguish between these two possibilities. Therefore, we aimed at designing conditions where Dvl levels are low but endocytosis is allowed. This could be achieved by washing out the sucrose from the cells after sufficient Dvl depletion and subsequent stimulation with Wnt-3a under osmotically normal conditions. Therefore, we changed the experimental paradigm: initial sucrose pretreatment led to a downregulation of Dvl levels and subsequent washing re-established physiologically normal osmotic conditions allowing endocytosis. Under these conditions, levels of Dvl are very low in comparison with control (Fig. 5). Importantly, Wnt-3a stimuMin 0 10 30 60 120 Sucrose Dvl2 Dvl3 0 10 30 60 240 K+ depletion Actin 120 Figure 2 Dishevelled dynamics upon inhibition of endocytosis over time. Time course of both sucrose treatment (a) and K+ depletion, (b) presenting protein levels of Dvl2/3 and actin. Dvl2 CTRL D4476 IC261 CTRL D4476 IC261 Min 0 10 60 0 10 60 0 10 60 Dvl2 Dvl3 Actin Sucrose (a) (b) Figure 3 The degradation of Dvl induced by block of endocytosis is independent of phosphorylation via casein kinases 1. The casein kinase 1 inhibitors D4476 and IC261 block the formation of the slow-migrating Dvl2 band in SN4741 cells (a). (b) shows immunoblotting detection of Dvl2/3 and actin as a loading control in the presence of D4476 (100 lm) or IC261 (50 lm) and 0, 10 and 60 min of sucrose treatment. 58 Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x Rapid Dvl degradation by endocytosis block ÆV Bryja et al. Acta Physiol 2007, 190, 55–61 lation for 2 h either 4 or 6 h after washing out 0.45 m sucrose did not induce b-catenin activation (ABC) compared with the control cells, which showed normal Dvl levels (Fig. 5). These data demonstrate that depletion of Dvl induced by endocytosis block is sufficient to block completely the Wnt-3a-mediated signal to bcatenin, even when endocytosis processes are functional. This suggests that Dvl itself and not clathrin-mediated endocytosis per se, is critical for transduction of the Wnt-induced signal to b-catenin. Discussion Endocytosis of receptors has been described both as a means of terminating signalling but also of propagating intracellular signalling according to the so-called endosome signalling hypothesis (Howe & Mobley 2004, Reiter & Lefkowitz 2006). Recently, a role for clathrinmediated endocytosis has been established for the Wnt/ Frizzled signalling pathway (Chen et al. 2003, Blitzer & Nusse 2006, Marois et al. 2006, Rives et al. 2006, Seto & Bellen 2006). By studying the role of endocytosis in the b-catenin Wnt signalling, we now provide evidence for a mechanism based on the regulation of Dvl levels, which fulfils all criteria for a possible negative feedback loop in the Wnt/FZD signalling pathway. SN4741 cells endogenously express different isoforms of FZDs and an elaborate Wnt/FZD signalling machinery (G. Schulte, unpublished results, see also Schulte et al. 2005, Bryja et al. 2007a,b). In those cells, we investigated the effects of endocytosis inhibition on Dvl, which is known to be regulated by phosphorylation and polyubiquitination (Miyazaki et al. 2004, Simons et al. 2005, Angers et al. 2006, Malbon & Wang 2006). Previous experiments with the protein synthesis inhibitor cycloheximide (Miyazaki et al. 2004, Simons et al. 2005) were performed to investigate the half-life of Dvl, showing a T1/2 (Dvl) in the range of several hours (estimated 8–12 h). Our results, showing a complete degradation of Dvl after just 2 h of sucrose treatment implicate therefore an active degradation process being triggered by the block of endocytosis. FZD4-GFP, which we used to visualize the behaviour of FZDs, was localized intracellularly in the basal state. In response to hyperosmolaric sucrose or K+ depletion (It should be noted that the treatments used for blocking endocytosis might have unspecific side effects on, e.g. cytoskeleton and actin polymerisation or protein distribution in the membrane, which might be of some importance but were not studied here.), FZD4-GFP Dvl2 ABC Actin – +– + – – – + – +Sucrose Wnt-3a + + ++ – – K+ depl. Wnt-3a Dvl2 ABC Actin (a) (b) Figure 4 Inhibition of endocytosis blocks Wnt-3a-induced canonical signalling. Dvl depletion from SN4741 cells after hyperosmolaric sucrose (a) or K+ depletion (b) prevented the Wnt-3a-induced (100 ng mL)1 , 2 h) activation of b-catenin as assessed by immunoblotting using the anti-active-b-catenin antibody (ABC) recognizing the active, dephosphorylated form of b-catenin. Actin was used as a loading control. Dvl2 ABC Actin + + ++ + + + + – – – – – ––– 4 h after wash Wnt-3a 6 h after wash Sucrose Figure 5 Lack of Dvl rather than block of endocytosis per se is the reason for impairment of Wnt-3a signalling. After sucrose treatment, leading to Dvl2 downregulation, cells were washed in order to re-establish normal osmotic conditions allowing endocytosis. The figure shows the levels of Dvl2 and the activation of b-catenin by Wnt-3a (100 ng mL)1 , 4 and 6 h after washing out the sucrose) under conditions allowing endocytosis. Actin was used as a loading control. Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x 59 Acta Physiol 2007, 190, 55–61 V Bryja et al. ÆRapid Dvl degradation by endocytosis block shifted to the membrane as a result of endocytosis block. This suggests that observed effects on degradation of Dvl might depend on increased FZD surface expression and increased autocrine signalling. In fact, long-term retrovirus-mediated overexpression of Wnt- 3a in P0 neural stem cells reduced the expression of Dvl2 and Dvl3 in a similar way (D. Kunke and V. Bryja, unpublished results) supporting the suggestion that the observed mechanism can be a part of a negative feedback loop brought about by agonist stimulation. Dvl has been shown previously to undergo posttranslational modification in response to both canonical and non-canonical Wnts (for a summary, see e.g. Wharton 2003). We have recently shown that Wnt-3a and Wnt-5a activate CK1, which in turn phosphorylates Dvl (Bryja et al. 2007a,b) and contributes to the further transduction of the signal. Interestingly, treatment with the CK1 inhibitors D4476 and IC261 prevented the phosphorylation and mobility shift, but did not affect the degradation of Dvl after sucrose treatment. Therefore, we conclude that Dvl is degraded independently of its phosphorylation status on CK1-target sites. This suggests that (i) Wnt-induced posttranslational modification is not necessary for Dvl degradation induced by blocking endocytosis, and that (ii) the block of endocytosis activates (presumably by increased membranous localization of FZD and increased autocrine signalling) an unknown and CK1-independent pathway devoted to shutting down signalling by removing Dvl from the cell. Recently, Blitzer & Nusse (2006) investigated the role of endocytosis inhibition on Wnt-3a-induced canonical signalling measured both by b-catenin stabilization and TCF/Lef signalling. The authors concluded that Wntinternalization through endocytosis and the formation of clathrin-coated pits were important for the stabilization of b-catenin. Furthermore, the crucial point of cross-talk was suggested to be localized on the level of APC and GSK-3. However, with our results in hand mechanisms might become more defined. We identify Dvl downregulation as the mechanism underlying the impaired canonical signalling under endocytosis-inhibiting conditions. Moreover, we show that Wnt-signalling is abolished when the endocytotic machinery is functional but Dvl is lacking due to increased degradation. In contrast to previously published results (Blitzer & Nusse 2006), this implies that not endocytosis per se but sufficient Dvl levels are required for successful signalling to b-catenin and raises the need for the re-evaluation of previously published data. Taken together, our data and results from others shed new light on the regulatory events proximal to the membrane being crucial for the transduction of the Wnt signal as well as for the adaptation to long-term and high-dose treatment. Our results point to the cytoplasmatic protein Dvl as a possible central figure for mediating homologous and heterologous desensitization of Wnt/FZD signalling and strongly support the idea that endocytosis plays an important regulatory role in this signalling pathway. The discovery of negative feedback mechanisms for signalling pathways, such as the Wnt/Frizzled pathway, which regulates many processes during embryonic development, cancerogenesis and neurological disorders (Logan & Nusse 2004, Nusse 2005) could provide insight into the molecular basis of disease and possible novel targets for therapy. Conflict of interest The authors declare that no conflict of interest exists. We would like to thank Prof. Ernest Arenas for general support, Prof. Robert J. Lefkowitz for valuable discussion and expression vectors for FZD4-GFP, and Janet Holme´n for proofreading the manuscript. Financial support was from Karolinska Institutet, Junior Faculty, and the Foundation Lars Hiertas Minne. G.S. received postdoctoral fellowships from the Swedish Society for Medical Research (SSMF) and the Swedish Brain Foundation (Hja¨rnfonden). References Angers, S, Thorpe, C.J, Biechele, T.L. et al. 2006. The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wnt-beta-catenin pathway by targeting Dishevelled for degradation. Nature Cell Biol 8, 348–357. Blitzer, J.T. & Nusse, R. 2006. A critical role for endocytosis in Wnt signaling. BMC Cell Biol 7, 28. Bryja, V., Pachernı´k, J., Soucek, K., Horvath, V., Dvora´k, P. & Hampl, A. 2004. Increased apoptosis in differentiating p27deficient mouse embryonic stem cells. Cell Mol Life Sci 61, 1384–1400. Bryja, V., Schulte, G. & Arenas, E. 2007a. Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate b-catenin. Cell Signal 19, 610–616. Bryja, V., Schulte, G., Rawal, N., Grahn, A. & Arenas, E. 2007b. Wnt-5a induces the phosphorylation of dishevelled and the differentiation of dopaminergic neurons by a CK1dependent mechanism. J Cell Sci 120, 586–595. Chen, W., ten Berge, D., Brown, J. et al. 2003. Dishevelled 2 recruits beta-arrestin 2 to mediate Wnt5A-stimulated endocytosis of Frizzled 4. Science 301, 1391–1394. Foord, S.M., Bonner, T.I., Neubig, R.R. et al. 2005. International union of pharmacology. XLVI. G protein-coupled receptor list. Pharmacol Rev 57, 279–288. Heuser, J.E. & Anderson, R.G. 1989. Hypertonic media inhibit receptor-mediated endocytosis by blocking clathrin-coated pit formation. J Cell Biol 108, 389–400. Howe, C.L. & Mobley, W.C. 2004. Signaling endosome hypothesis: a cellular mechanism for long distance communication. J Neurobiol 58, 207–216. 60 Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x Rapid Dvl degradation by endocytosis block ÆV Bryja et al. Acta Physiol 2007, 190, 55–61 Huelsken, J. & Birchmeier, W. 2001. New aspects of Wnt signaling pathways in higher vertebrates. Curr Opin Genet Dev 11, 547–553. Kikuchi, A. & Yamamoto, H. 2007. Regulation of Wnt signaling by receptor-mediated endocytosis. J Biochem EPub ahead of print (doi: 10.1093/jb/mvm061). Larkin, J.M., Brown, M.S., Goldstein, J.L. & Anderson, R.G. 1983. Depletion of intracellular potassium arrests coated pit formation and receptor-mediated endocytosis in fibroblasts. Cell 33, 273–285. Larkin, J.M., Donzell, W.C. & Anderson, R.G. 1986. Potassium-dependent assembly of coated pits: new coated pits form as planar clathrin lattices. J Cell Biol 103, 2619–2627. Logan, C.Y. & Nusse, R. 2004. The Wnt signaling pathway in development and disease. Ann Rev Cell Dev Biol 20, 781–810. Malbon, C.C. 2004. Frizzleds: new members of the superfamily of G-protein-coupled receptors. Front Biosci 9, 1048–1058. Malbon, C.C. & Wang, H.Y. 2006. Dishevelled: a mobile scaffold catalyzing development. Curr Top Dev Biol 72, 153–166. Marois, E., Mahmoud, A. & Eaton, S. 2006. The endocytic pathway and formation of the Wingless morphogen gradient. Development 133, 307–317. Mashhoon, N., DeMaggio, A.J, Tereshko, V. et al. 2000. Crystal structure of a conformation-selective casein kinase-1 inhibitor. J Biol Chem 275, 20 052–20 060. Miyazaki, K., Fujita, T., Ozaki, T. et al. 2004. NEDL1, a novel ubiquitin-protein isopeptide ligase for dishevelled-1, targets mutant superoxide dismutase-1. J Biol Chem 279, 11 327– 11 335. Nusse, R. 2003. Wnts and Hedgehogs: lipid-modified proteins and similarities in signaling mechanisms at the cell surface. Development 130, 5297–5305. Nusse, R. 2005. Wnt signaling in disease and in development. Cell Res 15, 28–32. Reiter, E. & Lefkowitz, R.J. 2006. GRKs and beta-arrestins: roles in receptor silencing, trafficking and signaling. Trends Endocrinol Metabol 17, 159–165. Rena, G., Bain, J., Elliott, M. & Cohen, P. 2004. D4476, a cellpermeant inhibitor of CK1, suppresses the site-specific phosphorylation and nuclear exclusion of FOXO1a. EMBO Rep 5, 60–65. Rives, A.F., Rochlin, K.M., Wehrli, M., Schwartz, S.L. & DiNardo, S. 2006. Endocytic trafficking of wingless and its receptors, arrow and DFrizzled-2, in the Drosophila wing. Dev Biol 293, 268–283. Schulte, G., Bryja, V., Rawal, N., Castelo-Branco, G., Sousa, K.M. & Arenas, E. 2005. Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation. J Neurochem 92, 1550–1553. Seto, E.S. & Bellen, H.J. 2006. Internalization is required for proper wingless signaling in Drosophila melanogaster. J Cell Biol 173, 95–106. Simons, M., Gloy, J., Ganner, A. et al. 2005. Inversin, the gene product mutated in nephronophthisis type II, functions as a molecular switch between Wnt signaling pathways. Nat Genet 37, 537–543. Son, J.H., Chun, H.S., Joh, T.H., Cho, S., Conti, B. & Lee, J.W. 1999. Neuroprotection and neuronal differentiation studies using substantia nigra dopaminergic cells derived from transgenic mouse embryos. J Neurosci 19, 10–20. Wharton, K.A. 2003. Runnin’ with the Dvl: proteins that associate with Dsh/Dvl and their significance to Wnt signal transduction. Dev Biol 253, 1–17. Yamamoto, H., Komekado, H. & Kikuchi, A. 2006. Caveolin is necessary for Wnt-3a-dependent internalization of LRP6 and accumulation of beta-catenin. Dev Cell 11, 213–223. Ó 2007 The Authors Journal compilation Ó 2007 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2007.01688.x 61 Acta Physiol 2007, 190, 55–61 V Bryja et al. ÆRapid Dvl degradation by endocytosis block       Vítězslav Bryja, 2014    Attachments      #5      Bryja  V,  Gradl  D,  Schambony  A,  Arenas  E*,  Schulte  G*  (2007):  ‐arrestin  is  a  necessary  component of Wnt/‐catenin signaling in vitro and in vivo. Proc. Natl. Acad. Sci. USA 104:  6690‐6695.   * corresponding authors      Impact factor (2007): 9.380  Times cited (without autocitations, WoS, Feb 21st 2014): 45  Significance: Identification of β‐arrestin, a Dishevelled binding partner implicated in the  signaling  via  seven‐transmembrane  receptor,  as  a  crucial  component  of  the  Wnt/β‐catenin  pathway.  Positioning  of  β‐arrestin  into  canonical  Wnt  pathway  pointed to the fact that Fzd receptors probably act upstream of Dishevelled and  cooperate with trimeric G proteins.  Contibution of the author/author´s team: All in vitro work.                ␤-Arrestin is a necessary component of Wnt/␤-catenin signaling in vitro and in vivo Vı´teˇzslav Bryja*† , Dietmar Gradl‡ , Alexandra Schambony‡ , Ernest Arenas*§ , and Gunnar Schulte¶ʈ *Laboratory of Molecular Neurobiology, Department of Medical Biochemistry and Biophysics, and ¶Section of Receptor Biology and Signaling, Department of Physiology and Pharmacology, Karolinska Institutet, S-171 77 Stockholm, Sweden; and ‡Zoologisches Institut II, Universita¨t Karlsruhe (TH), Kaiserstrasse 12, D-76131 Karlsruhe, Germany Edited by Robert J. Lefkowitz, Duke University Medical Center, Durham, NC, and approved March 1, 2007 (received for review December 20, 2006) The Wnt/␤-catenin signaling pathway is crucial for proper embryonic development and tissue homeostasis. The phosphoprotein dishevelled (Dvl) is an integral part of Wnt signaling and has recently been shown to interact with the multifunctional scaffolding protein ␤-arrestin. Using Dvl deletion constructs, we found that ␤-arrestin binds a region N-terminal of the PDZ domain of Dvl, which contains casein kinase 1 (CK1) phosphorylation sites. Inhibition of Wnt signaling by CK1 inhibitors reduced the binding of ␤-arrestin to Dvl. Moreover, mouse embryonic fibroblasts lacking ␤-arrestins were able to phosphorylate LRP6 in response to Wnt-3a but decreased the activation of Dvl and blocked ␤-catenin signaling. In addition, we found that ␤-arrestin can bind axin and forms a trimeric complex with axin and Dvl. Furthermore, treatment of Xenopus laevis embryos with ␤-arrestin morpholinos reduced the activation of endogenous ␤-catenin, decreased the expression of the ␤-catenin target gene, Xnr3, and blocked axis duplication induced by X-Wnt-8, CK1␧, or Dsh⌬DEP, but not by ␤-catenin. Thus, our results identify ␤-arrestin as a necessary component for Wnt/ ␤-catenin signaling, linking Dvl and axin, and open a vast array of signaling avenues and possibilities for cross-talk with other ␤arrestin-dependent signaling pathways. canonical Wnt signaling ͉ dishevelled ͉ Frizzled ͉ G protein-coupled receptor ͉ Xenopus The Wnt/Frizzled pathway is a crucial element in cellular communication (1), which is highly conserved through evolution (2) and is required for both embryonic development and tissue homeostasis. Dysfunction and deregulation of this pathway cause different diseases, including cancer and developmental defects (3). Wnts are secreted lipoglycoproteins (4) that bind to their cognate receptors of the Frizzled family (FZD1–10) as well as their coreceptors low-density lipoprotein receptor-related protein 5/6 (LRP5/6) (1). The phosphoprotein dishevelled (Dvl) is one of the most upstream modules in this pathway (5). Dvl associates with many different intracellular proteins and is phosphorylated by different kinases (for reviews, see e.g., refs. 5 and 6). In response to Wnt ligand, Dvl is phosphorylated and activated by kinases [as casein kinase 1 (CK1)␦/␧; ref. 7], resulting in a electrophoretic mobility shift (7, 8), hereafter referred to as PS-Dvl (phosphorylated and shifted Dvl). In the Wnt/␤-catenin** signaling pathway, Dvl activation is followed by an inhibition of a destruction complex composed of axin, glycogen synthase kinase 3 and adenomatous polyposis coli, which results in a stabilization of ␤-catenin, which is now capable of regulating transcription by means of T cell factor (TCF)/lymphoid enhancer factor (Lef) transcription factors (1). Recently, the multifunctional scaffolding protein ␤-arrestin was shown to interact with Dvl (9, 10). ␤-Arrestins are known to regulate G protein-coupled receptor desensitization and internalization as well as signaling (11). Initially, ␤-arrestin-mediated endocytosis was seen as a means of desensitizing a receptor system. However, recent findings point also to a role of ␤-arrestins as a signaling scaffold for diverse signaling modules, including kinases, phosphatases, small GTPases, phosphodiesterases, ubiquitin E3 ligases, and I␬B␣ (for review, see refs. 12 and 13). With regard to Wnt signaling, ␤-arrestin1 was identified as a positive modulator of the Wnt/␤-catenin pathway and ␤-arrestin2 as a mediator for the agonist-induced internalization of FZD4 (9, 10). However, despite the information from previous overexpression studies, the mechanism of action and role of endogenous ␤-arrestin in Wnt signaling are still unknown. Here, we aimed at characterizing the nature of the Dvl–␤-arrestin interaction and its importance for Wnt/␤-catenin signaling both in vitro and in vivo. In mouse embryonic fibroblasts (MEFs) genetically depleted of ␤-arrestin1 and/or 2, we show that endogenous ␤-arrestin is necessary for the Wnt-3a-induced activation of Dvl and for signaling to ␤-catenin. We identify axin as a ␤-arrestin-binding partner, and we suggest that ␤-arrestin forms a functional, trimeric Dvl–␤-arrestin–axin complex. Moreover, loss-of-function experiments in Xenopus laevis embryos further indicated that ␤-arrestin is necessary during embryonic development for Wnt/␤-catenin signaling in vivo. Results Based on previous findings on the interaction between ␤-arrestin and Dvl (9, 10), we aimed at characterizing that interaction in more detail. We first used as a model SN4741 cells (14), a dopaminergic cell line in which Wnt signaling has been characterized (7, 15, 16). We found that immunoprecipitation of either Myc-Dvl2 or HA-␤arrestin2, in cells ectopically expressing Myc-Dvl2 and FLAG-␤arrestin2 (Fig. 1A) or FLAG-Dvl3 and HA-␤-arrestin2 (Fig. 1B), resulted in the pulldown of FLAG-␤-arrestin2 and FLAG-Dvl3, respectively. Furthermore, expression of ␤-arrestin2-GFP and MycDvl2 in SN4741 cells showed a strong colocalization of the two proteins in characteristic punctate Dvl aggregates (17, 18), as detected by immunocytochemistry and confocal laser scanning microscopy (Fig. 1C). Similar results have been obtained with Myc-Dvl2 and HA-␤-arrestin2 or FLAG-␤-arrestin2 (data not Author contributions: V.B., D.G., A.S., E.A., and G.S. designed research; V.B., D.G., A.S., and G.S. performed research; V.B., D.G., A.S., and G.S. analyzed data; and V.B., E.A., and G.S. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Abbreviations: CK1, casein kinase 1; D4476, 4-[4-(2,3-dihydro-benzo[1,4]dioxin-6-yl)-5pyridin-2-yl-1H-imidazol-2-yl]benzamide; ⌬axin, aa1–477 axin; Dsh, Drosophila melanogaster dishevelled; Dvl, mammalian dishevelled; FZD, Frizzled; KO, knockout; LEF, lymphoid enhancer factor; MEF, mouse embryonic fibroblast lacking ␤-arrestin1 (␤-arr1KO), ␤-arrestin2 (␤-arr2KO), or both (␤-arr1/2dKO); MO, morpholino; PS-Dvl, phosphorylated and shifted Dvl; TCF, T cell factor; XDsh, Xenopus dishevelled. †Present address: Institute of Experimental Biology, Faculty of Science, Masaryk University and Laboratory of Cytogenetics, Institute of Biophysics, Academy of Sciences of the Czech Republic, 601 77 Brno, Czech Republic. §To whom correspondence may be addressed. E-mail: ernest.arenas@ki.se. ʈTo whom correspondence may be addressed. E-mail: gunnar.schulte@ki.se. **Synonymous to the term ‘‘canonical signaling,’’ we rather use ‘‘Wnt/␤-catenin signaling.’’ This article contains supporting information online at www.pnas.org/cgi/content/full/ 0611356104/DC1. © 2007 by The National Academy of Sciences of the USA 6690–6695 ͉ PNAS ͉ April 17, 2007 ͉ vol. 104 ͉ no. 16 www.pnas.org͞cgi͞doi͞10.1073͞pnas.0611356104 shown), to ensure that the association of ␤-arrestin with Dvl represents a true colocalization. To narrow the interaction interface of ␤-arrestin2 with Dvl, we used a set of deletion mutants of FLAG-Dvl3 (19) (Fig. 2A) and FLAG-␤-arrestin2 (Fig. 2B) expressed in COS-7 cells. Coimmunoprecipitation of HA-␤-arrestin2 in cells overexpressing also FLAG-Dvl3 mutants revealed that HA-␤-arrestin2 can interact with FLAG-Dvl3 mutants 1, 2, 5, 6, and 7 but not with 3 and 4 [supporting information (SI) Fig. 6A], which lack a region Nterminal of the PDZ domain of Dvl (Fig. 2A). Immunoprecipitation of Myc-Dvl2 resulted in the pulldown of full-length FLAG-␤arrestin2 (construct 1) as well as the deletion mutants 2, 3, 5 but not 4 (SI Fig. 6B). To confirm the weak interaction between ␤-arrestin2 construct 2 and Dvl2-Myc (SI Fig. 6B), we performed coimmunoprecipitation of construct 2 with XDsh-Myc (20) and found a strong interaction between them (SI Fig. 6C). Thus, our results identify the region aa163–300 as the Dvl-interacting domain on ␤-arrestin2. The small portion of Dvl that was found to bind ␤-arrestin, N-terminal to the PDZ domain, was previously shown to contain Ser/Thr residues recognized by CK1␧ (21) and PAR-1 (partitioning defective mutant in C. elegans) (20, 22). Moreover, endogenous CK1␦/␧ is induced by Wnt ligands (16, 23), phosphorylates Dvl, and results in formation of active PS-Dvl (7, 16). To assess the importance of CK1-mediated phosphorylation in the Dvl–␤-arrestin interaction, we treated cells with the CK1-specific inhibitor D4476 (24), which interferes with the activation of Dvl. In cells overexpressing Myc-Dvl2, treatment with D4476 reduced the formation of PS-Myc-Dvl2 (16) and reduced the interaction of Myc-Dvl2 with HA-␤-arrestin2 (Fig. 2C), suggesting that phosphorylation of Dvl by CK1 increases the affinity for ␤-arrestin. We have shown previously (16) that expression of CK1␧ in SN4741 cells promotes an even cytosolic distribution of Myc-Dvl2, whereas a kinase-dead mutant, CK1␧K3R, promotes the typical punctate Dvl appearance in the cytosol. In triple-expression experiments using HA-␤arrestin2, Myc-Dvl2, and CK1␧ or CK1␧K3R (Fig. 2D), we found that HA-␤-arrestin2 follows the localization of CK1␧–Dvl. In the presence of CK1␧, HA-␤-arrestin2, Myc-Dvl2, and CK1␧ were distributed evenly throughout the cytoplasm, whereas in cells expressing the kinase-dead mutant of CK1␧, the proteins adopted a punctate distribution. This finding demonstrated that ␤-arrestin overexpression does not affect the CK1␧-induced translocation of Myc-Dvl2 and suggests that ␤arrestin probably acts downstream from Dvl, phosphorylated by CK1. This finding also strengthens the biochemical data of ␤-arrestin interaction with both Dvl and PS-Dvl. To examine whether endogenous ␤-arrestin is required for Wnt-3a signaling, we examined whether MEFs lacking (knockout, KO) ␤-arrestin1, ␤-arrestin2, or both (␤-arr1KO, ␤-arr2KO, ␤-arr1/ 2dKO) respond to Wnt-3a stimulation (Fig. 3A and SI Figs. 7 and 8). The wild-type (WT) MEFs showed time courses and dose responses of Dvl activation similar to those reported previously for SN4741 cells (7). However, the ability of ␤-arr2KO or ␤-arr1/2dKO MEFs to induce the formation of PS-Dvl in response to Wnt-3a was delayed and severely reduced (Fig. 3 A and B and SI Fig. 7), suggesting an important role of ␤-arrestin in Dvl activation and PS-Dvl formation. Interestingly, complete ablation of ␤-arrestins resulted in a higher basal level of PS-Dvl compared with the WT, ␤-arr1KO, and ␤-arr2KO, suggesting up-regulation of yet unknown compensatory mechanisms. Moreover, endogenous ␤-arrestin2 seemed to be more important compared with ␤-arrestin1 because the defects on PS-Dvl formation were more pronounced in MEFs lacking ␤-arrestin2 compared with cells lacking ␤-arrestin1 (SI Fig. 8). Thus, our results suggest that ␤-arrestins are crucial components of the molecular machinery mediating Dvl activation and formation of PS-Dvl in response to Wnt-3a. Because the formation of PS-Dvl at the 2-h time point studied here is necessary for the substantial activation of ␤-catenin in response to Wnt-3a (7), we also investigated dephosphorylation and stabilization of ␤-catenin. Moreover, we also analyzed the activation and phosphorylation of Ser-1490 in LRP6, using a phosphospecific antibody (25). Notably, LRP6 phosphorylation was strongly induced by Wnt-3a treatment but was not affected by genetic deletion of arrestins (Fig. 3C). In contrast, agonist-induced ␤-catenin signaling in ␤-arr2KO and ␤-arr1/2dKO MEFs was almost completely abolished (Fig. 3C). ␤-arr1KO MEFs showed a pattern of ␤-catenin dephosphorylation similar to that of the Wnt-3a-stimulated WT MEFs (SI Fig. 8). Because dephosphorylated ␤-catenin is stabilized and leads to activation of transcription in a TCF/Lef-dependent manner, we also investigated the role of ␤-arrestin in Wnt-3a-induced activation of the TOPflash luciferase reporter. Although WT MEFs strongly responded to Wnt-3a stimulation with TCF/Lef reporter activity, the lack of both ␤-arrestins completely inhibited this response (Fig. 3D). MEFs lacking ␤-arrestin2 showed a partially reduced response to Wnt-3a stimulation, supporting the idea of redundancy between ␤-arrestin1 and ␤-arrestin2 in Wnt signaling. Again, depletion of ␤-arrestins 1 and 2 increased basal TCF/Lef signaling slightly, indicating the existence of compensatory mechanisms. Thus, our in vitro data argue for an important function of ␤-arrestin in Wnt signaling, in particular for the Wnt-induced formation of PS-Dvl and subsequent activation of the Wnt/␤-catenin pathway. In addition, we sought to characterize further the potential ␤-arrestin-dependent mechanisms upstream and downstream from Dvl. First, we investigated the role of ␤-arrestin in the FZD-induced redistribution of Dvl to the plasma membrane (26). Overexpression of FZD4-GFP together with Myc-Dvl2 in both WT and ␤-arr1/ 2dKO MEFs, resulted in complete redistribution of Myc-Dvl2 from cytoplasmic punctae to the plasma membrane (SI Fig. 9). Thus, ␤-arrestin is not crucial for the FZD4-induced Dvl2 translocation. Second, we analyzed the interaction of axin with ␤-arrestin by overexpression in COS-7 cells. As shown in Fig. 4A, a full-length axin or a deletion mutant 1–477 axin (⌬axin) lacking the DIX domain (27), which is necessary for the interaction with Dvl (28), coimmunoprecipitated with ␤-arrestin. Thus, ␤-arrestin can interact directly with axin, and such interaction is most likely not mediated by Dvl. These data suggest that a trimeric Dvl–␤-arrestin– axin complex might be central for the recruitment of Dvl to axin and the subsequent dissociation of the destruction complex (27–29). To test this possibility, we transfected in Fig. 4B the indicated combinations of vectors encoding for Dvl (XDsh), ␤-arrestin, and full- A Dvl2-mycarr2-GFP merge C Myc-Dvl2 Flag- -arr2 - + WB: Flag Flag WB: Myc IP:Myc-Dvl2 TCL: + + - + B TCL: HA- arr2 Flag-Dvl3 - + WB: Flag Flag WB: HA IP:HA-arr2 + + - + Fig. 1. Interaction and colocalization of Dvl and ␤-arrestin2. Expression of Myc-Dvl2 and FLAG-␤-arrestin2 (A) or HA-␤-arrestin2 and FLAG-Dvl3 (B) in SN4741 cells allowed coimmunoprecipitation of Myc-Dvl2 together with FLAG-␤-arrestin2 (A) or of HA-␤-arrestin2 with FLAG-Dvl3 (B). Total cell lysates (TCL) analyzed for the presence of transfected FLAG-tagged proteins by immunoblotting. (C) SN4741 cells transfected with ␤-arrestin2-GFP (␤arr2GFP) and Myc-Dvl2 show substantial colocalization (indicated in yellow in merged picture) of those proteins in cytoplasmatic, nonvesicular punctae, which are the typical subcellular distribution of Dvl as analyzed by confocal microscopy. WB, Western blotting. Bryja et al. PNAS ͉ April 17, 2007 ͉ vol. 104 ͉ no. 16 ͉ 6691 CELLBIOLOGY length/⌬axin for coimmunoprecipitation. We found that ␤-arrestin, XDsh, and axin are precipitated together and are, thus, likely to form a trimeric complex. Interestingly, ⌬axin is also present in such complexes, suggesting that it might be recruited to XDsh by ␤-arrestin. However, ␤-arrestin is not necessary for the XDsh–axin interaction, at least not in an overexpression system, because overexpressed XDsh interacts to a similar extent with axin in both WT and ␤-arr1/2dKO MEFs (data not shown). In addition it appears that axin could act as a stabilizing component of XDsh/ ␤-arrestin binding because the association of XDsh and ␤-arrestin is stronger when axin is coexpressed (compare lanes 3 and 4 with lane 5, WB: XDsh, Fig. 4B). D A B + 1 163 300 410 1 2 3 4 5 FLAG FLAG FLAG FLAG FLAG + + + - + DEPPDZDIXFLAG DEPPDZFLAG DEPFLAG FLAG DIXFLAG PDZDIXFLAG DEPPDZDIXFLAG 1 2 3 4 5 6 7 + + - - + + 1 71682 333 492 CKI HA- arr2 Dvl2-Myc merge CKI K R HA- arr2 Dvl2-Myc merge C WB: Myc WB: HA WB: HA IPMyc Dvl2-Myc HA- -arrestin D4476 - - - - + + - + + + + 1.47 0.57 1.70 0.05 Fig. 2. Mapping of the interaction interfaces of ␤-arrestin2 and Dvl. Deletion mutants of FLAG-Dvl3 (A; constructs 1–7; ref. 19) and of FLAG-␤-arrestin2 were used in combination with HA-␤-arrestin2 and Myc-Dvl2, respectively, for coim- munoprecipitation.ϩandϪindicateinteractionwith␤-arrestin(A)orDvl(B).For detailed presentation of coimmunoprecipitation data, see SI Fig. 6. (C) ␤-Arrestin2 preferentially binds phosphorylated Dvl2. SN4741 cells expressing Dvl2-Myc and HA-␤-arrestin2 were grown with and without the CK1 inhibitor D4476. D4476 treatment prevents formation of PS-Dvl and decreases the amount of HA-␤-arrestin2 that can be precipitated with Myc-Dvl2, suggesting a decreased affinity of ␤-arrestin to unphosphorylated Dvl. Numbers represent OD of ␤-arrestin signal and PS-Dvl/Dvl ratio. The arrowhead points to the localization of PS-Dvl band. (D) Overexpressed HA-␤-arrestin2 (red; HA-␤-arr2) and Dvl2-Myc (green) as well as CK1␧ (b/w; kinase-dead mutant CK1␧ K3R) are analyzed by wt -arr2KO -arr1/2dKO 0 1 2 3 4 control Wnt-3a (100 ng/ml) * Relativeluciferaseactivity Dvl2 A B D Wnt-3a wt arr2KO arr1/2dKO 0 5 30 60 12015 min Wnt-3a 0 3 30 10010 ng/ml MEFs: C wt arr2KO arr1/2dKO Dvl2 MEFs: Wnt-3a - + ABC actin - + - + P-LRP6 wtMEFs: -catenin arr2 KO arr1/2 dKO Fig. 3. ␤-Arrestin2 is a necessary component of canonical Wnt signaling in MEFs. (A) WT, ␤-arr2KO, or ␤-arr1/2dKO MEFs are treated with 100 ng/ml Wnt-3a for the indicated times. Dvl2 activation is assessed as the formation of PS-Dvl2 (Dvl, open; PS-Dvl, filled arrowheads) by immunoblotting. (B) Wnt-3a dose responses were performed at 2 h. (C) Quantification of data is available in SI Fig. 7. Phosphorylation of LRP6 at Ser-1490 (P-LRP6) and ␤-catenin signaling was analyzed in WT, ␤-arr2KO, and ␤-arr1/2dKO MEFs after 100 ng/ml Wnt-3a treatment for 2 h. Dephosphorylation of ␤-catenin (ABC) and ␤-catenin stabilization show that Wnt-3a-induced signaling to ␤-catenin is reduced in ␤-arr2KO MEFs and completely abolished in ␤-arr1/2dKO MEFs. For responses in ␤-arr1KO MEFs, see SI Fig. 8. (D) SuperTOPflash activity upon 100 ng/ml Wnt-3a stimulation is shown (n ϭ 3). *, P Ͻ 0.01 as analyzed by ANOVA and TukeyЈs post hoc tests. confocallaserscanningmicroscopy.InthepresenceofWTCK1␧,␤-arrestin2,Dvl2, and CK1 are distributed evenly throughout the cytoplasm. In the presence of CK1␧K3R, ␤-arrestin2, Dvl2, and CK1␧ are confined to nonvesicular punctae characteristic for Dvl aggregates. Overlap of distribution of ␤-arrestin and Dvl2Myc is indicated in yellow in the merged pictures. 6692 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0611356104 Bryja et al. To examine the importance of ␤-arrestin for Wnt signaling in vivo, during embryonic development, we turned to one of the most classical and robust models to study Wnt signaling, the early X. laevis embryo. The axis duplication assay, as induced by injection of XWnt-8 or Dsh⌬DEP mRNA in the ventral blastomeres, was used to examine the role of ␤-arrestin overexpression in canonical signaling. The lack of effect of ␤-arrestin RNA injection on axis duplication under basal conditions or in blastomeres injected with XWnt-8 or Dsh⌬DEP suggested that ␤-arrestin per se is not sufficient to induce Wnt/␤-catenin signaling. In contrast, down-regulation of Xenopus ␤-arrestins (for sequence and alignment, see SI Fig. 10) with ␤-arrestin morpholinos (␤-arrMO) showed that ␤-arrestin is required for Wnt/␤-catenin pathway (Fig. 5 B and C) because the XWnt-8induced axis duplication was blocked (percent axis duplication: XWnt-8, 51.0; XWnt-8 ϩ ␤-arrMO, 8.5). Control injections with 0 10 20 30 40 50 60 XWnt-8 Dsh DEP -arrestin - + - ++ n= 202 55 189 113122 %secondaryaxis 0 10 20 30 40 50 XWnt-8 Dsh DEP -arrMO - + - ++ n= -catenin - 202 258189 5758 45 CK1 70110 - + %secondaryaxis A D C +-arrMO--arrMO E A1CT ABC --arrMO + 0.0 0.5 1.0 control -arrMO relative expressionof Xnr3 B Fig. 5. ␤-Arrestin is not sufficient but necessary for Wnt/␤-catenin signaling in vivo. (A) Percentage of secondary axis induction in Xenopus embryos injected with XWnt-8, Dsh⌬DEP, and ␤-arrestin RNA. n, no. of analyzed embryos per condition. (B) Effect of ␤-arrestin morpholino (␤-arrMO) treatment on XWnt-8-, Dsh⌬DEP-, CK1␧-, and ␤-catenin-induced axis duplication; n, no. of analyzed embryos per condition. (C) Representative photographs of embryos. Secondary axis formation (arrowheads) is best seen by the appearance of a secondary neural tube in neurula (XWnt-8, ␤-catenin), tailbud (CK1␧), and tadpole stages (Dsh⌬DEP). Arrows indicate the endogenous axes. (D) Immunoblotting of Xenopus lysates analyzed for ␤-catenin dephosphorylation (ABC) and ␤-arrestin levels (A1CT). Filled arrowheads indicate Xenopus ␤-arrestin signal. Open arrowhead indicates an unspecific band serving as loading control. Gene expression (Xnr3) was examined in Xenopus embryos by quantitative RT-PCR. (E) Error bars show mean Ϯ SD. - + - + - + + + + - + + - - + + - - - - -arr2 axin axin XDsh - + - + - + + + + - + + - - + + - - - - - + - + - + + + + - + + - - + + - - - - B WB: axin WB: XDsh WB: -arr WB: axin WB: -arr -arr2 axin axin - - - + - + + + - + + - - - + - + + + - + + A IP: -arrTCL axinIP: -arr XDsh Fig. 4. ␤-Arrestin interaction with axin. (A) Immunoprecipitation (IP) experiments in COS-7 cells show the interaction of ␤-arrestin (Right) with full-length and aa1–477 axin (⌬axin). (Left) Total cell lysates (TCL). The arrowhead points to the IgG band in the immunoprecipitate. WB, Western blot. (B) Overexpression of HA-␤-arrestin2, Myc-Dishevelled (Xenopus dishevelled, XDsh) and full length (FLAG-tagged) as well as ⌬axin reveals the existence of trimeric Dvl-␤–arrestin–axin complexes. Bryja et al. PNAS ͉ April 17, 2007 ͉ vol. 104 ͉ no. 16 ͉ 6693 CELLBIOLOGY ␤-arrMO alone did not result in axis duplication (data not shown; percent axis duplication: ␤-arrMO, 0, n ϭ 261). In addition, ␤-arrMO (Fig. 5 B and C) completely blocked Dsh⌬DEP Ϫ (percent axis duplication: Dsh⌬DEP, 21.2; Dsh⌬DEP ϩ ␤-arrMO, 0) as well as CK1␧-induced secondary axis formation (percent axis duplication: CK1␧, 13.6; CK1␧ ϩ ␤-arrMO, 1.4). However, ␤-arrMO did not affect secondary axis induced by ␤-catenin (% axis duplication: ␤-catenin, 44.4; ␤-catenin ϩ ␤-arrMO, 38.6), arguing for ␤-arrestin acting upstream from ␤-catenin and downstream from CK1␧ and Dvl, which is in good agreement with our in vitro data. As control for the CK1␧ injection we also injected a kinase-dead mutant of CK1␧K3R, which alone or in combination with ␤-arrMO failed to induce axis duplication (data not shown; % axis duplication, CK1␧, 0, n ϭ 87; CK1␧ ϩ ␤-arrMO, 0, n ϭ 117). Detection of activated, dephosphorylated ␤-catenin (ABC, Fig. 5D) and expression of the ␤-catenin target gene Xnr3 (Fig. 5E) in control and ␤-arrestin depleted Xenopus embryos further showed that ␤-catenin activity is reduced in ␤-arrestin-depleted embryos. In summary, the Xenopus experiments demonstrate that endogenous ␤-arrestin is required for normal development and for Wnt/␤-catenin signaling in vivo. Discussion Here, we identify the multipurpose scaffolding protein ␤-arrestin as a necessary component in the Wnt/␤-catenin pathway in vitro and in vivo. We show that ␤-arrestin binds to and forms a trimeric complex with axin and Dvl. Moreover, endogenous ␤-arrestin does not affect phosphorylation of LRP6 but is necessary for proper activation of Dvl, signaling along the Wnt/␤-catenin pathway, activation of endogenous Wnt target genes, and XWnt- 8-, Dsh⌬DEP-, or CK1⑀-induced axis duplication in Xenopus embryos in vivo. We mapped the domain of physical interaction between ␤-arrestin and Dvl (9, 10) to a region N-terminal of the PDZ domain of Dvl and aa163–300 of ␤-arrestin. This stretch of Dvl contains important CK1␧ (21), PAR-1 (20, 22), and possibly also casein kinase 2 (CK2) (30) phosphorylation sites, which may work as potential regulators of ␤-arrestin–Dvl interaction. Importantly, CK1, CK2, and PAR-1 are all necessary for Wnt/␤catenin signaling (20, 31, 32), and at least CK1␧ and CK2 kinase activity is induced directly by Wnts (16, 23, 33). It has been reported that the association of Dvl with ␤-arrestin is potentiated by CK2 phosphorylation (9), and we report herein that this interaction is reduced by CK1 inhibition, a finding that also supports this hypothesis. We also found that cells lacking ␤-arrestin resemble cells with CK1␦/␧ knockdown (7, 16) in their inability to form PS-Dvl properly in response to Wnt. We thus suggest that the formation of the activated form of Dvl, PS-Dvl, is initiated by and depends on Dvl phosphorylation by Wntinduced kinases, as we demonstrated previously for CK1␧ (7, 16). In this model, ␤-arrestin recognizes phosphorylated (primed) Dvl and subsequently recruits other kinases to complete PS-Dvl formation and Dvl activation. A second function of ␤-arrestin that is likely to be crucial for Wnt signaling is its capacity to bind axin directly. This binding did not require the DIX domain of axin, which is necessary for the interaction of Dvl (7). This finding suggested the possibility of a trimeric complex consisting of Dvl, ␤-arrestin, and axin, which was confirmed by coimmunoprecipitation experiments. We hypothesize that such a trimeric complex could contribute to destabilize the degradation complex, resulting in the stabilization of ␤-catenin. Thus, our results suggest that the formation of a trimeric complex is an essential step in Wnt signaling that provides a molecular link between the LRP6/axin and FZD/Dvl branches of Wnt signaling after activation. Experiments in ␤-arrestin-deficient MEFs suggest that ␤-arrestins 1 and 2 might be at least partially redundant with respect to mediating Wnt signaling and activation. Interestingly, unstimulated ␤-arr1/2KO MEFs contain a higher amount of PS-Dvl and higher background activity in the TOPflash reporter assay compared with WT and ␤-arr1KO or ␤-arr2KO MEFs, indicating that mechanisms exist to accomplish the formation of PS-Dvl in the absence of ␤-arrestin. These compensatory mechanisms become evident also regarding the obvious delay and decrease in efficacy of Wnt-3a-induced PS-Dvl formation rather than a complete block. It is possible that high concentrations of Wnt or long-term exposure may to some extent lead to PS-Dvl formation in MEFs deficient in ␤-arrestins, although not to the same extent as in WT MEFs. These findings show that ␤-arrestin is not indispensable for PS-Dvl formation and suggest that redundant mechanism(s) exist, which is also evident from the quantification of PS-Dvl/Dvl data (SI Fig. 7), indicating that absence of ␤-arrestin2 alone has a more pronounced effect on Dvl activation compared with cells depleted of both ␤-arrestins. However, ␤-arrestin seems to be uniquely required for proper dynamics of the PS-Dvl formation and for correct Wnt signaling in vivo, as demonstrated by the dramatic defects in ␤-arrestin-depleted Xenopus embryos. It should be noted, however, that although ␤-arrestin is a necessary component for Wnt signaling, it is not sufficient to induce PS-Dvl formation (Fig. 1A) or axis duplication in Xenopus embryos (Fig. 5). Thus, our data suggest that ␤-arrestin is required, but not sufficient, for Wnt signaling and that ␤-arrestin is recruited to Wnt signaling upon Wnt activation. Our findings regarding the necessity of ␤-arrestin for Wnt signaling, together with the recognized function of ␤-arrestins as multifunctional adapter proteins, also open other interesting possibilities such as that ␤-arrestin might be necessary for other signaling pathways or that ␤-arrestin might mediate an interaction between Wnt and other signaling pathways (12). This assumption is further supported by structural data indicating that the domain of ␤-arrestin2 that binds Dvl (aa 163–300) lies in a region of the protein containing large parts of the C-terminal ␤-fold (␤ sheet X-XVI; see alignments of nonvisual vertebrate ␤-arrestins in SI Fig. 10). Thus, both the N and C termini of ␤-arrestin could remain open for interaction with other partners after Dvl binding. ␤-Arrestins are also known to play a very important role in clathrin-mediated endocytosis (34). In addition, blockade of endocytosis by, for example, hyperosmolar sucrose blocks Wnt-induced ␤-catenin signaling (35) and down-regulates Dvl (36). Combined, these findings together with the recently appreciated role of endocytosis in Wnt/Frizzled signaling (10, 35, 37–40) suggest the provocative possibility that ␤-arrestin-mediated endocytosis might be required for some aspects of Wnt signal transduction, an issue that remains to be explored. In summary, we hereby identify ␤-arrestin as an integral component of the Wnt signaling pathway. This finding opens up the possibility that Wnts may signal through novel additional pathways, allowing extensive cross-talk with other, e.g., G protein-coupled receptor-dependent, pathways. Future studies will explore such interactions and will focus on further elucidating the role of ␤-arrestin in Dvl activation and the function of the trimeric complex formed by Dvl–␤-arrestin–axin in Wnt signaling. Methods Cell Culture, Transfection, and Treatments. SN4741 cells were obtained from J. H. Son (14). WT MEFs and MEFs lacking ␤-arrestin1 (␤-arr1KO), ␤-arrestin2 (␤-arr2KO), or ␤-arrestins 1 and 2 (␤-arr1/2dKO) were a gift from R. J. Lefkowitz (41). SN4741 cells were propagated in DMEM/10% FCS/2 mM L-glutamine/50 units/ml penicillin/50 units/ml streptomycin/ 0.6% glucose. MEFs and COS-7 cells were grown in identical medium without glucose. Cells (40,000–60.000 per well) were seeded in 24-well plates either directly (for biochemical analysis) or on sterile coverslips (for microscopy). The next day, cells were transfected (plasmids are listed in SI Materials and Methods) by 6694 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0611356104 Bryja et al. using Lipofectamine 2000 or polyethylenimine at 0.8 ␮g/ml in PBS (42). For confocal analysis, 0.2 ␮g per construct were used according to the manufacturer’s instructions. Cells were harvested for immunoblotting or immunocytochemistry 24 h after transfection (for a detailed description of immunocytochemistry and confocal analysis, see SI Materials and Methods). Treatment with the D4476 (4-[4-(2,3-dihydro-benzo[1,4]dioxin-6-yl)-5pyridin-2-yl-1H-imidazol-2-yl]benzamide), dissolved in DMSO, was done in 24-well plates in the presence of 1 ␮l per well FuGENE 6 reagent to increase cell penetration. For analysis of cellular signaling, the cells were stimulated with mouse Wnt-3a (from R&D Systems, Minneapolis, MN) for 2 h if not otherwise stated. Control stimulations were done with 0.1% BSA in PBS. For the TOPflash assay, see SI Materials and Methods. Immunoprecipitation and Immunoblotting. Immunoprecipitation was done as described previously (16). For domain-mapping experiments, cells were lysed in 0.1% SDS instead of 0.5% Nonidet P-40. Immunoblotting and sample preparation were done as published previously (7). Protein extracts from X. laevis embryos were obtained by using 20 ␮l of lysis buffer (150 mM NaCl/1 mM EDTA/50 mM Tris⅐Cl, pH 7.4/0.5% Nonidet P-40) supplemented with protease inhibitors per embryo. After sonication, lysates were depleted of yolk and lipids by mixing with 30 ␮l of freon per embryo, vortexing, and spinning down (15 min at 12,000 ϫ g). The upper phase was mixed with 5ϫ Laemmli buffer (4:1), boiled, and analyzed further. A list of antibodies for immunoblotting and immunoprecipitation is available in the SI Materials and Methods. Injection and Analysis of X. laevis Embryos. Capped mRNAs were transcribed from linearized DNA templates (pCS2-Dsh⌬DEP, psp64T-XWnt-8, psp64T-␤-catenin, pcDNA-HA-␤-arrestin2) by using mMessage mMachine (Ambion, Austin, TX). For knockdown experiments in Xenopus, a morpholino antisense oligonucleotide targeted against Xenopus ␤-arrestin (␤-arrMO: 5ЈTCTCCCCCATCTTCCCAGCTCCGC-3Ј) was used. Eggs from human chorionic gonadotropin-treated females were fertilized by standard methods and staged according to (43). Morpholino and RNA were injected into the marginal zone of the ventral or dorsal blastomeres of four-cell stage embryos in a total volume of 4 nl. If not mentioned otherwise in the text, the following amounts were injected: XWnt-8, 20 pg; ␤-catenin, 250 pg; Dsh⌬DEP, 500 pg; ␤-arrestin2, 500 pg; CK1␧, 500 pg; ␤-arrMO, 0.4 pg. Embryos were cultivated as described previously (44). For real-time RT-PCR total RNA was extracted from stage 10.5 embryos (Nucleospin II; Macherey Nagel, Du¨ren, Germany) and reverse transcribed using Moloney murine leukemia virus reverse transcriptase (Promega, Madison, WI). Real-time PCR was carried out using iQ-SybrGreen Supermix on an iCycler instrument (Bio-Rad, Hercules, CA). Primer sequences were: ODC-U, 5Ј-gcc att gtg aag act ctc tcc att c; ODC-D, 5Ј-ttc ggg tga ttc ctt gcc ac (45); Xnr-3 forward, 5Ј-aag aga tca aac ccg agt gc; and Xnr-3 reverse, 5Ј-ctg tgg aac tgc aca agt gg. Expression levels were calculated relative to ornithine decarboxylase and normalized to uninjected controls. We thank Dr. R. J. Lefkowitz (Duke University Medical Center) for providing the ␤-arrestin KO MEFs, ␤-arrestin, and FZD4-GFP constructs, and ␤-arrestin antibodies as well as for valuable discussions. Drs. S. Yanagawa (Kyoto University, Kyoto, Japan), R. T. Moon (University of Washington School of Medicine, Seattle, WA), O. Ossipova (Mt. Sinai School of Medicine, New York, NY), T. C. Dale (Cardiff School of Biosciences, Cardiff, U.K.), and J. M. Graff (University of Texas Southwestern Medical Center, Dallas, TX) provided additional plasmids, and we thank Dr. Son for SN4741 cells. This work was supported by the Swedish Foundation for Strategic Research (INGVAR and Center of Excellence in Developmental Biology), Swedish Research Council (VR), European Union (Eurostemcell) (to E.A.), Karolinska Institutet (to E.A. and G.S.), the Stiftelse Lars Hiertas Minne and Tore Nilsson Foundation (to G.S.), MoEYS CR Grant MSM0021622430 (to V.B.), Swedish Society for Medical Research and the Swedish Brain Foundation (Hja¨rnfonden) postdoctoral fellowships (to G.S.), and Deutsche Forschungsgemeinschaft Grants DFG GR 1802 (to D.G.) and DFG SCHA 965 (to A.S.). 1. Gordon MD, Nusse R (2006) J Biol Chem 281:22429–22433. 2. Huelsken J, Birchmeier W (2001) Curr Opin Genet Dev 11:547–553. 3. Nusse R (2005) Cell Res 15:28–32. 4. Nusse R (2003) Development (Cambridge, UK) 130:5297–5305. 5. Wharton KA (2003) Dev Biol 253:1–17. 6. Malbon CC, Wang HY (2006) Curr Top Dev Biol 72:153–166. 7. Bryja V, Schulte G, Arenas E (2006) Cell Signal 19:610–616. 8. Gonza´lez-Sancho JM, Brennan KR, Castelo-Soccio LA, Brown AM (2004) Mol Cell Biol 24:4757–4768. 9. Chen W, Hu LA, Semenov MV, Yanagawa S, Kikuchi A, Lefkowitz RJ, Miller WE (2001) Proc Natl Acad Sci USA 98:14889–14894. 10. Chen W, ten Berge D, Brown J, Ahn S, Hu LA, Miller WE, Caron MG, Barak LS, Nusse R, Lefkowitz RJ (2003) Science 301:1391–1394. 11. Reiter E, Lefkowitz RJ (2006) Trends Endocrinol Metab 17:159–165. 12. Lefkowitz RJ, Shenoy SK (2005) Science 308:512–517. 13. Gurevich EV, Gurevich VV (2006) Genome Biol 7:236. 14. Son JH, Chun HS, Joh TH, Cho S, Conti B, Lee JW (1999) J Neurosci 19:10–20. 15. Schulte G, Bryja V, Rawal N, Castelo-Branco G, Sousa KM, Arenas E (2005) J Neurochem 92:1550–1553. 16. Bryja V, Schulte G, Rawal N, Grahn A, Arenas E (2007) J Cell Sci 120:586–595. 17. Smalley MJ, Signoret N, Robertson D, Tilley A, Hann A, Ewan K, Ding Y, Paterson H, Dale TC (2005) J Cell Sci 118:5279–5289. 18. Schwarz-RomondT,MerrifieldC,NicholsBJ,BienzM(2005)JCellSci118:5269–5277. 19. Angers S, Thorpe CJ, Biechele TL, Goldenberg SJ, Zheng N, MacCoss MJ, Moon RT (2006) Nat Cell Biol 8:348–357. 20. Ossipova O, Dhawan S, Sokol S, Green JB (2005) Dev Cell 8:829–841. 21. Klein TJ, Jenny A, Djiane A, Mlodzik M (2006) Curr Biol 16:1337–1343. 22. Sun TQ, Lu B, Feng JJ, Reinhard C, Jan YN, Fantl WJ, Williams LT (2001) Nat Cell Biol 3:628–636. 23. Swiatek W, Tsai IC, Klimowski L, Pepler A, Barnette J, Yost HJ, Virshup DM (2004) J Biol Chem 279:13011–13017. 24. Rena G, Bain J, Elliott M, Cohen P (2004) EMBO Rep 5:60–65. 25. Tamai K, Zeng X, Liu C, Zhang X, Harada Y, Chang Z, He X (2004) Mol Cell 13:149–156. 26. Axelrod JD, Miller JR, Shulman JM, Moon RT, Perrimon N (1998) Genes Dev 12:2610–2622. 27. Smalley MJ, Sara E, Paterson H, Naylor S, Cook D, Jayatilake H, Fryer LG, Hutchinson L, Fry MJ, Dale TC (1999) EMBO J 18:2823–2835. 28. Kishida S, Yamamoto H, Hino S, Ikeda S, Kishida M, Kikuchi A (1999) Mol Cell Biol 19:4414–4422. 29. Li L, Yuan H, Weaver CD, Mao J, Farr GH, Sussman DJ, Jonkers J, Kimelman D, Wu D (1999) EMBO J 18:4233–4240. 30. Willert K, Brink M, Wodarz A, Varmus H, Nusse R (1997) EMBO J 16:3089–3096. 31. Dominguez I, Mizuno J, Wu H, Song DH, Symes K, Seldin DC (2004) Dev Biol 274:110–124. 32. Peters JM, McKay RM, McKay JP, Graff JM (1999) Nature 401:345–350. 33. Gao ZH, Seeling JM, Hill V, Yochum A, Virshup DM (2002) Proc Natl Acad Sci USA 99:1182–1187. 34. Goodman OB, Krupnick JG, Santini F, Gurevich VV, Penn RB, Gagnon AW, Keen JH, Benovic JL (1996) Nature 383:447–450. 35. Blitzer JT, Nusse R (2006) BMC Cell Biol 7:28. 36. Bryja V, Caja´nek L, Grahn A, Schulte G (2007) Acta Physiol Scand 190:53–59. 37. Seto ES, Bellen HJ (2006) J Cell Biol 173:95–106. 38. Rives AF, Rochlin KM, Wehrli M, Schwartz SL, DiNardo S (2006) Dev Biol 293:268–283. 39. Marois E, Mahmoud A, Eaton S (2006) Development (Cambridge, UK) 133:307–317. 40. Yamamoto H, Komekado H, Kikuchi A (2006) Dev Cell 11:213–223. 41. Kohout TA, Lin FS, Perry SJ, Conner DA, Lefkowitz RJ (2001) Proc Natl Acad Sci USA 98:1601–1606. 42. Lundgren TK, Scott RP, Smith M, Pawson T, Ernfors P (2006) J Biol Chem 281:29886–29896. 43. Nieuwkoop PD, Faber J (1967) Normal Table of Xenopus laevis (Elsevier/ North–Holland, Amsterdam). 44. Ku¨hl M, Finnemann S, Binder O, Wedlich D (1996) Mech Dev 54:71–82. 45. Yokota C, Kofron M, Zuck M, Houston DW, Isaacs H, Asashima M, Wylie CC, Heasman J (2003) Development (Cambridge, UK) 130:2199–2212. Bryja et al. PNAS ͉ April 17, 2007 ͉ vol. 104 ͉ no. 16 ͉ 6695 CELLBIOLOGY VIII human β−arrestin 1 isoform A human β−arrestin 1 isoform B human β−arrestin 2 isoform 1 human β−arrestin 2 isoform 2 mouse β−arrestin 1 isoform A mouse β−arrestin 1 isoform B mouse β−arrestin 2 Xβ−arrestin 2-like Xβ−arrestin 2-like human β−arrestin 1 isoform A human β−arrestin 1 isoform B human β−arrestin 2 isoform 1 human β−arrestin 2 isoform 2 mouse β−arrestin 1 isoform A mouse β−arrestin 1 isoform B mouse β−arrestin 2 Xβ−arrestin 2-like Xβ−arrestin 2-like human β−arrestin 1 isoform A human β−arrestin 1 isoform B human β−arrestin 2 isoform 1 human β−arrestin 2 isoform 2 mouse β−arrestin 1 isoform A mouse β−arrestin 1 isoform B mouse β−arrestin 2 Xβ−arrestin 2-like Xβ−arrestin 2-like human β−arrestin 1 isoform A human β−arrestin 1 isoform B human β−arrestin 2 isoform 1 human β−arrestin 2 isoform 2 mouse β−arrestin 1 isoform A mouse β−arrestin 1 isoform B mouse β−arrestin 2 Xβ−arrestin 2-like Xβ−arrestin 2-like high consensus – red low consensus – blue neutral color - black XX XVII XVII XIX VIIVIVIII IVII α-I α-I XVIXII XIII XIV XVIX X XI I Bryja et al. Figure S1 A B DEP Flag-Dvl3 mutants WB: HA IP:HA-β-arrestin TCL: 100 75 50 37 75 37 kDa 100 50 1 2 3 4 5 6 7 WB: Flag WB: Flag 2* WB: Flag WB: Myc WB: Flag IP:Myc-Dvl2IP:Flag-β-arrestin2 1 2 3 4 5 37 20 25 15 50 kDa Flag-β-arrestin2 mutants 37 20 25 15 50 WB: Flag WB: Myc IP:FLAG-β-arrestin2 WB: Flag WB: Myc IP:XDsh-Myc XDsh-Myc- + + + + - - - FLAG-β-arr2(163-300) XDsh-Myc- + + + + - - - FLAG-β-arr2(163-300) C Bryja et al. Figure S2 Bryja et al. Figure S3 A B C β-arr1KO MEFs Wnt-3a - + ABC β-catenin actin P-Lrp6 β-arr1KO MEFs Wnt-3a 100 ng/ml Dvl2 0 30 60 min 0 30 100 ng/ml Dvl2 Wnt-3a wt β-arr1KOβ-arr2KOβ-arr1/2dKO β-arr1 β-arr2 Dvl2-Myc -FZD4 +FZD4 MEF wt MEF β-arr1/2dKO Bryja et al. Figure S4 Bryja et al. Figure S5 0 20 40 60 80 100 120 0 20 40 60 80 100 120 MEF wt MEF β-arr2KO MEF β-arr1/2dKO time [min] ratioPS-Dvl/Dvl %ofmaxresponse 0 10 100 0 20 40 60 80 100 120 log Wnt-3a [ng/ml] ratioPS-Dvl/Dvl %ofmaxresponse * ** *** *** *** ** ***       Vítězslav Bryja, 2014    Attachments      #6      Bryja V 1 , Schambony A 1 , Čajánek L, Dominguez I, Arenas E*, Schulte G* (2008): beta‐ Arrestin and casein kinase 1/2 define distinct branches of non‐canonical WNT signalling  pathways. EMBO Rep. 9(12): 1244‐50. Featured article.  1  equal contribution, * corresponding author      Impact factor (2008): 7.099  Times cited (without autocitations, WoS, Feb 21st 2014): 15  Significance: This study defined the role of β‐arrestin in non‐canonical Wnt pathways.  We first described, both in vitro and in Xenopus system, that involvement of β‐ arrestin  can  serve  as  a  discriminant  of  the  branch  involving  activation  of  small  GTPase Rac1. In contrast, activity of casein kinase 1 (and phosphorylation of Dvl  by this kinase) prevents Rac1 activation and triggers other, not yet well defined  pathway.  Contibution  of  the  author/author´s  team:  All  in  vitro  work,  design  of  the  in  vivo  experiments.                b-Arrestin and casein kinase 1/2 define distinct branches of non-canonical WNT signalling pathways Vı´te˘zslav Bryja1,2*, Alexandra Schambony3w*, Luka´s˘Cˇaja´nek1, Isabel Dominguez4, Ernest Arenas1+ & Gunnar Schulte5++ 1Laboratory of Molecular Neurobiology, Department of Medical Biochemistry & Biophysics, Karolinska Institutet, Stockholm, Sweden,2InstituteofBiophysicsofAcademyofSciencesoftheCzechRepublic& InstituteofExperimentalBiology,FacultyofScience, Masaryk University, Brno, Czech Republic, 3Zoologisches Institut II, Universita¨t Karlsruhe (TH), Karlsruhe, Germany, 4Boston University, Boston, Massachusetts, USA, and 5Section of Receptor Biology & Signaling, Department of Physiology & Pharmacology, Karolinska Institutet, Stockholm, Sweden Recent advances in understanding b-catenin-independent WNT (non-canonical) signalling suggest an increasing complexity, raising the question of how individual non-canonical pathways are induced and regulated. Here, we examine whether intracellular signalling components such as b-arrestin (b-arr) and casein kinases 1 and 2 (CK1 and CK2) can contribute to determining signalling specificity in b-catenin-independent WNT signalling to the small GTPase RAC-1. Our findings indicate that b-arr is sufficient and required for WNT/RAC-1 signalling, and that casein kinases act as a switch that prevents the activation of RAC-1 and promotes other non-canonical WNT pathways through the phosphorylation of dishevelled (DVL, xDSH in Xenopus). Thus, our results indicate that the balance between b-arr and CK1/2 determines whether WNT/RAC-1 or other non-canonical WNT pathways are activated. Keywords: convergent extension movements; RAC-1; RHO-like GTPases; Xenopus laevis; WNT-5A EMBO reports (2008) 9, 1244–1250. doi:10.1038/embor.2008.193 INTRODUCTION WNTs are secreted glycolipoproteins that activate b-catenindependent and -independent (non-canonical) pathways. Several recent reports have provided evidence for the complexity of the non-canonical WNT pathways (Semenov et al, 2007). The individual b-catenin-independent WNT branches, which mediate planar cell polarity (for a list of abbreviations, see supplementary information online) or convergent extension movements, are so diverse that it has become necessary to name them by the signalling components involved. Here, we refer to the following pathways as defined in the original studies: WNT/RHO (Habas et al, 2003), WNT/RAC-1 (Habas et al, 2003), WNT/Ca2 þ (Ku¨hl et al, 2001), WNT/ROR2/CDC42 (Schambony & Wedlich, 2007) and WNT/ CK1/RAP1 (Tsai et al, 2007). b-Arrestin (b-arr) was recently found to regulate WNT/b-catenin signalling (Bryja et al, 2007b) and convergent extension movements in Xenopus laevis through RHOA (Kim & Han, 2007). The frequent involvement of b-arr in WNT signalling prompted us to examine whether it mediates the effects of WNTs on other branches of non-canonical signalling. We found that b-arr was required for WNT/RAC-1 signalling, but was not essential for the WNT/ROR2/CDC42 pathway. Furthermore, in vitro and in vivo data support that CK1/2 prevent the activation of RAC-1, but promote the formation of phosphorylated and shifted dishevelled (PS-DVL; dishevelled, xDSH in Xenopus). This indicates that the recruitment of b-arr to DVL selects and is required for WNT/RAC-1 signalling, and that casein kinases switch signalling from WNT/RAC-1 towards non-canonical pathways mediated by PS-DVL. RESULTS AND DISCUSSION WNT-induced activation of RAC-1 requires b-arr First, we examined the activation of small GTPases by WNT-5A in mouse embryonic fibroblasts (MEFs) using GTPase pull-down assays. Although WNT-5A, which does not affect b-catenin signalling (supplementary Fig S1A online), activated RAC-1 (Fig 1A; supplementary Fig S1B online; 4.1-fold±0.8 standard Received 11 August 2008; accepted 8 September 2008; published online 24 October 2008 *These authors contributed equally to this work + Corresponding author. Tel: þ 46 8 52487663; Fax: þ 46 8 34 1960; E-mail: ernest.arenas@ki.se ++ Corresponding author. Tel: þ 46 8 52487933; Fax: þ 46 8 34 1280; E-mail: gunnar.schulte@ki.se 1 Laboratory of Molecular Neurobiology, Department of Medical Biochemistry & Biophysics, Karolinska Institutet, Scheeles v¨ag 1, S-171 77 Stockholm, Sweden 2 Institute of Biophysics of Academy of Sciences of the Czech Republic & Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlarska 2, 611 37 Brno, Czech Republic 3 Zoologisches Institut II, Universita¨t Karlsruhe (TH), Kaiserstrasse 12, D-76131 Karlsruhe, Germany 4 Boston University, Boston, 650 Albany Street, Massachusetts 02118, USA 5 Section of Receptor Biology & Signaling, Department of Physiology & Pharmacology, Karolinska Institutet, Nanna Svartz va¨g 2, S-171 77 Stockholm, Sweden w Present address: Developmental Biology Unit, Biology Department, University of Erlangen-Nuremberg, Staudtstrasse 5, 91058 Erlangen, Germany EMBO reports VOL 9 | NO 12 | 2008 &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION scientificreportscientific report 1244 error of the mean, s.e.m.), it activated neither RHO-A nor CDC42 (supplementary Fig S1C online). Interestingly, the activation of RAC-1 by WNT-5A was blocked in MEFs lacking b-arr1 and b-arr2 (b-arr1/2 double knock out (dKO)), arguing that b-arr is required for WNT/RAC-1 signalling. In the same cells, WNT-5A-induced formation of PS-DVL (Gonza´lez-Sancho et al, 2004; Bryja et al, 2007c) was less efficient and delayed compared with wild-type MEFs (Fig 1B). It is important to note that although the lack of b-arr interferes with WNT-induced dynamics of PS-DVL, it increases the basal PS-DVL (Fig 1B). This suggests that b-arr is not necessary for PS-DVL but modulates its formation. Re-transfection of b-arr1 in dKO MEFs reduces the basal PS-DVL and restores the WNT-5Ainduced activation of RAC-1 (supplementary Fig S1D online). To analyse whether b-arr is sufficient for the activation of RAC-1, we overexpressed b-arr2-FLAG in MEFs and observed increased activation of RAC-1 (Fig 1C; supplementary Fig S1E online). To confirm these results, we overexpressed b-arr2-FLAG and downregulated xb-arr by using antisense morpholinos (Xb-arr MO; Bryja et al, 2007b) in X. laevis embryos. We found that overexpression of b-arr increased basal RAC-1 activity and that it was decreased on downregulation (supplementary Fig S1F online). Furthermore, the effect of the xb-arr MO could be reversed by overexpression of MO-insensitive, murine b-arr, emphasizing the specificity of the xb-arr MO. These findings identify b-arr as a signalling component required for the regulation of RAC-1 in vitro and in vivo. For functional analysis, we performed Keller open-face explant elongation assays in X. laevis embryos, an established measure of convergent extension (Keller et al, 2003; Schambony & Wedlich, 2007). Both overexpression and downregulation of b-arr modulated convergent extension, and the latter was reversed by co-injection of the MO-insensitive haemagglutinin (HA)-b-arr (supplementary Fig S1F online), confirming previous results (Kim & Han, 2007). Kim & Han (2007) showed that b-arr2 regulates convergent extension movements in 0 20 40 60 80 100 120 * * Percentageofelongatedexplants n/#exp 131/9 97/8 27/2 57/4 xβ-Arr MO xDSHΔDIX caRHO-A caRAC-1 ++ + + + + + + + 58/4 36/3 30/2 caCDC42+ xNR3+ Elongation – + – +WNT-5A GTP-RAC-1 WT dKO RAC-1 β-Arr1/2A C GTP-RAC-1 FLAG RAC-1 DVL2 β-Arr2-FLAG – + A1CT RAC-1 GTP-RAC-1 DVL2-MYC – + – + DVL2-MYC siRNA Ctrl β-Arr1/2E WNT-5A (ng/ml) β-Arr1/2 WT β-Arr1/2 dKO DVL2 DVL2 0 10 30 100 300 1000 5WNT-5A (min) β-Arr1/2 WT β-Arr1/2 dKO DVL2 DVL2 0 15 30 60 120 B D Fig 1 | b-Arrestin is necessary for WNT-induced activation of RAC-1. (A) WNT-5A induced the activation of RAC-1 (GTP-RAC-1) in wild-type MEFs but not in MEFs lacking b-arrestin (b-arr1/2 double knock out (dKO)). One representative experiment is shown; in total, three are summarized in supplementary Fig S1B online. (B) WNT-5A-induced formation of phosphorylated and shifted (DVL2, open arrowhead; PS-DVL2, filled arrowhead) in a dose- and time-dependent manner in wild-type and b-arr1/2dKO MEFs is shown. (C) The activation and expression of RAC-1 were monitored in MEFs overexpressing b-arr2-FLAG. (D) Downregulation of b-arr is compensated by constitutively active RAC-1 and RHO-A. Keller open-face explants were scored for elongation on injection with Xb-arr MO and indicated RNAs. Asterisks indicate values that are significantly different (P40.95, t-test) from xb-arr MO. (E) Downregulation of b-arr1/2 expression by short interfering RNA (siRNA) in HEK293 cells does not affect the DVL2-MYCinduced activation of RAC-1 (GTP-RAC-1). Levels of b-arr are detected by the A1CT antibody. ca, constitutively active; DVL/DSH, dishevelled; HEK, human embryonic kidney; MEF, mouse embryonic fibroblast; MO, morpholino; WT, wild type. b-Arrestin in the WNT/RAC-1 pathway V. Bryja et al &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION EMBO reports VOL 9 | NO 12 | 2008 scientificreport 1245 X. laevis through RHO-A. Thus, our results, combined with those reported by Kim & Han, suggest that various subsets of RHO family GTPases are regulated by b-arr. b-Arr acts upstream from RAC-1 in convergent extension To examine the regulation of RAC-1 by b-arr, we analysed the effect of the following RNAs encoding for WNT signalling components on the convergent extension phenotype in b-arr knockdowns: xDSHDDIX (a mutant lacking the DIX domain), a constitutively active WNT/planar cell polarity DVL mutant, constitutively active caRHO-A, caRAC-1 or caCDC42, and xNR3, a WNT/b-catenin transcriptional target (Fig 1D; Ku¨hl et al, 2001; Bryja et al, 2007b). Both RAC-1 and RHO-A compensated for the lack of b-arr, and xDSHDDIX partly restored explant elongation, whereas CDC42 and xNR3 had no effect (Fig 1D). Complementary results from b-arr1/2 short interfering RNA (siRNA) experiments indicate that b-arr is not required for the DVL-induced activation of RAC-1 (Fig 1E). Other epistatic experiments showed that RHO-A and RAC-1 act downstream from xDSH in the regulation of convergent extension (supplementary Fig S1H online). When discussing the WNT-induced b-arr-dependent formation of PS-DVL, it is worthwhile mentioning that WNT-3A does not induce the activation of RAC-1 in MEFs (supplementary Fig S1I online), despite the fact that it induces b-arr-dependent PS-DVL (Bryja et al, 2007b) with a similar time course (Bryja et al, 2007c). Taken together, these results indicate that: b-arr acts upstream or on the level of DVL; RHO-A and RAC-1 but not CDC42 are mediators downstream from b-arr and DVL; and the WNT/ b-catenin pathway is not involved in the observed phenotype. b-Arr is not required for xWNT-5A/ROR signalling It has been shown previously that WNT-regulated convergent extension in X. laevis Keller explants is mediated by several distinct molecular mechanisms (Tada et al, 2002). Although xWNT-11 exerts its effects on convergent extension by activation of the WNT/RAC-1, WNT/RHO and WNT/FZD7/Ca2 þ pathways (Winklbauer et al, 2001; Habas et al, 2003), xWNT-5A is not required for Keller explant elongation but instead for explant constriction, and activates the WNT/ROR2/CDC42 pathway (Unterseher et al, 2004; Schambony & Wedlich, 2007). To examine the function of b-arr for xWNT-11 or xWNT-5A signals, we examined whether b-arr mediated their effects in Keller explants. Strikingly, only xWNT-11 (Fig 2A), but not xWNT-5A (Fig 2B), was reversed by Xb-arr MO co-injection. Consistently, the negative effects of xWNT-11 MO on explant elongation were partly reversed by the overexpression of b-arr (Fig 2A); however, b-arr did not restore constriction in xWNT-5A-depleted explants (Fig 2B), arguing that b-arr is not involved in the xWNT-5A/ROR2 pathway. To relate the xWNT-11-induced pathway to the small GTPases, we examined whether caRHO-A, caRAC-1 and caCDC42 or dnRHO-A and dnRAC-1 (supplementary Fig S2A online) rescued the effects of xWNT-11 MO or xWNT-11, respectively. Reduced explant elongation in xWNT-11-depleted explants was partly rescued by coexpression of caRHO-A and caRAC-1 (but not caCDC42), and xWNT-11 overexpression was compensated by dnRHO-A, dnRAC-1 and xDSHDDEP, a mutant lacking the RAC-1-activating DEP domain (supplementary Fig S2B online; 0 20 40 60 80 100 120 Percentageofelongatedexplants 0 20 40 60 80 100 120 Percentageofconstrictedexplants + n/#exp 65/3 24/2 97/7 xWNT-5A xβ-Arr MO xWNT-5A MO xβ-Arr + + + + 69/5 + Elongation Constriction B A xβ-Arr+ Elongation n/#exp 131/9 99/5 36/3 108/7 xWNT-11 xβ-Arr MO xWNT-11 MO + + + + 44/3 + 0 20 40 60 80 100 120 * * Percentageofelongatedexplants Fig 2 | b-Arrestin mediates WNT-11- but not WNT-5A-induced gastrulation movements in Xenopus laevis embryos. Keller open-face explant elongation (A,B, upper panels) and constriction (B, lower panel) were scored in X. laevis embryos injected with indicated RNAs and morpholinos (MOs). Asterisks indicate values that are significantly different (P40.95, t-test) from xWNT-11 and xWNT-11 MO. b-Arrestin in the WNT/RAC-1 pathway V. Bryja et al EMBO reports VOL 9 | NO 12 | 2008 &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION scientificreport 1246 Habas et al, 2003). These experiments confirmed the crucial function of RAC-1 and RHO-A downstream from xWNT-11 in convergent extension movements and indicated that xWNT-11 signalling to RHO/RAC also includes xDSH. CK1/2 as switches in non-canonical WNT signalling It is known that WNTs directly activate at least two kinases: casein kinase (CK)1 and CK2 (Willert et al, 1997; Swiatek et al, 2004; Gao & Wang, 2006; Bryja et al, 2007d), which, in turn, phosphorylate DVL to form PS-DVL. In line with these findings, it has been shown previously that CK1d/e are crucial components of the b-catenin-independent WNT signalling machinery (McKay et al, 2001; Klein et al, 2006; Strutt et al, 2006; Bryja et al, 2007d). Furthermore, it was recently shown that CK1e acts through the small GTPases RAP1 (Tsai et al, 2007). Interestingly, CK1 and CK2 can physically interact with b-arr (Xiao et al, 2007), and we have already shown that CK1 mediates communication between b-arr and DVL in the WNT/b-catenin pathway (Bryja et al, 2007b). To test whether PS-DVL and CK1 have a similar function in b-arr-dependent WNT/RAC-1 signalling, we inhibited CK1 pharmacologically with D4476 (4-[4-(2,3-dihydro-benzo[1,4]dioxin-6-yl)-5-pyridin-2-yl-1H-imidazol-2-yl]benzamide). As expected, D4476 led to a reduction of basal PS-DVL2 levels but, surprisingly, it increased the activation of RAC-1 (Fig 3A). Similar results were obtained with the inhibition of CK2 by TBBt (4,5,6,7-tetrabromo-2-azabenzotriazole; Fig 3A). The presence of the DVL protein per se seems to be required for both the activation of RAC-1 and the maintenance of RAC-1 expression levels (supplementary Fig S3A online). Thus, DVL as a protein, but not PS-DVL, which is dependent on CK1 and CK2, is required for the activation of RAC-1. In summary, our data suggest that CK1 and CK2, and possibly also PS-DVL, negatively regulate WNT/RAC-1 signalling in vitro. Importantly, WNTs seem to concomitantly evoke the activation of RAC-1/RHO-A (Habas et al, 2003; Fig 1A) and the formation of PS-DVL (Gonza´lez-Sancho et al, 2004; Schulte et al, 2005; Bryja et al, 2007d; Fig 1B), which depend on CK1 (Bryja et al, 2007d) and CK2 (Gao & Wang, 2006; Fig 3A). Interestingly, PS-DVL is usually formed 30–45 min after stimulation with WNT-5A (Bryja et al, 2007d; Fig 1B), whereas RAC-1 was shown to be already activated 5 min after stimulation (Kurayoshi et al, 2006), raising the possibility that the activation of RAC-1 precedes the formation of PS-DVL and activation of other branches of non-canonical signalling. Can CK1/CK2 thus act as a switch between individual B CK1ε DVL3-FLAG – – + + – + – + Actin DVL3-FLAG RAC-1 GTP-RAC-1 CK1ε 0 20 40 60 80 100 120 * Percentageofelongated explants Elongation n/#exp 131/9 97/8 50/4 xβ-Arr MO CK1ε CK2α CK2β ++ + + + + + + + + + + + ++ – – – – –––––– – – – – – – + 43/3 48/3 40/3 + + 200 30 pg CA GTP-RAC-1 RAC-1 DVL2 CK2CK1 – + – + Elongation n/#exp 131/9 38/3 xWNT-11 MO CK1ε CK2α CK2β + + + + + + + + + + 41/3 60/4 46/3 + + 200 30 pg 148/10 37/3 36/3 0 20 40 60 80 100 120 Percentageofelongatedexplants D * * * E xWNT-11 MO CK2 inhibitor RAC-1 GTP-RAC-1 CK1 inhibitor Inhibitor CK1 inhibitor CK2 inhibitor + + * Fig 3 | Casein kinases 1 and 2 differentially regulate the formation of PS-DVL and the activation of RAC-1. (A) The activation of RAC-1 (GTP-RAC-1), and the expression levels of RAC-1 and DVL2/PS-DVL2 (DVL2, open arrowhead; PS-DVL2, filled arrowhead) in MEFs were monitored in the absence (À) or presence ( þ ) of the CK1 (100 mM) or CK2 inhibitor (100 mM). (B) HEK293 cells were transfected with RAC-1-MYC in the presence of DVL3-FLAG and/or CK1e. RAC-1-MYC activity (GTP-RAC-1) was determined by pull down, and RAC-1-MYC, DVL3-FLAG, CK1e and actin levels were determined in cell lysates. Note the CK1e-induced mobility shift of DVL3-FLAG. (C,D) Explant elongation after downregulation of b-arrestin (C; xb-arr MO) or xWNT-11 (D; xWNT-11 MO), in combination with CK1e, CK2a, CK2b and CK2a/b, or CK1 and CK2 inhibitors, was evaluated. Asterisks indicate values that are significantly different (P40.95, t-test) from (C) xb-arr MO and (D) xWNT-11 MO. (E) Effect of xWNT-11 MO on the activation of RAC-1 in the absence (À) and presence ( þ ) of CK1 or CK2 inhibitors was measured by active GTP-RAC-1 pull down from Xenopus embryo lysates. Levels of RAC-1 in the lysates are shown in the lower panel. CK, casein kinase; HEK, human embryonic kidney; MEF, mouse embryonic fibroblast; MO, morpholino; PS-DVL, phosphorylated and shifted dishevelled. b-Arrestin in the WNT/RAC-1 pathway V. Bryja et al &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION EMBO reports VOL 9 | NO 12 | 2008 scientificreport 1247 b-catenin-independent WNT pathways at the level of DVL? To test this possibility, we induced the activation of RAC-1 directly by DVL3 and studied the effects of CK1 on this process (Fig 3B). Although overexpression of DVL3 increased the activation of RAC-1, this positive effect on RAC-1 was abolished when the formation of PS-DVL3 was induced by simultaneous CK1e co-expression. Furthermore, we found, by using the AP-1 reporter assay, that basal and DVL3––but not phorbolester––or ca-RAC-1induced AP-1 activity was enhanced by treatment with D4476 (supplementary Fig S3B online). These data suggest that: PS-DVL (mediated by CK1/CK2 activity) does not participate in the WNT/ b-arr/RAC-1/AP-1 pathway; and that the activation of CK1/CK2 and formation of PS-DVL leads to a change in the signalling function of DVL, displacing it from the RAC-1 pathway. CK 1 and 2 in the regulation of convergent extension In vivo, we confirmed that both CK1 and CK2 are involved in and required for the regulation of convergent extension movements in Keller explants (supplementary Fig S3C online). However, we observed differences in the ability of CK1 and CK2 to restore elongation in xb-arr or xWNT-11-depleted explants. xb-arr MO was not rescued by CK1e, CK2a or CK2 holoenzyme and only partly by CK2b (Fig 3C). By contrast, CK2a, CK2b and CK2 holoenzyme, but not CK1e rescued xWNT-11 depletion (Fig 3D). Consistently, xWNT-11 overexpression was rescued by dominant negative (dn) CK2a, but not by dnCK1e (supplementary Fig S3D online). Furthermore, in agreement with our biochemical data, chemical inhibition of CK1 or CK2 restored explant elongation (Fig 3D) and RAC-1 activity (Fig 3E) in xWNT-11 MO-injected explants. Simultaneous downregulation of xWNT-11 and xDSH (for optimizing xDSH MOs, see supplementary Fig S4A online) resulted in an additive phenotype that was rescued by co-injection of CK2a, but not CK2b (Fig 4A). Overall, these results indicate that in vivo CK1e and CK2a do not positively regulate Xb-arr-dependent branches of non-canonical signalling, and that xWNT-11 activates not only b-arr/xDSH/RAC-1 signalling but also xDSH-MYC/DAPI CK1ε xDSH-MYC/DAPI CK1 inhibitor xDSH-MYC/DAPI xDSH-MYC/DAPI xDSH-MYC/DAPI xDSH-MYC/DAPI ctrl xWNT-5A xWNT-1 1xWNT-8 D Other branches of PS-DVL signalling β-Arrestin RAC-1 CK2 CK1 WNT PS-DVL DVL DVL 0 20 40 60 80 100 120 * * Percentageofelongatedexplants A n/#exp 131/9 46/4 xDSH MO xWNT-11 MO CK2α + 24/2 23/2 + + CK2β+ + + + +++ Elongation 148/10 41/4 57/4 55/10 + + + + + + + CK1 inhibitor CK2 inhibitor B C xDSH-MYC xDSH-MYC B62A12 + CK1ε CK1εK-RCK1inhibitor –+ + + + + + xW NT-8xW NT-5AxW NT-11 WNT-5A in MEFs xWNT-11 in Xenopus Fig 4 | DVL-dependent and -independent function of casein kinases. (A) Convergent extension was quantified in embryos treated with both xDSH MO and xWNT-11 MO in combination with overexpression of CK2a, CK2b or CK1, CK2 inhibition. Elongation of Keller explants was quantified. Asterisks indicate values that are significantly different (P40.95, t-test) from xWNT-11 MO/xDSH MO. (B) Immunocytochemical localization of xDSH-MYC (nuclear counterstain DAPI) in Xenopus mesoderm in control (ctrl), WNT- or CK1e-overexpressing or D4476-treated embryos. Arrowheads point at membraneous xDSH-MYC. (C) Corresponding biochemical data. xDSH-MYC was detected in embryo lysates and analysed for the expression of xDSH-MYC (xDSH, open arrowhead; PS-xDSH, filled arrowhead). The antigen B62A12 was used as a loading control. (D) Schematic view of the role of casein kinases and b-arrestin in the regulation of formation of PS-DVL and WNT-induced signalling to RAC-1; for details, see text. CE, convergent extension; CK, casein kinase; DAPI, 4,6-diamidino-2-phenylindole; DVL/DSH, dishevelled; MO, morpholino; PS-DVL, phosphorylated and shifted DVL. b-Arrestin in the WNT/RAC-1 pathway V. Bryja et al EMBO reports VOL 9 | NO 12 | 2008 &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION scientificreport 1248 other pathways that involve CK2, most likely WNT/Ca2 þ signalling. This assumption is supported by the observation that CK2a rescues convergent extension in PTX-treated explants (unpublished observation). Interestingly, the positive effect of inhibition of CK1 on convergent extension was strictly xDSH dependent, whereas inhibition of CK2 led to explant elongation (Fig 4A) and the activation of RAC-1 (supplementary Fig S4B online) even after knockdown of xWNT-11/xDSH. To address the requirement of DVL for the CK-modulated activation of RAC-1 in mammalian the cells, we used pretreatment with hyperosmolaric sucrose to deplete cells of DVL (Bryja et al, 2007a). Under these conditions, only inhibition of CK2 increased the activation of RAC-1, further supporting the idea that CK2 but not CK1 mediates the activation of RAC-1 at least partly independently of DVL (supplementary Fig S4C online). When we analysed xDSH subcellular localization (Fig 4B) in response to D4476, we observed that inhibition of CK1 activity resulted in translocation of xDSH-MYC to the cell membrane, similar to the overexpression of xWNT-11, but not of xWNT-5A or xWNT-8, whereas xDSH-MYC was found predominantly in the cytoplasm in CK1e-overexpressing cells. The effect of the treatment was also monitored by the decrease or increase of PS-xDSH, respectively (Fig 4C). This indicates that submembraneous DVL could represent a pool of DVL that is active in the RAC-1 pathway and that is not phosphorylated by CK1/2, but does not form PS-DVL. In summary, our data first suggest that b-arr is not essential for the WNT/ROR2/CDC42 and non-canonical pathways involving CK1/2 and PS-DVL. However, we found that b-arr is selectively required for the WNT/RAC-1 pathway. This finding selectively links WNT/RAC-1 signalling with other b-arr-regulated pathways and functions, such as receptor endocytosis, internalization and desensitization. Second, our in vitro and in vivo data support that the formation of PS-DVL (which requires CK1 and/or CK2) and the activation of RAC-1 (which requires inhibition of CK1/2) represent independent events, and potential ways to measure two b-catenin-independent pathways activated by a common ligand (WNT5A or XWNT-11). We propose that CK1 and CK2 can act as a central relay to select the pathways and signalling components being positively and negatively regulated (PS-DVL and RAC-1, respectively, schematized in Fig 4D). Thus, CK1 and CK2 can be considered as switches between individual branches of b-catenin-independent WNT signalling. METHODS Cell culture, transfection and treatments. Wild-type MEFs and MEFs lacking b-arr 1 and 2 (b-arr1/2dKO) were a gift from Dr Lefkowitz (Kohout et al, 2001). MEFs and human embryonic kidney (HEK) 293 cells were grown in DMEM, 10% FCS, L-glutamine (2 mM), penicillin (50 U/ml) and streptomycin (50 U/ml). Transfection and treatment with D4476, casein kinase 2 inhibitor I (Calbiochem, Nottingham, UK)-TBBt or recombinant mouse WNT-5A (from R&D Systems, Abingdon, UK) were carried out as described previously (Bryja et al, 2007c). For a list of expression vectors, see the supplementary information online. GTPase pull-down assay and immunoblotting. GST-PAK-CRIB (for glutathione-S-transferase-p21-activated kinase-CDC42/RAC interactive binding domain), GST-WASP-CRIB (WASP for Wiskott– Aldrich syndrome protein) and GST-RHOtekin recombinant proteins were coupled to glutathione sepharose beads for the detection of activated RAC-1, CDC42 and RHO-A, respectively, and the assay performed as described in the supplementary information online. Immunoblotting and sample preparation were carried out as described previously (Bryja et al, 2007c). Xenopus oocytes and Keller explants. For RNA transcription and oocyte injection, see the supplementary information online. Embryos were cultivated until Nieuwkoop and Faber stage 10.5, and Keller explants were prepared, cultivated and scored as described previously (Unterseher et al, 2004). For inhibitor treatment, D4476 and CK2 inhibitors were added to the medium at 1 mM at the blastula stage (stage 8). xDSH-MYC translocation was examined after injection of 100 pg xDSH-MYC RNA. At onset of gastrulation (stage 10.5), the dorsal marginal zone and the blastocoel roof were prepared and fixed in 4% formalin. Tissue was stained with 9E10 anti-MYC, mounted and examined with a laser scanning confocal microscope. Supplementary information is available at EMBO reports online (http://www.emboreports.org). ACKNOWLEDGEMENTS The study was supported by the Swedish Foundation for Strategic Research (INGVAR and Center of Excellence for Developmental Biology), Swedish Research Council, European Union (Eurostemcell), Karolinska Institutet, A˚ ke Wiberg, Signe & Olof Wallenius, Jeanssons, Tore Nilsson Foundations, German Research Foundation (DFG), Ministry of Education and Grant Agency of the Academy of Sciences of the Czech Republic and the European Molecular Biology Organization (for more details, see supplementary information online). CONFLICT OF INTEREST The authors declare that they have no conflict of interest. REFERENCES Bryja V, Caja´nek L, Grahn A, Schulte G (2007a) Inhibition of endocytosis blocks Wnt signalling to b-catenin by promoting dishevelled degradation. Acta Physiol 190: 55–61 Bryja V, Gradl D, Schambony A, Arenas E, Schulte G (2007b) b-Arrestin is a necessary component of Wnt/b-catenin signaling in vitro and in vivo. Proc Natl Acad Sci USA 104: 6690–6695 Bryja V, Schulte G, Arenas E (2007c) Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate b-catenin. Cell Signal 19: 610–616 Bryja V, Schulte G, Rawal N, Grahn A, Arenas E (2007d) Wnt-5a induces dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J Cell Sci 120: 586–595 Gao Y, Wang HY (2006) Casein kinase 2 is activated and essential for Wnt/b-catenin signaling. J Biol Chem 281: 18394–18400 Gonza´lez-Sancho JM, Brennan KR, Castelo-Soccio LA, Brown AM (2004) Wnt proteins induce dishevelled phosphorylation via an LRP5/6-independent mechanism, irrespective of their ability to stabilize b-catenin. Mol Cell Biol 24: 4757–4768 Habas R, Dawid IB, He X (2003) Coactivation of Rac and Rho by Wnt/ Frizzled signaling is required for vertebrate gastrulation. Genes Dev 17: 295–309 Keller R, Davidson LA, Shook DR (2003) How we are shaped: the biomechanics of gastrulation. Differentiation 71: 171–205 Kim GH, Han JK (2007) Essential role for b-arrestin 2 in the regulation of Xenopus convergent extension movements. EMBO J 26: 2513–2526 Klein TJ, Jenny A, Djiane A, Mlodzik M (2006) CKIe/discs overgrown promotes both Wnt-Fz/b-catenin and Fz/PCP signaling in Drosophila. Curr Biol 16: 1337–1343 b-Arrestin in the WNT/RAC-1 pathway V. Bryja et al &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION EMBO reports VOL 9 | NO 12 | 2008 scientificreport 1249 Kohout TA, Lin FS, Perry SJ, Conner DA, Lefkowitz RJ (2001) b-Arrestin 1 and 2 differentially regulate heptahelical receptor signaling and trafficking. Proc Natl Acad Sci USA 98: 1601–1606 Ku¨hl M, Geis K, Sheldahl LC, Pukrop T, Moon RT, Wedlich D (2001) Antagonistic regulation of convergent extension movements in Xenopus by Wnt/b-catenin and Wnt/Ca2+ signaling. Mech Dev 106: 61–76 Kurayoshi M, Oue N, Yamamoto H, Kishida M, Inoue A, Asahara T, Yasui W, Kikuchi A (2006) Expression of Wnt-5a is correlated with aggressiveness of gastric cancer by stimulating cell migration and invasion. Cancer Res 66: 10439–10448 McKay RM, Peters JM, Graff JM (2001) The casein kinase I family: roles in morphogenesis. Dev Biol 235: 378–387 Schambony A, Wedlich D (2007) Wnt-5A/Ror2 regulate expression of XPAPC through an alternative noncanonical signaling pathway. Dev Cell 12: 779–792 Schulte G, Bryja V, Rawal N, Castelo-Branco G, Sousa KM, Arenas E (2005) Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation. J Neurochem 92: 1550–1553 Semenov MV, Habas R, Macdonald BT, He X (2007) SnapShot: noncanonical Wnt signaling pathways. Cell 131: 1378 Strutt H, Price MA, Strutt D (2006) Planar polarity is positively regulated by casein kinase Ie in Drosophila. Curr Biol 16: 1329–1336 Swiatek W, Tsai IC, Klimowski L, Pepler A, Barnette J, Yost HJ, Virshup DM (2004) Regulation of casein kinase I e activity by Wnt signaling. J Biol Chem 279: 13011–13017 Tada M, Concha ML, Heisenberg CP (2002) Non-canonical Wnt signalling and regulation of gastrulation movements. Semin Cell Dev Biol 13: 251–260 Tsai IC, Amack JD, Gao ZH, Band V, Yost HJ, Virshup DM (2007) A WntCKIvare-Rap1 pathway regulates gastrulation by modulating SIPA1L1, a Rap GTPase activating protein. Dev Cell 12: 335–347 Unterseher F, Hefele JA, Giehl K, De Robertis EM, Wedlich D, Schambony A (2004) Paraxial protocadherin coordinates cell polarity during convergent extension via Rho A and JNK. EMBO J 23: 3259–3269 Willert K, Brink M, Wodarz A, Varmus H, Nusse R (1997) Casein kinase 2 associates with and phosphorylates dishevelled. EMBO J 16: 3089–3096 Winklbauer R, Medina A, Swain RK, Steinbeisser H (2001) Frizzled-7 signalling controls tissue separation during Xenopus gastrulation. Nature 413: 856–860 Xiao K, McClatchy DB, Shukla AK, Zhao Y, Chen M, Shenoy SK, Yates JR, Lefkowitz RJ (2007) Functional specialization of b-arrestin interactions revealed by proteomic analysis. Proc Natl Acad Sci USA 104: 12011–12016 b-Arrestin in the WNT/RAC-1 pathway V. Bryja et al EMBO reports VOL 9 | NO 12 | 2008 &2008 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION scientificreport 1250 1 Supplementary table: Table S1: Expression vector Supplied by/reference -arrestin2-FLAG R. J. Lefkowitz (HHMI, Durham, NC) -arrestin2-HA R. J. Lefkowitz (HHMI, Durham, NC) X -arrestin A. Schambony / D. Gradl XDSH DIX R. T. Moon (HHMI, University of Washington, Seattle, WA) XDSH DEP R. T. Moon (HHMI, University of Washington, Seattle, WA) XDSH-MYC R. T. Moon (HHMI, University of Washington, Seattle, WA) caRHO A A. Schambony (Unterseher et al. 2004) caRAC-1 A. Schambony (Unterseher et al. 2004) caCDC42 K. Giehl (University Ulm, Germany) dnRHO A A. Schambony (Unterseher et al. 2004) dnRAC-1 A. Schambony (Unterseher et al. 2004) XNR3 R. T. Moon (HHMI, University of Washington, Seattle, WA), Kuehl et al. 2001 XWNT-11 R. T. Moon (HHMI, University of Washington, Seattle, WA) XWNT-5A R. T. Moon (HHMI, University of Washington, Seattle, WA), Kuehl et al. 2001 XWNT-8 R. T. Moon (HHMI, University of Washington, Seattle, WA), Kuehl et al. 2001 CK1 J. M. Graff (University of Texas, Dallas, TX) DVL3-FLAG R. T. Moon (HHMI, University of Washington, Seattle, WA) RAC-1-MYC Alan Hall (University College London, UK) CK2 Isabel Dominguez (Boston University, MA, USA) CK2 Isabel Dominguez (Boston University, MA, USA) CK1 K-R J. M. Graff (University of Texas, Dallas, TX) dnCAMKII R. T. Moon (HHMI, University of Washington, Seattle, WA) pAP1-tk-Luc Philippe Lefebvre We are thankful to the above-mentioned researchers for generously providing vectors used in this study. 2 Unterseher F, Hefele JA, Giehl K, De Robertis EM, Wedlich D, Schambony A (2004) Paraxial protocadherin coordinates cell polarity during convergent extension via Rho A and JNK. EMBO Journal 23: 3259-3269. Kühl M, Geis K, Sheldahl LC, Pukrop T, Moon RT, Wedlich D (2001) Antagonistic regulation of convergent extension movements in Xenopus by Wnt/beta-catenin and Wnt/Ca2+ signaling. Mech Dev 106: 61-76 3 Supplementary information: List of abbreviations: AP-1 - activator protein 1 -arr1/2dKO MEFs - MEFs lacking both isoforms of -arrestin ca - constitutively active CE - convergent extension CK1 - casein kinase 1 CK1 K-R - K to R mutation of CK1 (kinase dead) CK2 - casein kinase 2 DAAM - Dishevelled-associated activator of morphogenesis D4476 - (4-[4-(2,3-dihydro-benzo[1,4]dioxin-6-yl)-5-pyridin-2-yl-1H -imidazol-2-yl]benzamide) DVL (PS-DVL) - mammalian Dishevelled (phosphorylated and shifted DVL) (X)DSH - Xenopus Dishevelled ERK1/2 - extracellular signal-regulated kinase 1/2 FZD - Frizzled GTPase - GTP hydrolase GST - Glutathione-S-transferase GST-PAK-CRIB - GST-p21-activated kinase-Cdc42/Rac interactive binding GST-WASP-CRIB - GST-Wiskott-Aldrich Syndrome protein-CRIB JNK - c-jun N terminal kinase MEF - mouse embryonic fibroblast MO - morpholino PCP - planar cell polarity PI3K - phosphatidylinositol-3’-kinase PKB - protein kinase B RAP1 - RAS-related protein 1, RAS proximate RAC-1, RHO A, CDC42 - RHO-like GTPases ROR2 - receptor tyrosine kinase ROR2 c-SRC - cytosolic tyrosine kinase identified from the oncogene src 4 SDS-PAGE - sodium dodecyl sulfate polyacrylamide gelelectrophosresis SEM - standard error of the mean WNT - acronym wingless (from Drosophila) and the oncogene int-1 X -arr - Xenopus -arrestin XNR3 - Xenopus Nodal-related 3 XWNT - Xenopus WNT 5 Supplementary material and methods: Small GTPase pull down MEFs were serum starved overnight and stimulated with recombinant WNT-5A (200 ng/ml, R&D Systems) for 2 h. They were pretreated with high molar sucrose (2 h) and a subsequent wash in norm-osmolaric medium or with CK1 (D4476) or CK2 inhibitors (100 M) as indicated. Transfected Hek293 cells were used 24 h posttransfection for small GTPase pull down. Xenopus samples were obtained as follows: embryos were injected into the marginal zone of the dorsal blastomeres at the 4cell stage and cultivated until they reached stage 10.5. Fifty embryos per sample were collected, lysed and processed. Cells were washed with ice cold PBS and subsequently allowed to lyse in ice cold lysis buffer (10 mM Tris-Cl – pH = 7.5, 110 mM NaCl, 1 mM EDTA, 10 mM MgCl2, 1% Triton X-100, 0.1% SDS, 20 mM glycerolphosphate, 1 mM dithiotreitol, complete protease inhibitors (Roche) for 5 min. Crude cell lysates were spun down in chilled tubes at 14000 rpm for 5 min at 4°C. Supernatants (5% saved as input) were supplemented with bait proteins coupled to glutathione sepharose beads and tubes were incubated rotating end-over-end at 4°C for 15 min. Beads were washed 3 times with washing buffer (lysis buffer without SDS and protease inhibitors) on ice and subsequently mixed with 2x Laemmli buffer. Each sample was boiled (5 min) before loading on SDS-PAGE. RNA injection and morpholino treatment of Xenopus oocytes Capped mRNAs were transcribed from linearized DNA templates using mMessage mMachine (Ambion). For knock-down experiments in Xenopus, morpholino antisense oligonucleotides targeted against Xenopus -arrestin (Bryja et al. 2007), XWNT-11 (Pandur et al. 2002), XWNT-5A (Schambony & Wedlich 2007) and XDSH (Sheldahl 6 et al. 2003) were used. Eggs from HCG-treated females were fertilized by standard methods and staged according to (Nieuwkoop., and Faber 1967). Morpholino oligonucleotides, RNA and DNA were injected into the marginal zone of the dorsal blastomeres of 4-cell stage embryos in a total volume of 4 nl. If not mentioned otherwise in the text, the following amounts were injected: arrMO: 1 pmol, arrestin2: 500 g, XWNT-5A: 100 pg, XWNT-11: 20 pg, XWNT-5A MO: 0.8 pmol, XWNT-11 MO: 1 pmol, XWNT-8: 20 pg, XDSH DIX: 500 pg, XDSH DEP: 500 pg, XDSH-myc: 100 pg, XDSH MO: 0,8 pmol, Xnr-3: 500 pg, CK1 : 500 pg, CK2 , CK2 : 200 pg, dnCAMKII: 1ng, dnJNK: 500 pg, dnRHO A: 50 pg, dnRAC-1: 100 pg, pCS2-caRHO A: 5 pg DNA, pCS2-caRAC-1: 10 pg DNA, pcDNA-caCDC42: 20 pg DNA. siRNA for -arrestin1/2 knock-down HEK293 cells were transfected with DVL2-MYC, RAC-1-MYC and either control siRNA ((AA)UUCUCCGAACGUGUCACGU; Xeragon (Qiagen) made as a nonsilencing siRNA which does not have matches with mammalian genes by a BLAST search) or -arrestin1/2 siRNA ((AA)ACCUGCGCCUUCCGCUAUG (positions 172-190 and 175-193 relative to the start codons of mouse -arrestin 1 and -arrestin 2, respectively) from Dharmacon using GeneSilencer (Genlantec) reagent (GestyPalmer et al. 2006). RAC-1 pull down assays were performed >72 h posttransfection after overnight serum deprivation. Levels of -arrestin1/2 were monitored with immunoblotting using the polyclonal A1CT antibody kindly provided by Prof RJ Lefkwitz, HHMI, Duke University Medical Center, Durham, NC, USA. AP-1 reporter assay 7 COS7 cells were transfected with with mixture of pAP1-tk-Luc (20 ng, (Benkoussa et al. 2002)), pRL-TK (200 ng, internal control, Promega) and indicated expression construct (200 ng of either pcDNA-DVL3-FLAG or 400 ng of pCS2-caRAC-1). pcDNA was used to equalize the total amount of DNA to 0.7 ug per 1 well in 24-well plate. Cells were serum starved 24 h after transfection by reducing serum to 1% and stimulated with D4476(100 uM)/TPA (0.8 uM) or DMSO (control) for additional 18 hours. Following treatments cells were lysed and further processed using Dual-Luciferase® Reporter 1000 Assay System (Promega) according to manufacturer´s instructions and measured on a MLX luminometer (Dynex Technologies). Each data point was run in duplicate and three independent experiments were performed. Data were normalized to control transfected COS7 cells. References: Benkoussa M, Brand C, Delmotte MH, Formstecher P, Lefebvre P (2002) Retinoic acid receptors inhibit AP1 activation by regulating extracellular signal-regulated kinase and CBP recruitment to an AP1-responsive promoter. Mol. Cell. Biol. 22: 4522-4534 Bryja V, Gradl D, Schambony A, Arenas E, Schulte G (2007) beta-Arrestin is a necessary component of Wnt/beta-catenin signaling in vitro and in vivo. Proc. Natl. Acad. Sci. U. S. A. 104: 6690-6695 Gesty-Palmer D, et al. (2006) Distinct beta-arrestin- and G protein-dependent pathways for parathyroid hormone receptor-stimulated ERK1/2 activation. J. Biol. Chem. 281: 10856-10864 Kim GH, Han JK (2007) Essential role for beta-arrestin 2 in the regulation of Xenopus convergent extension movements. EMBO J. 26: 2513-2526 Nieuwkoop, P.D. and Faber, J. (1967). Normal table of Xenopus laevis. Normal table of Xenopus laevis, (Amsterdam: Elsevier North-Holland Biochemical Press). 8 Pandur P, Läsche M, Eisenberg LM, Kühl M (2002) Wnt-11 activation of a noncanonical Wnt signalling pathway is required for cardiogenesis. Nature 418: 636- 641 Schambony A, Wedlich D (2007) Wnt-5A/Ror2 regulate expression of XPAPC through an alternative noncanonical signaling pathway. Dev. Cell 12: 779-792 Sheldahl LM, Slusarski DC, Pandur P, Miller JR, Kühl M, Moon RT (2003): Dishevelled activates Ca2+ flux, PKC and CamKII in vertebrate embryos. J. Cell Biol. 161: 769-777 9 Supplementary figures: Fig. S1: (A) WT MEFs were treated with increasing doses of WNT-5A for 2 h. Cell lysates were probed for DVL2 to monitor pathway activation. WNT-5A did not change -catenin phosphorylation as detected by the anti active -catenin antibody (ABC). (B) quantification of Fig. 1A. *** p=0.004 WT, WNT-5A vs. dKO, WNT- 5A; paired t-test. (C) GTPase pull down assays for CDC42, RHO A and RAC-1 show that WNT-5A (2 h) induces RAC-1 but not CDC42/RHO A activation in WT MEFs. (D) -arrestin1/2 double-deficient MEFs (dKO) and dKO MEFs with the re-expressed -arrestin1 (a kind gift of Robert Lefkowitz) were treated with 300 ng/ml Wnt5a. RAC-1 activity and the levels of RAC1 and DVL3 (open triangle)/PSDVL3 (full triangle) were analyzed. (E) The bar graph summarizes three experiments as shown in Fig. 3C: * p=0.0161, paired t-test. (F) RAC-1 activity was measured in Xenopus embryos after injection of -arr2-FLAG RNA and upon X -arr morpholino (MO) treatment. For a control, MO treatment was combined with overexpression of -arr2-FLAG. (G) Consistent with previous results (Kim & Han 2007), we found that the downregulation of -arrestin with X arr MO, or the overexpression of -arrestin, reduced the intrinsic elongation from 100% to 36.3±4.3% (± SEM) or 47.6±3.1%, respectively. Furthermore, the inhibition of explant elongation by X arr MO was reversed by simultaneous injection of arrestin RNA (Fig. 2A), demonstrating the specificity of the approach. (H) Inhibition of explant elongation induced by overexpression of Dsh DIX, a PCP activating mutant, was rescued by co-expression of dnRHOA or dnRAC1. (I) WNT-3A (2 h) stimulation of WT MEFs did not induce RAC 1 activation. 10 Fig. S2: RAC-1/RHO A involvement in XWNT-11-mediated CE. (A) Keller explant elongation was quantified in XWNT-11 MO-treated embryos in combination with constitutively active small GTPases and in XWNT-11 overexpressing embryos in combination with dominant negative small GTPases. (B) Effect of DSH DEP on XWNT-11-induced CE defects in Xenopus embryos. 11 Fig. S3: Effects of CK1 and CK2 and on CE movements. (A) AP-1 activity was measured by an luciferase-based reporter assay in control, DVL3-FLAGoverexpressing, TPA (1 M)-treated or caRAC-1 expressing cells in the presence and absence of the CK1 inhibitor D4476. Results were normalized to DMSO-treated control cells. (B) DVL protein is required for basal RAC-1 expression as well as activation. MEFs were pretreated with 0.45 M sucrose for 2 h to deplete endogenous DVL (Bryja et al 2007). Subsequently cells were treated as indicated for 2 hours and a RAC-1 pull down assay was performed. Immunoblotting of the pull down as well as the total cell lysate was done to analyze RAC-1 activation (GTP-RAC-1), as well as RAC-1 and DVL2 levels. (C) Explant elongation is shown after injection of CK1 , CK1 K-R, CK2 , and at different concentrations. (D) Keller explant elongation was determined in CK1 or CK2-inhibitor-treated oocytes and in XWNT-11/dnCK1 or CK2 expressing oocytes. Reference: Bryja V, Cajanek L, Grahn A, Schulte G (2007) Inhibition of endocytosis blocks Wnt signalling to beta-catenin by promoting dishevelled degradation. Acta Physiol 190: 55-61 12 Fig. S4: (A) optimization of DSH MO. Keller explant elongation and constriction was measured in the embryos that were treated with DSH MO alone and in combination with casein kinase inhibitors. (B) RAC-1 activation (GTP-RAC-1) was measured in embryo lysates that were subjected to parallel XWNT-11 and DSH MO treatment with and without CK1 or CK2 inhibtion. RAC-1 levels are shown in the lower immunoblot. (C) WT MEFs were treated with hyperosmolaric (0.45 M) sucrose for two hours in combination with pharmacological inhibition of CK1 or CK2. After RAC-1 pull down, cell lysates and precipitates were analysed by immunoblotting for RAC-1/ DVL-2 and for GTP-RAC-1, respectively. 13 Individual financial support Support for: V.B. from MSM0021622430 (Ministry of Education, Youth and Sports of the Czech Republic), Grant Agency of the Academy of Sciences of the Czech Republic (GAAV, KJB501630801) and EMBO Installation Grant; G.S. from Swedish Brain Foundation (hjärnfonden) and Swedish Research Council (VR; 2007-2769 and 2007- 5595); A.S. from Deutsche Forschungsgemeinschaft (DFG SCHA 965); I.D. from Karin Grunebaum Cancer Research Foundation. β-arr2-FLAG - + 0.0 0.5 1.0 1.5 2.0 2.5 * RAC-1activation (foldofpcDNA) Figure S1 Supplementary Figures B 1 2 3 4 5 *** RAC-1actication (foldofctrlWT) - +WNT-5A WT dKOβ-arr1/2 - + WNT-5A (ng/ml) DVL2 active β-catenin DVL2 0 10 30 100 300 1000 A GTP-CDC42 CDC42 GTP-RHO A RHO A - +WNT-5A WTβ-arr1/2 GTP-RAC-1 RAC-1 C F E - + - + GTP-RAC-1 control RAC-1 β-arr2 Xβ-arr MO * GTP-RAC-1 RAC- 1 DVL2 - + WNT-3A 0 20 40 60 80 100 120 %elongatedexplants n/#exp 131/9 97/8 36/3 46/3 Xβ-arr MO HA-β-arr + + + + elongation G D GTP-RAC-1 RAC-1 DVL3 - + - +WNT-5A dKO dKO+ β-arr1β-arr1/2 H 0 20 40 60 80 100 120 * * %elongatedexplants elongation n/#exp 131/9 85/5 40/3 40/3 DSHΔDIX dn RHO A + + + I dn RAC-1+ + Figure S2 Supplementary Figures elongation n/#exp 131/9 99/5 46/3 XWNT-11 DSHΔDEP + + + B 0 20 40 60 80 100 120 * %elongatedexplants n/#exp 108/7 61/4 60/4 XWNT-11 MO caRHO A caRAC-1 caCDC42 + ++ + + + + 40/3 elongation 0 20 40 60 80 100 120 * * * * %elongatedexplants 131/9 33/2 41/3 39/3 + ++ + + XWNT-11 dnRHO A dnRAC-1 A ctrl D VL3-FLA G TPA caR A C -1 0 1 2 3 4 DMSO CK1 inhib. AP-1activity foldofcontrol 0 20 40 60 80 100 120 * * * * %elongatedexplants Figure S3 Supplementary Figures A RAC- 1 DVL2 - + sucrose GTP-RAC-1 B + n/#exp 131/9 31/3 32/3 CK1 inhib. CK2 inhib. + 148/10 37/3 + dnCK1ε dnCK2α XWNT-11+ + 0 20 40 60 80 100 120 %elongatedexplants n/#exp 131/9 44/343/3 CK1ε CK2α CK2β 43/3 39/3 43/3 28/2 CK1ε K-Rctrl 200 pg 30 pg CK2α/β CK2α/β elongation C D + XWNT-11 MO+ ++ + + + elongation 36/3 99/5 40/3 33/2 Figure S4 A n/#exp 131/9 49/4 46/4 DSH MO (pmol) CK1 inhib 0.8 24/2 23/2 CK2 inhib 1.6 1.6 1.6 + + 0 20 40 60 80 100 120 %constrictedexplants constriction C DVL2 - - +CK2 inhib. GTP-RAC-1 ctrl sucrose RAC-1 - - + - + -CK1 inhib. - + - Supplementary Figures B CK2 inhib RAC-1 GTP-RAC-1 CK1 inhib DSH MO - - + - - - + - + + + XWNT-11 MO + + + 0 20 40 60 80 100 120 * %elongatedexplants elongation Vítězslav Bryja, 2014    Attachments      #7      Bryja V1 , Andersson ER1 , Schambony A1 , Esner M, Bryjová L, Biris KK, Hall AC, Kraft B,  Cajanek L, Yamaguchi TP, Buckingham M, Arenas E. (2009): The Extracellular Domain of  Lrp5/6 Inhibits Non‐Canonical Wnt Signaling in vivo. Mol Biol Cell. 20: 924‐936. Cover  article.  1  equal contribution      Impact factor (2009): 5.979  Times cited (without autocitations, WoS, Feb 21st 2014): 28  Significance: The first evidence of the fact that extracellular domain of Lrp6, a receptor  dedicated to Wnt/β‐catenin signaling can block non‐canonical Wnt signaling via  sequestration  of  Wnt  ligands  activating  non‐canonical  Wnt  pathway.  This  interaction is physiologically relevant and is demonstrated both in mouse and in  frog at several developmental processes.  Contibution  of  the  author/author´s  team:  Coordination  of  the  study.  Biochemical  analysis and generation/phenotype analysis of compound mouse mutants.                Molecular Biology of the Cell Vol. 20, 924–936, February 1, 2009 The Extracellular Domain of Lrp5/6 Inhibits Noncanonical Wnt Signaling In Vivo Vitezslav Bryja,*†‡ Emma R. Andersson,*‡ Alexandra Schambony,‡§ሻ Milan Esner,¶# Lenka Bryjova´,*† Kristin K. Biris,@ Anita C. Hall,*, ** Bianca Kraft,§ Lukas Cajanek,* Terry P. Yamaguchi,@ Margaret Buckingham,¶ and Ernest Arenas* *Laboratory of Molecular Neurobiology, Department of Medical Biochemistry and Biophysics, Center for Developmental Biology and Regenerative Medicine, Karolinska Institute, SE-171 77 Stockholm, Sweden; § Zoologisches Institut II, Universita¨t Karlsruhe (TH), D-76131 Karlsruhe, Germany; ¶ Centre National de la Recherche Scientifique Unite´ de Recherche Associe´e 2578, Department of Developmental Biology, Pasteur Institute, 75 724 Paris Cedex 15, France; @ Cancer and Developmental Biology Laboratory, Center for Cancer Research, National Cancer Institute-Frederick, National Institutes of Health, Frederick, MD 21702; and † Department of Cytokinetics, Institute of Biophysics, Academy of Sciences of the Czech Republic, 601 77, Brno, Czech Republic and Institute of Experimental Biology, Faculty of Science, Masaryk University, 602 00, Brno, Czech Republic Submitted July 11, 2008; Revised November 20, 2008; Accepted November 21, 2008 Monitoring Editor: Kunxin Luo Lrp5/6 are crucial coreceptors for Wnt/␤-catenin signaling, a pathway biochemically distinct from noncanonical Wnt signaling pathways. Here, we examined the possible participation of Lrp5/6 in noncanonical Wnt signaling. We found that Lrp6 physically interacts with Wnt5a, but that this does not lead to phosphorylation of Lrp6 or activation of the Wnt/␤-catenin pathway. Overexpression of Lrp6 blocks activation of the Wnt5a downstream target Rac1, and this effect is dependent on intact Lrp6 extracellular domains. These results suggested that the extracellular domain of Lrp6 inhibits noncanonical Wnt signaling in vitro. In vivo, Lrp6؊/؊ mice exhibited exencephaly and a heart phenotype. Surprisingly, these defects were rescued by deletion of Wnt5a, indicating that the phenotypes resulted from noncanonical Wnt gain-of-function. Similarly, Lrp5 and Lrp6 antisense morpholino-treated Xenopus embryos exhibited convergent extension and heart phenotypes that were rescued by knockdown of noncanonical XWnt5a and XWnt11. Thus, we provide evidence that the extracellular domains of Lrp5/6 behave as physiologically relevant inhibitors of noncanonical Wnt signaling during Xenopus and mouse development in vivo. INTRODUCTION Wnts are extracellular lipoglycoproteins that activate several downstream signaling pathways depending on cellular context. The best defined pathways include the canonical Wnt/ ␤-catenin pathway and the noncanonical Wnt/planar cell polarity (PCP) pathway. These two pathways regulate distinct biological processes. Certain components of Wnt signaling machinery are, based on current evidence, believed to be dedicated to only one of these two paths. Such components include Wnt ligands, receptors/coreceptors, and cytoplasmic components, in which Wnt1/Wnt3a, Lrp5/6 (XLRP5, and XLRP6 in Xenopus) and axin/APC/GSK3 are usually associated with the canonical Wnt pathway (Clevers, 2006) and Wnt5a/Wnt-11, Vangl1/2, Celsr1, and Rho/ Rho kinase/c-Jun NH2-terminal kinase are associated with the noncanonical Wnt pathway (Seifert and Mlodzik, 2007). Deficiency in the core components of the vertebrate Wnt/ PCP pathway results in defects in embryonic development that are different from the defects found in Wnt/␤-catenin pathway mutants (van Amerongen and Berns, 2006; Seifert and Mlodzik, 2007). The Wnt/PCP pathway is usually associated with the regulation of cell polarity and/or cell migration. In line with this, both gain-of-function (GOF) and lossof-function (LOF) in Wnt/PCP often produce similar/ identical phenotypes (Fanto and McNeill, 2004; Klein and Mlodzik, 2005; Schambony and Wedlich, 2007). It was suggested that membrane proteins and proteoglycans that act as Wnt coreceptors, e.g., Lrp5/6, Ror1/2, and Knypek (Wehrli et al., 2000; Topczewski et al., 2001; Oishi et al., 2003; Mikels and Nusse, 2006; Schambony and Wedlich, 2007), may be the factors deciding the predominant direction of Wnt signaling. A large body of evidence suggests that Lrp5 and Lrp6 are crucial coreceptors for Wnt/␤-catenin signaling (Pinson et al., 2000; Tamai et al., 2000; Wehrli et al., 2000). However, a recent report by Tahinci et al. (2007) suggests that the intracellular part of Lrp6 can also act as an This article was published online ahead of print in MBC in Press (http://www.molbiolcell.org/cgi/doi/10.1091/mbc.E08–07–0711) on December 3, 2008. ‡ These authors contributed equally to this work. Present addresses: ࿣ Developmental Biology Unit, Department of Biology, University of Erlangen-Nuremberg, Staudtstrasse 5, 91058 Erlangen, Germany; # Max-Planck Institute of Molecular Cell Biology and Genetics, Screening facility, Pfotenhauerstrasse 108, D-01307 Dresden, Germany; **Division of Cell and Molecular Biology, Imperial College London, United Kingdom. Address correspondence to: Vitezslav Bryja (bryja@sci.muni.cz) or Ernest Arenas (ernest.arenas@ki.se). 924 © 2009 by The American Society for Cell Biology inhibitor of noncanonical Wnt signaling in Xenopus. Yet, it is not clear whether there is a general role of endogenous Lrp5/6 as an inhibitor of noncanonical Wnt signaling, and how Lrp5/6 achieves its inhibitory action. Here, we performed in vitro and in vivo experiments, both in Xenopus and mouse, to further define the involvement of endogenous Lrp5/6 in noncanonical Wnt signaling. We show that Wnt5a physically interacts with Lrp6, and overexpression of Lrp6 inhibits the activity of the Rho GTPase Rac1. Moreover, Lrp5 and/or Lrp6 deficiency in Xenopus and mouse caused noncanonical Wnt gain of function (GOF) defects, which could be rescued by ablation of noncanonical Wnts. These data provide for the first time the evidence that extracellular parts of Lrp5/6 can sequester noncanonical Wnt ligands and act as physiologically relevant inhibitors of noncanonical Wnt signaling in multiple organs of Xenopus and mouse, including the heart and neural tube, during vertebrate development. MATERIALS AND METHODS Tissue Culture SN4741 cells were obtained from Dr. J. H. Son (Son et al., 1999). B1A and B1A overexpressing hemagglutinin (HA)-Wnt5a (HA-Wnt5a-B1A) fibroblasts were a kind gift of Jan Kitajewski (Columbia University, New York, NY; Shimizu et al., 1997). SN4741 and human embryonic kidney (HEK) 293 cells were grown and treated as described previously (Schulte et al., 2005; Bryja et al., 2007b). Wnt3a and Wnt5a (R&D Systems, Minneapolis, MN) were tested for activity using previously established protocols (Bryja et al., 2007a,b). Concentrations required for maximal activity (Dvl shift) varied from batch to batch. Western Blotting, Rac1 Activity Assay and Immunoglobulin G (IgG) Pull-Down Immunoblotting and sample preparation were done as published previously (Bryja et al., 2007a). When required, signal intensity was quantified using Scion Image densitometry software (Scion, Frederick, MD). Antibody details are given in Supplemental Material. Activity of Rac1, Rho, and Cdc42 was analyzed essentially as published previously (Unterseher et al., 2004). Briefly, glutathione transferase (GST)-p21-activated kinase (PAK)-CDC42/Rac interactive binding domain (CRIB), GST-Wiskott-Aldrich syndrome protein (WASP)-CRIB, GST-RHOtekin recombinant proteins were coupled to glutathione-Sepharose beads for detection of activated RAC-1, CDC42, and RHO A, respectively. Cells were washed with ice-cold phosphate-buffered saline (PBS) and subsequently allowed to lyse in ice-cold lysis buffer (10 mM Tris-Cl, pH 7.5, 110 mM NaCl, 1 mM EDTA, 10 mM MgCl2, 1% Triton X-100, 0.1% SDS, 20 mM ␤-glycerophosphate, 1 mM dithiothreitol, and complete protease inhibitors (Roche Molecular Biochemicals, Indianapolis, IN) for 5 min. Crude cell lysates were spun down in chilled tubes at 14,000 rpm for 5 min at 4°C. Supernatants (5% saved as input) were supplemented with bait proteins coupled to glutathione-Sepharose beads, and tubes were incubated rotating end-over-end at 4°C for 15 min. Beads were washed three times with washing buffer (lysis buffer without SDS and protease inhibitors) on ice and subsequently mixed with 2ϫ Laemmli buffer. Each sample was boiled (5 min) before loading on SDS-polyacrylamide gel electrophoresis (PAGE). For analysis of the interaction of Lrp6 and Wnt5a by IgG pull-down, B1A and HA-Wnt5a-B1A cells were transfected with Lipofectamine 2000 (Invitrogen, Carlsbad, CA) with a vector encoding the extracellular part of Lrp6 fused with human Fc fragment (Lpr6N-Fc; Tamai et al., 2000). After 2 d, culture media were collected and supplemented with 0.5% NP-40, and remaining cells were extracted for 15 min in ice-cold lysis buffer (50 mM Tris-HCl, pH 7.4, 150 mM sodium chloride, 0.5% NP-40, 1 mM EDTA, 1ϫ protease inhibitor cocktail [Roche Diagnostics]) to generate samples of conditioned media and cell lysate, respectively. Then samples were cleared by centrifugation at 15,000 ϫ g for 5 min at 4°C and incubated with protein G-coupled Sepharose beads (GE Healthcare, Little Chalfont, Buckinghamshire, United Kingdom) overnight. Next day, beads were washed with lysis buffer five times, mixed with 2ϫ Laemmli buffer, and subjected to SDS-PAGE. Frog Handling, Microinjections, and Keller Explants Embryos were obtained by in vitro fertilization, cultured, and injected as described previously (Unterseher et al., 2004). Embryos were injected at the four-cell stage in both dorsal blastomeres or at the eight-cell stage in one dorsal blastomere. If not indicated otherwise, injection amounts of plasmids were 60 pg of mLRP5 or mLRP6; 5 pg of constitutively active (ca) RhoA; 10 pg of ca Rac 1; 100 pg of ␤-catenin, XWnt5a, dominant-negative (dn) RhoA, or dn Rac 1; 20 pg of XWnt-11; and 100 pg of ␤-galactosidase; MOs were 0.8 pmol (all MOs: GeneTools, Philomath, OR; sequences are given in Supplemental Material). Keller open face explants for analysis of convergent extension movements were prepared at stage 10.5 and cultured, imaged, and scored as described previously (Unterseher et al., 2004). Statistical evaluation was performed using Student’s t test. Whole mount in situ hybridizations were carried out using the digoxigenin/alkaline phosphatase detection system (Roche Molecular Biochemicals) as described previously (Hollemann et al., 1999). Mouse Strains and Genotyping Lrp6 (Pinson et al., 2000) and Wnt5a (Yamaguchi et al., 1999a) mutant mice were housed, bred, and treated in accordance with the ethical approval for animal experimentation granted by Stockholms Norra Djurfo¨rso¨ks Etiska Na¨mnd. The genotyping was performed by polymerase chain reaction and is described in detail in Supplemental Material. Whole-Mount in Situ Hybridization (ISH) of Mouse Embryos The original cDNA clones described in the literature were used as templates for the generation of cRNA probes. Details are available upon request. Wholemount in situ hybridization was performed as described previously (Wilkinson and Nieto, 1993). Embryos were photographed on a stereoscope (Leica, Wetzlar, Germany) or an Axiophot (Carl Zeiss, Jena, Germany) compound microscope. Unless indicated otherwise, at least three mutant embryos were examined with each probe, and all yielded similar results. Heart Analysis in Mouse Before embedding, embryos were fixed in 4% paraformaldehyde overnight, and then they were incubated 12 h at 4°C in 15% sucrose in PBS. Embryos at stage embryonic day (E) 10.5 were embedded in 7.5% gelatin/sucrose and at stage E14.5 in Tissue-Tek OCT (Labonord, Villeneuve d’Ascq, France). Ten- to 16-␮m-thick sections were obtained using a cryostat. Slides were washed 2 ϫ 10 min in PBS at 37°C and stained for 30 min in 0.5% eosin solution (Labonord), progressively dehydrated to 100% ethanol, and mounted into Cytoseal (Richard Allan Scientific, Kalamazoo, MI). RESULTS Lrp6 Interacts with Wnt5a and Blocks Noncanonical Wnt Signaling in Vitro To analyze the involvement of Lrp6 in the noncanonical Wnt pathway, we treated SN4741 cells with Wnt5a and used Wnt3a (a canonical Wnt) for comparison. Treatment with either Wnt (100 ng/ml) led to the phosphorylation of Dvl2 and Dvl3, detected by a mobility shift of the protein on SDS-PAGE, as shown previously (Gonzalez-Sancho et al., 2004; Schulte et al., 2005). Although both Wnt3a and Wnt5a induced Dvl phosphorylation, only Wnt3a induced ␤-catenin activation and Lrp6 phosphorylation at Ser1490 (Tamai et al., 2004), as assessed by antibodies recognizing active ␤-catenin (ABC; the form of ␤-catenin dephosphorylated on Ser37 and Thr41; van Noort et al., 2002) or pSer1490-Lrp6 (Figure 1A). These data demonstrated the ability of Wnt3a, but not Wnt5a, to induce Wnt/␤-catenin signaling via phosphorylation of Lrp6. Wnt-induced phosphorylation of Lrp6 at Ser1490 was shown to recruit axin to Lrp6 and promote further downstream signaling to ␤-catenin (Tamai et al., 2004; Davidson et al., 2005; Zeng et al., 2005). Wnt5a failed to activate Lrp6 and the ␤-catenin pathway in SN4741 cells (Figure 1A), although it induced phosphorylation of Dvl. Such activation of Dvl was shown to be Lrp6-independent (Gonzalez-Sancho et al., 2004). Based on this finding, it was expected that Wnt5a should not interfere with Wnt3a-induced phosphorylation of Lrp6. However, in Wnt3a-treated (20 ng/ml) SN4741 cells, Wnt5a efficiently reduced, in a dose-dependent manner, theWnt3a-induced phosphorylation of Lrp6 at Ser1490 (Figure 1, B and C). These data suggested that Wnt5a directly or indirectly interfered with the phosphorylation of Lrp6 induced by Wnt3a. We therefore examined whether Wnt5a could bind or physically interact with Lrp6. Lrp5/6 in Noncanonical Wnt Signaling Vol. 20, February 1, 2009 925 To explore this possibility, we overexpressed the extracellular part of Lrp6 fused to the Fc fragment of human IgG (Lrp6N-Fc; Tamai et al., 2000) in B1A fibroblasts and in B1A fibroblasts overexpressing HA-tagged Wnt5a (HA-Wnt5a; Shimizu et al., 1997). Cell lysates and serum free-conditioned media were subjected to hFc (IgG) pull-down by incubation Figure 1. Wnt5a can bind Lrp6, inhibit Lrp6 phosphorylation and Lrp6 inhibits Rac1 activation. (A) SN4741 cells were stimulated with 100 ng/ml Wnt3a or Wnt5a. Stimulation with either of these led to the phosphorylation of Dvl2 and Dvl3 (indicated by open arrowheads), as assessed by a mobility shift in a Western blot. Only Wnt3a led to an increase in active ␤-catenin and phosphorylation of Lrp6 at Ser1490. (B) Wnt5a inhibits Wnt3a-induced phosphorylation of Lrp6 at Ser1490, but not Dvl2 phosphorylation, in a dose-dependant manner. Total Lrp6 did not significantly change by Wnt treatment. Actin was used as loading control. The level of Lrp6 phosphorylation from three independent experiments is quantified in C. (D) Lrp6 Fc can associate with HA-Wnt5a. Lrp6 Fc was overexpressed in HA-Wnt5a expressing B1A fibroblasts (B1A fibroblasts were used as a control). Total cell lysates (TCL) or conditioned media (CM) were subjected to IgG pull-down, and HA-Wnt5a was detected only in samples also expressing Lrp6-Fc. (E) Myc-tagged Lrp6 mutants lacking either E1 and E2 (Myc-Lrp6⌬E1-E2) or E3 and E4 (Myc-Lrp6⌬E3-E4) were overexpressed in B1A, or B1A cells stably expressing HA-tagged Wnt5a. Cells lysates were immunoprecipitated using antibody directed against HA-tag. Expression of Myc-Lrp6 and HA-Wnt5a in immunoprecipitates was determined by Western blotting. (F) HEK cells were transfected with Myc-Rac1 and indicated Lrp6 constructs. The activation of Rac1 was determined using Rac1 activation assay, and Western blot for Myc-Rac1. The signal ratio for GTP-Rac1/Rac1 was quantified and demonstrates that only full length Lrp6, but not mutants in any of the extracellular domains of Lrp6, inhibited Rac1 activity. V. Bryja et al. Molecular Biology of the Cell926 with protein G-Sepharose beads and subsequent Western blotting. As shown in Figure 1D, after IgG pull-down HAWnt5a was present only in the samples expressing Lrp6-Fc but not in any of the control conditions. Extracellular domains E1 and E2 (YWTD EGF repeats), but not E3 and E4, of Lrp6 are required for binding of Wnt5a to Lrp6 (Figure 1E) as shown previously for Wnts activating the Wnt/␤-catenin pathway (Mao et al., 2001). These results demonstrated that the extracellular part of Lrp6 can physically interact with Wnt5a. Because recombinant Wnt5a can induce the small Figure 2. Lrp5/6 are crucial regulators of convergent extension (CE) movements in Xenopus. (A and B) Injection of Lrp5 or Lrp6 mRNA or XLRP5/6 MOs all inhibit convergent extension of Keller explants from stage 10.5 that are rescued by mLrp5/6 but cannot be rescued by ␤-catenin coinjection (*, significant difference from control; **, significant rescue of MO, p Ͼ 0.95). (C and D) The CE defects induced by XLRP5 MO or XLRP6 MO are not rescued by Wnt5a or Wnt11 overexpression or constitutively active (ca) RhoA and Rac1. However, down-regulation of noncanonical signaling by XWnt5a or XWnt11 MO rescued XLRP5 and XLRP6 depletion phenotypes. XLRP5 MO induced inhibition of elongation was also rescued by dn RhoA and Rac1 (**, significant rescue of MO; p Ͼ 0.95). (E) Typical morphology of Keller explants injected with XLRP5 and XLRP6 MOs. The numbers under the graphs indicate number of injected embryos/number of independent experiments. Lrp5/6 in Noncanonical Wnt Signaling Vol. 20, February 1, 2009 927 Rho GTPase Rac1 (Andersson et al., 2008), a downstream component of the Wnt/PCP pathway, we next tested whether activity of Rac1 was Lrp6 dependent. As we show in Figure 1F, the overexpression of Lrp6 reduced the activity of Rac1 (Figure 1F). Importantly, Lrp6 mutants lacking either the Wnt-5a binding (⌬E1-E2; Figure 1E), or Dkk1 binding (⌬E3-E4) or the entire extracellular domain (⌬E1-E4) of Lrp6 (Mao et al., 2001) were not sufficient to reduce the activity of Rac1. These findings suggested that extracellular domains of Lrp5/6 may work as inhibitors of noncanonical Wnts and prompted us to test this hypothesis in vivo. XLRP5 Is Essential for Convergent Extension Movements in Xenopus In vertebrate embryos, ␤-catenin–independent Wnt pathways regulate several developmental processes, including convergent extension (CE) movements in the gastrulating Xenopus embryo. To determine the importance of Lrp5 and Lrp6 in Xenopus CE in vivo, we modulated the levels of Lrp5/6 by injecting mRNAs encoding mouse Lrp5 or Lrp6 or antisense morpholino-oligonucleotides directed against XLRP5 and XLRP6 (XLRP5 MO and XLRP6 MO, Supplemental Figure S1A) and examined Keller explants of the dorsal marginal zone. The scheme of timing of injections performed in Xenopus embryos and subsequent analysis is shown in Supplemental Figure S1B. We show in Figure 2A, in agreement with Tahinci et al. (2007) that XLRP5 MO strongly affected explant elongation, whereas XLRP6 MO had less effect. However, depletion of XLRP6 affected explant constriction (Figure 2B), which indicated defective noncanonical Wnt signaling (Unterseher et al., 2004; Schambony and Wedlich, 2007). Coinjection of both XLRP MOs blocked elongation similarly to XLRP5 MO alone. As expected, overexpression of mLrp5 or mLrp6 also affected explant elongation similarly (Figure 2A). Coinjection of XLRP5 MO or XLRP6 MO with mLRP5 or mLRP6 induced a partial Figure 3. (A) Embryos injected with mLRP5 or XLRP5 MO were lysed and analyzed for the activity of small GTPases Rac1 and Cdc42. (B) XWnt5a- or XWnt11-induced CE defects can be partially rescued by coinjection of Lrp5 or Lrp6 (**, significant rescue of XWnt-11 and XWnt-5a, respectively; p Ͼ 0.95). (C) Effects of FL and mutant hLRP6 on CE defects induced by XLRP5 MOs. hLRP6 and also hLRP6 lacking cytoplasmic domain (hLRP6⌬C) can efficiently rescue XLRP5 MO defects, whereas hLRP6 lacking the extracellular domains E1-E4 cannot (**, significant rescue of MO; p Ͼ 0.95). (D) Lrp6 lacking cytoplasmic domain (hLRP6⌬C) can rescue elongation defects caused by overexpression of XWnt-11 or XWnt-5a. The numbers under the graphs indicate number of injected embryos/number of independent experiments. V. Bryja et al. Molecular Biology of the Cell928 rescue of explant elongation and constriction, respectively (Figure 2, A and B), indicating partial redundancy of LRP5/6 and demonstrating that the effects of XLRP MOs were specific. Importantly, the effects of XLRP5 depletion seem to be ␤-catenin independent because coinjection of ␤-catenin RNA with XLRP5 MO did not rescue the phenotype (Figure 2A). It should be noted that, when injected into the dorsal marginal zone, XLRP5 MO induced other defects, which are associated with a ␤-catenin LOF, e.g., ventralization. We scored these defects by calculating the dorso-anterior index (Kao and Elinson, 1988) and found that XLRP5 MO-induced ventralization was rescued by coinjection of mLrp5 and to the same extent also by ␤-catenin (Supplemental Figure S1C), indicating that these phenotypes are the result of defective canonical Wnt signaling caused by XLRP5 depletion. To identify the molecular mechanism responsible for the observed phenotypes of XLRP5 knockdown in Xenopus embryos, we performed a set of rescue experiments. We hypothesized that if the XLRP5 knockdown caused an increase in noncanonical Wnt signaling, as our in vitro experiments suggested, negative but not positive regulators of noncanonical Wnt signaling should rescue the phenotype. Coinjection of XWnt5a, XWnt11, and ca forms of known downstream effectors of Wnt/PCP pathway—ca RhoA and ca Rac1 (Habas et al., 2003) did not rescue the phenotype (Figure 2C). Instead, inhibition of noncanonical Wnt pathways by knockdown of XWnt5a or XWnt11 or dn forms of Rac1 or RhoA were able to rescue XLRP5 MO. In XLRP6 MO, we observed that coinjection of XWnt-11 MO but not XWnt-5a MO restored constriction (Figure 2D). These and the previous experiments suggest that the depletion of XLRP5/6 induced a noncanonical Wnt GOF phenotype. These results indicate that Lrp5/6 interact with and sequester noncanonical Wnts, partially preventing their physiological activity and thereby inhibiting noncanonical signaling. To test the possibility that Lrp5/6 overexpression directly blocks the activity of small GTPases, we measured the activity of Rac1 and Cdc42, two GTPases that have been shown to be under the control of noncanonical Wnt signaling in Xenopus embryos (Habas et al., 2003; Penzo-Mendez et al., 2003; Kim and Han, 2005; Schambony and Wedlich, 2007). As we show in Figure 3A, overexpression of XLRP5 negatively regulates the activity of Rac1 and Cdc42. The XLRP5 MOs have a weak positive effect, which, especially in Cdc42, might be due to its effects on total Rac1/Cdc42. We were unable to detect significant differences in the activity of RhoA (data not shown). These results provide biochemical support for the rescue experiments and demonstrate that Lrp5 affects noncanonical Wnt signaling via modulation of the activity of small GTPases. To further test our model, we overexpressed XWnt11 and XWnt5a, which resulted in a strong inhibition of explant elongation by noncanonical GOF (Figure 3B). This inhibition of explant elongation by XWnt-11 or XWnt-5a overexpression was rescued by coinjection of either mLrp5 or mLrp6 (Figure 3A). To test which LRP-domains are required for its function in CE, we attempted to rescue the stronger XLRP5 knockdown phenotype with hLrp6 deletion mutants (Figure 3C). The mutant lacking the extracellular domains (⌬E1-E4; Mao et al., 2001) failed to rescue XLRP5 MO, whereas the mutant without cytoplasmic domain (Lrp6 ⌬C; Tamai et al., 2004) efficiently improved CE defects induced by XLRP5 MOs. Importantly, Lrp6 lacking the intracellular domain (Lrp6 ⌬C) was sufficient to rescue elongation defects caused by overexpression of XWnt-11 or XWnt-5a (Figure 3D). This data confirmed that Lrp5/6, and specifically the extracellular domains, in addition to playing a role in the Wnt/␤catenin pathway, are physiologically relevant inhibitors of noncanonical Wnt signaling. XWnt11 MO Rescues the Heart Phenotype Induced by XLRP5/6 MO in Xenopus embryos Noncanonical Wnts are known to regulate not only CE movements but also heart development in Xenopus (Pandur et al., 2002). To analyze the role of Lrp5/6 in heart development we investigated the expression of cardiac marker genes in embryos injected with XLRP MOs targeted to the presumptive cardiac mesoderm in eight–cell-stage embryos to avoid convergent extension and primary-axis defects (Supplemental Figure S1B). XLRP5 MO, and to a lesser extent XLRP6 MO, caused a down-regulation of the cardiac markers Nkx 2.5 and troponin (Figure 4A), two defects also observed in the XWnt-11 knockdown (Pandur et al., 2002). However, heart development is regulated by both canonical and noncanonical Wnt signaling (Brade et al., 2006) and both pathways could be potentially affected by XLRP5 depletion. To better distinguish the contribution of each pathway and to demonstrate specificity of the observed defects, we performed a set of rescue experiments. The cardiac markers Troponin and Nkx 2.5 were rescued by coinjection of mLRP5, which demonstrates the specificity of XLRP5 MO. However, only a partial rescue was observed after coinjection of ␤-catenin (Figure 4, B and C), suggesting that the XLRP5 knockdown phenotype is at least partially ␤-catenin independent. To test whether this XLRP5 MO cardiac phenotype is contributed to by noncanonical Wnt signaling, as shown previously for CE movements, we coinjected XWnt11 MO. We observed a partial rescue of the XLRP5 MO cardiac phenotype by XWnt11 MO (Figure 4, B and C), suggesting that XLRP5 MO induces a noncanonical Wnt GOF as shown for CE movements above. In summary, these data demonstrate that endogenous XLRP5/6 regulates not only canonical Wnt signaling but also CE movements and heart development, two processes driven by noncanonical Wnt signaling in Xenopus. Furthermore, our results suggest that the inhibition of noncanonical Wnt signaling by Lrp5/6 is achieved by an interaction of Lrp5/6 with noncanonical Wnts, which may reduce the availability of noncanonical Wnts for signaling. Lrp6-deficient Mice Display Heart Malformations That Are Partially Rescued by Deletion of Wnt5a Based on our analysis of Wnt signaling in vitro and in vivo, in Xenopus embryos, we hypothesized that Lrp5 or Lrp6 should also interact with noncanonical Wnts and regulate heart development in mice. To test this hypothesis, we analyzed heart morphology at E10.5 and E14.5 in Lrp6 and Wnt5a mutant mice (Yamaguchi et al., 1999a; Pinson et al., 2000). At E10.5, no obvious defects in heart morphology were detected, although Lrp6Ϫ/Ϫ hearts were a little bit smaller (Supplemental Figure S2), likely reflecting the overall smaller size of Lrp6 mutant embryos (Pinson et al., 2000). However, at E14.5 major outflow tract deformities were observed in either Lrp6 or Wnt5a mutants (Figure 5, A and B). Wnt5a mutants exhibited persistent truncus arteriosus (PTA) and the right atrium was expanded, whereas in Lrp6 mutants the great arteries were separated, resulting in a transposition of the great arteries. Analysis of Wnt5a single mutants revealed not only defects in intraventricular septation, with PTA, but also ventricular septal defects (VSD). Defects in the closure of the intraventricular septum (IVS) were observed in two of three analyzed embryos. The myocardium in the Wnt5a mutants Lrp5/6 in Noncanonical Wnt Signaling Vol. 20, February 1, 2009 929 was abnormally thin, the right atria were expanded, and the pericardium showed an abnormal morphology compared with wild type. In Lrp6 single mutants, we also observed VSD, the septum seemed porous, and the myocardium abnormally thin. No pericardial deformities were found in the Lrp6 mutants. In summary, at E14.5, both Wnt5a and Lrp6 Figure 4. XLRP5/6 regulate heart development. (A) Knockdown of XLrp5 and XLrp6 affects heart development. Embryos injected with XLRP5 and XLRP6 MO in one or both dorsal blastomeres were analyzed by whole mount ISH for the expression of heart markers Nkx2.5 and Troponin lc (Tnlc) at stage 28. Injected side shown by blue circle in single-blastomere injections, open triangle shows Tnlc/Nkx2.5 on injected side, closed triangle shows TnIc/Nkx2.5 on uninjected side. (B) XLRP5 MO-induced reduction in heart markers can be rescued with mLrp5 and XWnt-11 MO (Tnlc) and only partially rescued with ␤-catenin overexpression (Nkx2.5). (C) Quantification of ISH analysis of cardiac markers (*, significantly differs from ␤-galactosidase controls; p Ͼ 0.95); **, significantly differs from XLRP5MO; p Ͼ 0.95). V. Bryja et al. Molecular Biology of the Cell930 single mutants displayed arterial pole defects. Similar defects were observed in other mutants of the noncanonical signaling pathway, such as that for Vang-like 2 factor (Henderson et al., 2006) and Dvl2 mutants (Hamblet et al., 2002). These data suggest that both Lrp6 and Wnt5a are necessary for proper heart development in mouse. Based on the previous results we concluded that the heart phenotype of Lrp6 mutants could in part be due to an excess of noncanonical signaling. We therefore decided to analyze the effect of removing one allele of Wnt5a at a later stage of heart development (E14.5) in Lrp6 mutants, expecting that an excess of noncanonical signaling in Lrp6 mutants may be mitigated by reducingWnt5a. Strikingly, analysis of Lrp6 null mice in which one allele of Wnt5a had been removed led to a partial or complete rescue of the heart defects seen in Lrp6 single mutants. Specifically, two of four compound Wnt5aϩ/Ϫ Lrp6Ϫ/Ϫ embryos did not display any heart defects at E14.5 (Figure 5C), and the heart defects in the two remaining compound Wnt5aϩ/Ϫ Lrp6Ϫ/Ϫ embryos showed a marked improvement in myocardium and IVS, and the phenotype was less severe than that in Lrp6Ϫ/Ϫ mice (Figure 5B). E14.5 Wnt5aϪ/Ϫ Lrp6ϩ/Ϫ embryos showed identical heart defects to Wnt5aϪ/Ϫ Lrp6ϩ/ϩ (Figure 5, A and B). Our analysis of heart development in mice thus supports and extends our findings obtained in Xenopus. Together, these data indicate that Lrp6 deficiency causes noncanonical Wnt pathway GOF phenotypes and that Lrp6 can inhibit noncanonical Wnt signaling in vivo. Furthermore these findings underline the importance of Wnt/PCP signaling in heart development and particularly in outflow tract morphogenesis. Deletion of Wnt5a Completely Rescues Exencephaly in Lrp6 Mutant Mice To define in more detail the functional interaction between Lrp6- and Wnt5a-driven pathways in vivo and to ascertain whether that interaction is more widespread than anticipated previously, we studied the neural phenotype of compound Wnt5a and Lrp6 mutants. Both Wnt5a and Lrp6 deficiencies are embryonal lethal and we thus crossed Wnt5aϩ/Ϫ mice with Lrp6ϩ/Ϫ mice to generate double heterozygous Wnt5aϩ/Ϫ Lrp6ϩ/Ϫ animals. These double Figure 5. Lrp6 mutant mice exhibit heart defects, which can be rescued by Wnt-5a heterozygosity. (A) Heart morphology in Lrp6 and Wnt5a mutants at E14.5. Heart of E14.5 wild type, Lrp6Ϫ/Ϫ and Wnt5aϪ/Ϫ mouse embryos was dissected and the morphology was analyzed. Representative examples are shown. ao, aorta; pt, pulmonary artery; PTA, persistent truncus arteriosus. (B) Hearts of embryos (E14.5) with indicated genotypes were sectioned and stained with hematoxylin/eosin. Typical heart defects are indicated by arrows. peric., pericardium; pIVS, porous intraventricular septum; VSD, ventricular septal defects; RA, right atrium. (C) Quantification of arterial pole deformities of embryos in individual genotypes. Lrp5/6 in Noncanonical Wnt Signaling Vol. 20, February 1, 2009 931 heterozygous animals were born with the expected Mendelian frequency (25.13%, expected 25%), and we did not notice any gross morphological, fertility, or behavioral defects in comparison with parental Wnt5aϩ/Ϫ or Lrp6ϩ/Ϫ mutants. A small proportion (Ͻ5%) of Wnt5a ϩ/Ϫ Lrp6ϩ/Ϫ mice showed tail deformities, which were also present at comparable frequency in Lrp6ϩ/Ϫ mice. After crossing of double-heterozygous mice, we recovered Lrp6Ϫ/ϪWnt5aϪ/Ϫ embryos at E10.5 at the expected frequency. All Lrp6Ϫ/ϪWnt5aϪ/Ϫ double null embryos were severely developmentally delayed (Figure 6A), and we did not obtain any Wnt5a/Lrp6 double knockout embryos at E12.5. We observed exencephaly (neural tube completely open in cephalic region) in ϳ30% of Wnt5aϩ/ϩLrp6Ϫ/Ϫ embryos at E10.5 (Figure 6A, open arrow) and in 25% of Lrp6Ϫ/Ϫ embryos at E12.5 (n ϭ 20). Neural tube closure defects such as exencephaly are usually associated with aberrant CE movements and are observed in several other mutants of the Wnt/PCP pathway components such as Dvl2, Vangl2, and Wnt5a-deficient mice (Torban et al., 2004; Wang et al., 2006; Qian et al., 2007). Previous studies directly proved the importance of Wnt5a for neural tube closure in mouse (Qian et al., 2007) and demonstrated that similar defects can also be caused by GOF of noncanonical Wnt ligands (Shariatmadari et al., 2005). Interestingly, the proportion of Lrp6 mutants with exencephaly can be efficiently reduced by Wnt5a heterozygosity and in the double Wnt5a/Lrp6-deficient embryos we observed a complete rescue of exencephaly (Figure 6B). This observation is in good agreement with our in vitro findings, analyses of CE in Xenopus and heart development in Xenopus and mouse. Thus, our results provide evidence that Lrp5/6 deficiency results in noncanonical Wnt GOF defects and attributes Lrp5/6 a role as an inhibitor of noncanonical Wnt signaling in multiple vertebrate systems. Wnt5a and Lrp6 Cooperate in Early Mouse Embryonic Development Although it is known that Lrp5/6 is required as a coreceptor for canonical Wnt signaling and we hereby report that Lrp6 can inhibit noncanonical Wnt signaling, we also found evidence that these are not the only two modalities of interaction between Lrp6 and Wnt5a. We noticed that some of the compound Wnt5a/Lrp6 mutants displayed intermediate phenotypes at E10.5, and so we analyzed allelic combinations of Wnt5a and Lrp6. The analyzed phenotypic features included general developmental delay (defined as an embryo Ͻ40% size of WT littermates) and a lack of embryo turning, which usually takes place at E9.0. At E10.5, Lrp6Ϫ/ϪWnt5aϪ/Ϫ embryos showed a complex phenotype that included a developmental delay/lack of turning with almost complete penetrance. A similar phenotype, but less penetrant, was observed in Lrp6Ϫ/ϪWnt5aϩ/Ϫ embryos (Figure 6C). These data suggest that in addition to Wnt/PCP pathway defects (heart deformities and exencephaly), which were rescued by Wnt-5a deficiency, other mechanistically unrelated interactions exist between Lrp6- and Wnt5a-driven pathways during early morphogenesis. We hypothesized that such defects could result from the combination of defects in the Wnt/␤catenin pathway of the Lrp6Ϫ/Ϫ mutants and the Wnt/PCP pathway of Wnt5aϪ/Ϫ mutants resulting in additive or synergistic effects in other developmental processes such as embryonic turning. Analysis of embryos with various combinations of Lrp6 and Wnt5a null alleles at E8.5 (Figure 7) confirmed that Lrp6 deficiency resulted in decreased levels of genes directly or indirectly regulated by ␤-catenin, such as mesogenin (msgn), brachyury (T), and Mesp2 (Yamaguchi et al., 1999b). The expression of these genes is abolished in Wnt3a mutants, and in conditional ␤-catenin LOF mutants (Dunty et al., 2008). In contrast, Wnt5aϪ/Ϫ embryos did not show any alterations in the expression of ␤-catenin target genes but showed smaller somites as a result of affected CE movements (Figure 7A). Compound Lrp6Ϫ/ϪWnt5aϪ/Ϫ mutants showed an additive phenotype, with no evidence for further perturbation of ␤-catenin target gene expression (in comparison to the Lrp6 null) or of somite condensation (in comparison to the Wnt5a null). Thus, our analysis suggests that the early phenotype of Wnt5a; Lrp6 double mutants (growth retardation and lack of embryo turning) results from defects in the regulation of common biological process by the Wnt/␤catenin and Wnt/PCP pathways, rather than a direct interaction between the two signaling pathways. DISCUSSION It is well documented that Lrp5/6 are coreceptors necessary for Wnt signal transduction toward ␤-catenin (Pinson et al., Figure 6. Analysis of Wnt5a;Lrp6 double null mice at E10.5: Wnt5a deficiency rescues exencephaly but promotes general retardation of Lrp6Ϫ/Ϫembryos. (A) Wnt5a and Lrp6 single mutants occur as described previously at E10.5, with some Lrp6 embryos exhibiting exencephaly (arrow), compound Wnt5aϪ/Ϫ Lrp6Ϫ/Ϫ display developmental delay and fail to undergo turning. (B) Loss of Wnt5a in Lrp6 null mice rescued exencephaly. Exencephaly was observed in 30% of Lrp6 null mutants, this frequency was partially rescued by loss of one allele of Wnt5a and completely rescued by loss of both Wnt5a alleles. (C) Developmental delay and a failure to undergo turning was only observed in 10% of Lrp6Ϫ/Ϫ mice but was exacerbated by loss of Wnt-5a and occurred in almost in 100% of the Wnt5aϪ/Ϫ Lrp6Ϫ/Ϫ double knockout mice. V. Bryja et al. Molecular Biology of the Cell932 2000; Tamai et al., 2000; Wehrli et al., 2000). Here, we demonstrate an additional role for the extracellular domain of Lrp5/6 as an inhibitor of noncanonical Wnt signal transduction. Based on a series of experiments in vitro and in vivo, in Xenopus and mouse, we suggest that Lrp5/6 is a physiologically relevant inhibitor of noncanonical Wnt signaling, which acts by interacting with and reducing the availability of noncanonical Wnts for signaling. This conclusion is based on the following critical pieces of evidence: 1) Wnt5a physically interacts with Lrp6; 2) Lrp6 overexpression blocks activation of Rac1 in vitro; 3) Lrp5 and/or Lrp6 deficiency in Xenopus or mouse results in ␤-catenin–independent defects in CE and heart development; 4) these defects can be partially rescued by depletion of noncanonical Wnt ligands both in Xenopus and in mouse; and 5) Lrp6 deficiency in mice results in exencephaly, which is completely rescued by deletion of Wnt5a. Can our findings in XLRP5/6 depleted Xenopus or Lrp6deficient mouse embryos be explained simply as ␤-catenin lack-of-function phenotypes influenced by the loss of Wnt5a and/or Wnt11 rather than Lrp5/6 regulating noncanonical Wnt signaling? First, although inhibition of ␤-catenin signaling has been reported to block CE in Xenopus (Kuhl et al., 2001), we were not able to rescue XLRP5/6 MO phenotypes with ␤-catenin in Keller explants but rather rescued the phenotype with dn Rac1 or dn RhoA, suggesting that the phenotype is caused by a noncanonical GOF. In contrast, the overexpression of XWnt-8 (Kuhl et al., 2001) and ␤-catenin (Schambony, unpublished observation) has no negative effect on CE, but we observed CE defects after overexpression of mLrp5 and mLrp6. Consistently, overexpression of hLrp6 ⌬E1-E4, which acts as a constitutive activator of canonical Wnt-signaling, had no negative effect on CE (data not shown). Moreover, the effects of XLRP5 MO on dorsal axis formation, which is a well-defined ␤-catenin-dependent process, were almost abolished by coinjection of ␤-catenin. Thus, our results suggest that the mechanisms of CE disruption by blocking ␤-catenin signaling and by XLRP5/6 knockdown are distinct. Second, the analysis of Wnt5a mutant mice either alone or in combination with Lrp6 deletions did not show any alteration (either negative or positive) in the level of ␤-catenin target genes. Moreover, Wnt5a did not activate canonical Wnt signaling in vitro. Thus, Wnt5a deficiency or treatment does not affect ␤-catenin target genes despite the fact that selected ␤-catenin target genes are expressed in the Wnt5a expression domain (Yamaguchi et al., Figure 7. Analysis of ␤-catenin target genes in Wnt5a, Lrp6 compound mutants: Wnt-5a deficiency does not affect the expression of ␤-catenin target genes and Lrp6 deficiency does not affect the somite compression. (A) Embryos with indicated genotype were analyzed by two-color whole-mount ISH at E8.5. Somite marker Uncx4.1 is in red and ␤-catenin target gene mesogenin is in purple. (B) The expression of ␤-catenin target gene brachyury (T) is in red and a marker of segmentation clock Mesp2 is shown in purple. Lrp5/6 in Noncanonical Wnt Signaling Vol. 20, February 1, 2009 933 1999a). Wnt5a has also been shown to interfere with the ␤-catenin pathway, but such inhibition requires activation of canonical Wnt signaling and/or specific cellular and receptor contexts (He et al., 1997; Tao et al., 2005; Mikels and Nusse, 2006; Bryja et al., 2007b; Kofron et al., 2007). Thus, our results support the idea that deletion of Lrp6 or knockdown of Lrp5 results not only in ␤-catenin-dependent phenotypes, as described previously, but also in a disinhibition of noncanonical Wnt signaling that results in diverse noncanonical GOF phenotypes such as defects in CE movements, cardiac outflow tract morphogenesis, and neural tube closure. During cardiac development, canonical Wnt-signaling is required in the very early phase of cardiac precursor specification in the mesoderm (Naito et al., 2006; Ueno et al., 2007), followed by an inhibition of the same pathway by DKK1 and crescent (Marvin et al., 2001; Schneider and Nimpf, 2003) and activation of noncanonical pathways by Wnt-11 as shown by GOF and LOF (Eisenberg and Eisenberg, 1999; Pandur et al., 2002; Garriock et al., 2005; Ueno et al., 2007). Why does the increase in noncanonical Wnt signaling upon XLRP5 knockdown result in a loss of cardiac marker genes? Because both GOF and LOF of noncanonical Wnt signaling result in impaired cell polarity, deficits in migration of cardiac precursors to the anterior ventral side of the embryo would alter the exposure of cardiac precursors to inductive signals and therefore result in the loss of cardiac markers. This interpretation is also consistent with a recently proposed model of cardiac precursor migration and specification (Eisenberg and Eisenberg, 2006) and the regulation of cell–cell adhesion by noncanonical Wnts (Brade et al., 2006). It is believed that Lrp5/6 provides specificity to Wnt signaling by directing signaling to the ␤-catenin pathway (Mikels and Nusse, 2006). Although not empirically tested, binding specificity for canonical Wnts was thought to be a part of that mechanism (He et al., 2004). The physical interaction of Lrp6 and Wnt5a that we report seems to contradict that view. Our experiments demonstrate that Wnt5a can bind the extracellular soluble domain of Lrp6, although Wnt5a (in contrast to Wnt3a) cannot induce Ser1490-phosphorylation of Lrp6 or ␤-catenin activation. It is not clear whether the interaction between Lrp5/6 and Wnt5a is direct or mediated by other protein/proteoglycans such as the Wnt/PCP pathway component Knypek (Topczewski et al., 2001), which can bind Dkk1 (Caneparo et al., 2007), a highaffinity ligand for Lrp5/6 (Mao et al., 2001). Thus, our data reveal another level of complexity in Wnt signaling because Lrp5/6 binds a larger array of Wnts than previously anticipated, but only a subset of these ligands can activate specific Lrp6 phosphorylation and downstream signaling to ␤-catenin. Our functional in vivo studies suggest that Lrp5/6 (in Xenopus) and Lrp6 (in mouse) are negative regulators of noncanonical Wnt signaling. When this study was prepared for publication, Tahinci et al. (2007) reported that a stretch of 36 amino acids from the cytoplasmic domain of Lrp6 can block CE in Xenopus. Our data support the fact that Lrp5/6 inhibit CE via activation of noncanonical Wnt signaling, and they suggest that an additional mechanism of inhibition exists. Our experiments showed that mutations in the extracellular domain strongly interfere with the ability of Lrp6 to either block Rac1 or rescue CE defects caused by XLRP5 knockdown, whereas the membrane-tethered extracellular domain of Lrp6 (Lrp6⌬C) behaved as full-length Lrp6. This suggests that the mechanism of Lrp5/6 function in noncanonical Wnt signaling involves formation of complexes on the cell surface, which sequester noncanonical Wnts ligands and prevent their interaction with signaling receptors such as Frizzleds. The effects of Lrp6 on Rac1 activation require both E1ϩE2 Wnt-binding domains, and E3ϩE4 Dkk1-binding domains, pointing out the importance of physical interaction of noncanonical Wnts and Lrp5/6 for the function of Lrp5/6 in noncanonical signaling. This suggests that extracellular complexes organized by Lrp5/6 and inhibiting noncanonical Wnt signaling contain several components, e.g., Wnts, Dkk, and/or membrane glycoproteins such as Knypek (Caneparo et al., 2007). Our data support a model, where the availability of noncanonical Wnts and the degree of noncanonical Wnt signaling seems to be determined under physiological conditions by the molecular ratio of noncanonical Wnts and free Lrp5/6 available for binding. Indeed, developmental defects caused by the excess of noncanonical Wnt5a and Wnt11 were efficiently rescued by overexpression of Lrp5 or Lrp6, and developmental defects caused by the reduction of Wnt5a or Wnt11 were rescued by knockdown of Lrp5. In agreement with our model, a recent study by Caneparo and colleagues demonstrated that the Lrp5/6 ligand, Dkk1, can act as a ␤-catenin–independent positive regulator of Wnt/PCP pathway in vivo (Caneparo et al., 2007). Importantly, we hereby demonstrate that the interaction between Lrp5/6 and noncanonical Wnt ligands is physiologically important and regulates normal development. In summary, we provide several lines of evidence, both in vitro and in vivo, in Xenopus and mouse that define Lrp5 and Lrp6 as general inhibitors of the noncanonical Wnt signaling pathway. Moroever, our results indicate that Lrp5/6 receptors, by binding canonical or noncanonical Wnts, determine not only the level of Wnt signaling through Wnt/␤-catenin but also through noncanonical branches of Wnt signaling. ACKNOWLEDGMENTS We thank Andrew McMahon (Harvard University, Boston, MA) and William Skarnes (Sanger Institute, Cambridge, United Kingdom) for providing Wnt5a and Lrp6 mutant mice. We also thank Jan Kitajewski (Columbia University, New York, NY) for HA-Wnt5a-B1A cells; Dr. J. H. Son (Columbia University, New York, NY) for SN4741 cells; Christoph Niehrs (German Cancer Research Center, Heidelberg, Germany) for vectors encoding hLrp6 and its extracellular deletions; Mikhail Semenov and Xi He (Harvard Medical School) for vectors encoding mLrp5, mLrp6, Lrp6⌬C, and Lrp6-Fc; and Paul Krieg (University of Arizona, Tucson, AZ) for Xenopus Nkx2.5 and Troponin ISH probes. We would also like to thank Michael Ku¨hl (University of Ulm, Ulm, Germany), Sigolene Meilhac (Pasteur Institute, Paris, France), and the Arenas laboratory for critical reading of the manuscript and valuable discussions and Cecilia Olsson, Annika Ka¨ller, Claudia Winter, Caroline Berger, and C. Cimper for excellent technical and secretarial assistance. V. B. and M. E. were financed by the “EuroStemCell” project. This work was supported by European Molecular Biology Organization Installation Grant, Academy of Sciences of the Czech Republic (AVOZ50040507 and AVOZ50040702) and Ministry of Education, Youth and Sports of the Czech Republic (MSM0021622430) to V. B.; Swedish Foundation for Strategic Research, Swedish Royal Academy of Sciences, Knut and Alice Wallenberg Foundation, European Commission (Eurostemcell), Swedish Medical Research Council and Karolinska Institutet to E. A.; the Pasteur Institute, the Centre National de la Recherche Scientifique, the European Union Integrated Project Heart Repair to M. B., and the German Research Foundation (Scha965/2-3) to A. S. REFERENCES Andersson, E. R., Prakash, N., Cajanek, L., Minina, E., Bryja, V., Bryjova, L., Yamaguchi, T. P., Hall, A. C., Wurst, W., and Arenas, E. (2008). Wnt5a regulates ventral midbrain morphogenesis and the development of A9–A10 dopaminergic cells in vivo. PLoS ONE 3, e3517. Brade, T., Manner, J., and Kuhl, M. (2006). The role of Wnt signalling in cardiac development and tissue remodelling in the mature heart. Cardiovasc. Res. 72, 198–209. Bryja, V., Schulte, G., and Arenas, E. (2007a). Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate ␤-catenin. Cell Signal. 19, 610–616. V. Bryja et al. Molecular Biology of the Cell934 Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007b). Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J. Cell Sci. 120, 586–595. Caneparo, L., Huang, Y. L., Staudt, N., Tada, M., Ahrendt, R., Kazanskaya, O., Niehrs, C., and Houart, C. (2007). Dickkopf-1 regulates gastrulation movements by coordinated modulation of Wnt/beta catenin and Wnt/PCP activities, through interaction with the Dally-like homolog Knypek. Genes Dev. 21, 465–480. Castelo-Branco, G., Wagner, J., Rodriguez, F. J., Kele, J., Sousa, K., Rawal, N., Pasolli, H. A., Fuchs, E., Kitajewski, J., and Arenas, E. (2003). Differential regulation of midbrain dopaminergic neuron development by Wnt-1, Wnt-3a, and Wnt-5a. Proc. Natl. Acad. Sci. USA 100, 12747–12752. Clevers, H. (2006). Wnt/beta-catenin signaling in development and disease. Cell 127, 469–480. Davidson, G., Wu, W., Shen, J., Bilic, J., Fenger, U., Stannek, P., Glinka, A., and Niehrs, C. (2005). Casein kinase 1 gamma couples Wnt receptor activation to cytoplasmic signal transduction. Nature 438, 867–872. Dunty, W. C., Jr., Biris, K. K., Chalamalasetty, R. B., Taketo, M. M., Lewandoski, M., and Yamaguchi, T. P. (2008). Wnt3a/beta-catenin signaling controls posterior body development by coordinating mesoderm formation and segmentation. Development 135, 85–94. Eisenberg, C. A., and Eisenberg, L. M. (1999). WNT11 promotes cardiac tissue formation of early mesoderm. Dev. Dyn. 216, 45–58. Eisenberg, L. M., and Eisenberg, C. A. (2006). Wnt signal transduction and the formation of the myocardium. Dev. Biol. 293, 305–315. Fanto, M., and McNeill, H. (2004). Planar polarity from flies to vertebrates. J. Cell Sci. 117, 527–533. Garriock, R. J., D’Agostino, S. L., Pilcher, K. C., and Krieg, P. A. (2005). Wnt11-R, a protein closely related to mammalian Wnt11, is required for heart morphogenesis in Xenopus. Dev. Biol. 279, 179–192. Gonzalez-Sancho, J. M., Brennan, K. R., Castelo-Soccio, L. A., and Brown, A. M. (2004). Wnt proteins induce dishevelled phosphorylation via an LRP5/ 6-independent mechanism, irrespective of their ability to stabilize beta-catenin. Mol. Cell. Biol. 24, 4757–4768. Habas, R., Dawid, I. B., and He, X. (2003). Coactivation of Rac and Rho by Wnt/Frizzled signaling is required for vertebrate gastrulation. Genes Dev. 17, 295–309. Hamblet, N. S., Lijam, N., Ruiz-Lozano, P., Wang, J., Yang, Y., Luo, Z., Mei, L., Chien, K. R., Sussman, D. J., and Wynshaw-Boris, A. (2002). Dishevelled 2 is essential for cardiac outflow tract development, somite segmentation and neural tube closure. Development 129, 5827–5838. He, X., Saint-Jeannet, J. P., Wang, Y., Nathans, J., Dawid, I., and Varmus, H. (1997). A member of the Frizzled protein family mediating axis induction by Wnt-5A. Science 275, 1652–1654. He, X., Semenov, M., Tamai, K., and Zeng, X. (2004). LDL receptor-related proteins 5 and 6 in Wnt/beta-catenin signaling: arrows point the way. Development 131, 1663–1677. Henderson, D. J., Phillips, H. M., and Chaudhry, B. (2006). Vang-like 2 and noncanonical Wnt signaling in outflow tract development. Trends Cardiovasc. Med. 16, 38–45. Hollemann, T., Panitz, F., and Pieler, T. (1999). In situ hybridization techniques with Xenopus embryos. In: A Comparative Methods Approach to the Study of Oocytes and Embryos, ed. J. D. Richter, Oxford, United Kingdom: Oxford University Press. Kao, K. R., and Elinson, R. P. (1988). The entire mesodermal mantle behaves as Spemann’s organizer in dorsoanterior enhanced Xenopus laevis embryos. Dev. Biol. 127, 64–77. Kim, G. H., and Han, J. K. (2005). JNK and ROKalpha function in the noncanonical Wnt/RhoA signaling pathway to regulate Xenopus convergent extension movements. Dev. Dyn. 232, 958–968. Klein, T. J., and Mlodzik, M. (2005). Planar cell polarization: an emerging model points in the right direction. Annu. Rev. Cell Dev. Biol. 21, 155–176. Kofron, M., Birsoy, B., Houston, D., Tao, Q., Wylie, C., and Heasman, J. (2007). Wnt11/beta-catenin signaling in both oocytes and early embryos acts through LRP6-mediated regulation of axin. Development 134, 503–513. Kuhl, M., Geis, K., Sheldahl, L. C., Pukrop, T., Moon, R. T., and Wedlich, D. (2001). Antagonistic regulation of convergent extension movements in Xenopus by Wnt/beta-catenin and Wnt/Ca2ϩ signaling. Mech. Dev. 106, 61–76. Mao, B., Wu, W., Li, Y., Hoppe, D., Stannek, P., Glinka, A., and Niehrs, C. (2001). LDL-receptor-related protein 6 is a receptor for Dickkopf proteins. Nature 411, 321–325. Marvin, M. J., Di Rocco, G., Gardiner, A., Bush, S. M., and Lassar, A. B. (2001). Inhibition of Wnt activity induces heart formation from posterior mesoderm. Genes Dev. 15, 316–327. 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. Naito, A. T., Shiojima, I., Akazawa, H., Hidaka, K., Morisaki, T., Kikuchi, A., and Komuro, I. (2006). Developmental stage-specific biphasic roles of Wnt/ beta-catenin signaling in cardiomyogenesis and hematopoiesis. Proc. Natl. Acad. Sci. USA 103, 19812–19817. Oishi, I., et al. (2003). The receptor tyrosine kinase Ror2 is involved in noncanonical Wnt5a/JNK signalling pathway. Genes Cells 8, 645–654. Pandur, P., Lasche, M., Eisenberg, L. M., and Kuhl, M. (2002). Wnt-11 activation of a non-canonical Wnt signalling pathway is required for cardiogenesis. Nature 418, 636–641. Penzo-Mendez, A., Umbhauer, M., Djiane, A., Boucaut, J. C., and Riou, J. F. (2003). Activation of Gbetagamma signaling downstream of Wnt-11/Xfz7 regulates Cdc42 activity during Xenopus gastrulation. Dev. Biol. 257, 302–314. Pinson, K. I., Brennan, J., Monkley, S., Avery, B. J., and Skarnes, W. C. (2000). An LDL-receptor-related protein mediates Wnt signalling in mice. Nature 407, 535–538. Qian, D., Jones, C., Rzadzinska, A., Mark, S., Zhang, X., Steel, K. P., Dai, X., and Chen, P. (2007). Wnt5a functions in planar cell polarity regulation in mice. Dev. Biol. 306, 121–133. Schambony, A., and Wedlich, D. (2007). Wnt-5A/Ror2 regulate expression of XPAPC through an alternative noncanonical signaling pathway. Dev. Cell 12, 779–792. Schneider, W. J., and Nimpf, J. (2003). LDL receptor relatives at the crossroad of endocytosis and signaling. Cell Mol Life Sci. 60, 892–903. Schulte, G., Bryja, V., Rawal, N., Castelo-Branco, G., Sousa, K. M., and Arenas, E. (2005). Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation. J. Neurochem. 92, 1550–1553. Seifert, J. R., and Mlodzik, M. (2007). Frizzled/PCP signalling: a conserved mechanism regulating cell polarity and directed motility. Nat. Rev. Genet 8, 126–138. Shariatmadari, M., Peyronnet, J., Papachristou, P., Horn, Z., Sousa, K. M., Arenas, E., and Ringstedt, T. (2005). Increased Wnt levels in the neural tube impair the function of adherens junctions during neurulation. Mol. Cell Neurosci. 30, 437–451. Shimizu, H., Julius, M. A., Giarre, M., Zheng, Z., Brown, A. M., and Kitajewski, J. (1997). Transformation by Wnt family proteins correlates with regulation of beta-catenin. Cell Growth Differ. 8, 1349–1358. Son, J. H., Chun, H. S., Joh, T. H., Cho, S., Conti, B., and Lee, J. W. (1999). Neuroprotection and neuronal differentiation studies using substantia nigra dopaminergic cells derived from transgenic mouse embryos. J. Neurosci. 19, 10–20. Tahinci, E., Thorne, C. A., Franklin, J. L., Salic, A., Christian, K. M., Lee, L. A., Coffey, R. J., and Lee, E. (2007). Lrp6 is required for convergent extension during Xenopus gastrulation. Development 134, 4095–4106. Tamai, K., Semenov, M., Kato, Y., Spokony, R., Liu, C., Katsuyama, Y., Hess, F., Saint-Jeannet, J. P., and He, X. (2000). LDL-receptor-related proteins in Wnt signal transduction. Nature 407, 530–535. Tamai, K., Zeng, X., Liu, C., Zhang, X., Harada, Y., Chang, Z., and He, X. (2004). A mechanism for Wnt coreceptor activation. Mol. Cell 13, 149–156. Tao, Q., Yokota, C., Puck, H., Kofron, M., Birsoy, B., Yan, D., Asashima, M., Wylie, C. C., Lin, X., and Heasman, J. (2005). Maternal wnt11 activates the canonical wnt signaling pathway required for axis formation in Xenopus embryos. Cell 120, 857–871. Topczewski, J., Sepich, D. S., Myers, D. C., Walker, C., Amores, A., Lele, Z., Hammerschmidt, M., Postlethwait, J., and Solnica-Krezel, L. (2001). The zebrafish glypican knypek controls cell polarity during gastrulation movements of convergent extension. Dev. Cell 1, 251–264. Torban, E., Kor, C., and Gros, P. (2004). Van Gogh-like2 (Strabismus) and its role in planar cell polarity and convergent extension in vertebrates. Trends Genet. 20, 570–577. Ueno, S., Weidinger, G., Osugi, T., Kohn, A. D., Golob, J. L., Pabon, L., Reinecke, H., Moon, R. T., and Murry, C. E. (2007). Biphasic role for Wnt/ beta-catenin signaling in cardiac specification in zebrafish and embryonic stem cells. Proc. Natl. Acad. Sci. USA 104, 9685–9690. Unterseher, F., Hefele, J. A., Giehl, K., De Robertis, E. M., Wedlich, D., and Schambony, A. (2004). Paraxial protocadherin coordinates cell polarity during convergent extension via Rho A and JNK. EMBO J. 23, 3259–3269. Lrp5/6 in Noncanonical Wnt Signaling Vol. 20, February 1, 2009 935 van Amerongen, R., and Berns, A. (2006). Knockout mouse models to study Wnt signal transduction. Trends Genet 22, 678–689. van Noort, M., Meeldijk, J., van der Zee, R., Destree, O., and Clevers, H. (2002). Wnt signaling controls the phosphorylation status of beta-catenin. J. Biol. Chem. 277, 17901–17905. Wang, J., Hamblet, N. S., Mark, S., Dickinson, M. E., Brinkman, B. C., Segil, N., Fraser, S. E., Chen, P., Wallingford, J. B., and Wynshaw-Boris, A. (2006). Dishevelled genes mediate a conserved mammalian PCP pathway to regulate convergent extension during neurulation. Development 133, 1767–1778. Wehrli, M., Dougan, S. T., Caldwell, K., O’Keefe, L., Schwartz, S., VaizelOhayon, D., Schejter, E., Tomlinson, A., and DiNardo, S. (2000). Arrow encodes an LDL-receptor-related protein essential for Wingless signalling. Nature 407, 527–530. Wilkinson, D. G., and Nieto, M. A. (1993). Detection of messenger RNA by in situ hybridization to tissue sections and whole mounts. Methods Enzymol. 225, 361–373. Yamaguchi, T. P., Bradley, A., McMahon, A. P., and Jones, S. (1999a). A Wnt5a pathway underlies outgrowth of multiple structures in the vertebrate embryo. Development 126, 1211–1223. Yamaguchi, T. P., Takada, S., Yoshikawa, Y., Wu, N., and McMahon, A. P. (1999b). T (Brachyury) is a direct target of Wnt3a during paraxial mesoderm specification. Genes Dev. 13, 3185–3190. Zeng, X., Tamai, K., Doble, B., Li, S., Huang, H., Habas, R., Okamura, H., Woodgett, J., and He, X. (2005). A dual-kinase mechanism for Wnt co-receptor phosphorylation and activation. Nature 438, 873–877. V. Bryja et al. Molecular Biology of the Cell936 Supplementary Figure 1 A. ctrl XLRP5MO XLRP6MO mLrp5 mLrp6 pLrp5/6 total protein B. Dorsalized Normal Ventralized Control XLRP6MO XLRP5MO XLRP5MO +-catenin XLRP5MO +mLRP5 C. E-/N-cadherin Supplementary Figure 2 Supplementary Figure 3 E13.5 Wnt5a (relativeexpression) lrp6 +/+ lrp6 +/- lrp6 -/lrp6 +/+ lrp6 +/- lrp6 -/- Wnt5a (relativeexpression) E11.5 0 0,5 1 WT Wnt5a -/- Lrp6 (relativeexpression) E12.5 Supplementary Figure legends Supplementary Figure 1. (A) Efficiency of XLRP5/6 MO and mLrp5/6 overexpression in Xenopus embryos. Cell lysates were injected as indicated and analyzed by Western Blotting with phospho Lrp6 antibody, which also recognizes unphosphorylated Lrp5 and Lrp6, or XLRP5 and XLRP6, respectively. Levels of E-/N-cadherin and total protein were used as a loading control. (B) Scheme of Xenopus injections used in the study. (C) Effects of XLRP5/6 MOs on ventralization. Xenopus embryos were injected as indicated and the ventralization has been scored using DAI index (Kao and Elinson, 1988). Ventralization induced by XLRP5 MO is rescued by mLrp5 and β-catenin. Supplementary Figure 2. Heart morphology in Lrp6 and Wnt5a mutants at E10.5. Heart of E10.5 wild type, Lrp6-/- and Wnt5a-/- mouse embryos was dissected and the morphology was analyzed. No obvious abnormalities were detected at this stage. Representative examples are shown. RV – right ventricle, LV – left ventricle, RA - right atrium, LA - left atrium, OFT – outflow tract. Supplementary Figure 3. Expression of wnt5a in midbrain of lrp6 mutants. Ventral midbrains of mouse embryos with indicated genotypes were dissected and total RNA was isolated. Expression of wnt5a and lrp6 was determined by quantitative RT-PCR . Expression levels of Wnt5a were unchanged by loss of Lrp6, and expression levels of Lrp6 were unchanged by loss of Wnt5a. Supplementary Materials and Methods Mouse genotyping Ear or embryonic tissues were boiled at 95ºC for 40 minutes in 100-200 μl of 25mM NaOH/0,2mM EDTA and an equal volume of 40mM Tris HCl pH 5 was added. 3 μl of this solution was used for PCRs. Lrp6 wild type allele was identified with the previously described primers LRP6-U1 and LRP6-D1, while mice with the gene trap insertion (Lrp6 knockout allele) were recognized with the following primer set: CD4mix forward (5´GCACGGATGTCTCAGATCAAGAGG-3’) and CD4mix reverse (5’CGGGATCATCGCTCCCATATATG-3’), with an annealing temperature of 63ºC and an amplicon of 108bp. Wnt5a WT and null alleles were identified with the following primers: 5a WT For (5´-GACTTCCTGGTGAGGGTGCGTG-3´), Wnt5a WT Rev (5´GGAGAATGGGCACACAGAATCAAC-3´), Wnt5a null For (5´GGGAGCCGGTTGGCGCTACCGGTGG-3´) and Wnt5a null Rev (5´- GGAGAATGGGCACACAGAATCAAC-3´). Sequences of antisense MOs XLRP5 MO (5’-ctc cca tgg cct cgt acc cct ctcc), XLRP6 MO (5’-gct caa tgc tcc ccc gta acc cgac) or standard control MO directed against Lpr5 and Lrp6 were used at 0.8 pmol (all MOs: GeneTools, Philomath, OR, USA). Sequences of XWnt5a and XWnt11 MOs were published before (Pandur et al., 2002; Schambony and Wedlich, 2007). Antibodies Following antibodies were used: anti P-Ser1490-Lrp6 (#2568), anti Lrp6 (#2560) from Cell Signaling Technologies), anti-Dvl2, anti-Dvl3 (sc-8027), anti-c-Myc (sc-40) (from Santa Cruz Biotechnology), anti-active β-catenin (ABC, #05-389) and anti-Rac1 (-389) from Upstate Biotechnology , anti--actin (Ab6276) and anti HA (Ab9110) for IP from Abcam, anti HA (HA.11) for WB from Nordic Biosite, and anti-human IgG Fc fragment (109-005-098, Jackson ImmunoResearch Laboratories). Cadherin levels in Xenopus lysates were detected using the mixture of antibodies against E- and N-cadherin (610182 and 610921, BD Biosciences). Quantitative-PCR cDNA was generated as described previously (Castelo-Branco et al., 2003) RNA from. ventral midbrains of lrp6+/+, lrp6+/- and lrp6-/- mouse embryos at E11.5 and E13.5.5 (where both Lrp6 and Wnt5a are known to be expressed) was extracted using RNeasy Mini Kit (Qiagen). 1 g of RNA was reverse-transcribed using SuperScript II Reverse Transcriptase (Invitrogen) and random primers (Invitrogen). Wnt5a primers used in this study have been described previously (Castelo-Branco et al., 2003), Lrp6 primers were as follows: forward 5-GCTACAAATGGCAAAGAGAATGC-3, reverse CAGTATACAAGCCATGACCAAACA.       Vítězslav Bryja, 2014    Attachments      #8      Andersson T, Södersten E, Duckworth JK, Cascante A, Fritz N, Sacchetti P, Cervenka I,  Bryja  V,  Hermanson  O  (2009):  CXXC5  is  a  novel  BMP4‐regulated  modulator  of  Wnt‐ signaling in neural stem cells. J Biol Chem. 284(6):3672‐81.      Impact factor (2009): 5.328  Times cited (without autocitations, WoS, Feb 21st 2014): 24  Significance:  Identification  of  a  previously  unknown  protein,  CXXC5,  which  acts  as  a  negative regulator of canonical Wnt signaling via interaction with Dishevelled.  Contibution  of  the  author/author´s  team:  Analysis  of  the  role  of  CXXC5  in  the  Wnt  signaling pathway.                CXXC5 Is a Novel BMP4-regulated Modulator of Wnt Signaling in Neural Stem Cells*□S Received for publication,October 22, 2008 Published, JBC Papers in Press,November 10, 2008, DOI 10.1074/jbc.M808119200 Therese Andersson‡ , Erik So¨dersten‡ , Joshua K. Duckworth‡ , Anna Cascante‡1 , Nicolas Fritz§2 , Paola Sacchetti§ , Igor Cervenka¶ , Vitezslav Bryja¶3 , and Ola Hermanson‡4 From the ‡ CoE in Developmental Biology for Regenerative Medicine (CEDB/DBRM), Department of Neuroscience, and § CoE in Developmental Biology for Regenerative Medicine (CEDB/DBRM), Department of Medical Biochemistry and Biophysics (MBB), Karolinska Institutet, SE17177 Stockholm, Sweden and the ¶ Institute of Experimental Biology, Faculty of Science, Masaryk University and the Department of Cytokinetics, Institute of Biophysics, Academy of Sciences of the Czech Republic, 61137 Brno, Czech Republic Bone morphogenetic proteins such as BMP4 are essential for proper development of telencephalic forebrain structures and induce differentiation of telencephalic neural stem cells into a variety of cellular fates, including astrocytic, neuronal, and mesenchymal cells. Little is yet understood regarding the mechanisms that underlie the spatiotemporal differences in progenitor response to BMP4. In a screen designed to identify novel targets of BMP4 signaling in telencephalic neural stem cells, we found the mRNA levels of the previously uncharacterized factor CXXC5 reproducibly up-regulated upon BMP4 stimulation. In vivo, CXXC5 expression overlapped with BMP4 adjacent to Wnt3a expression in the dorsal regions of the telencephalon, including the developing choroid plexus. CXXC5 showed partial homology with Idax, a related protein previously shown to interact with the Wnt-signaling intermediate Dishevelled (Dvl). Indeed CXXC5 and Dvl co-localized in the cytoplasm and interacted in co-immunoprecipitation experiments. Moreover, fluorescence resonance energy transfer (FRET) experiments verified that CXXC5 and Dvl2 were located in close spatial proximity in neural stem cells. Studies of the functional role of CXXC5 revealed that overexpression of CXXC5 or exposure to BMP4 repressed the levels of the canonical Wnt signaling target Axin2, and CXXC5 attenuated Wnt3a-mediated increase in TOPflash reporter activity. Accordingly, RNA interference of CXXC5 attenuated the BMP4-mediated decrease in Axin2 levels and facilitated the response to Wnt3a in neural stem cells. We propose that CXXC5 is acting as a BMP4–induced inhibitor of Wnt signaling in neural stem cells. Members of the TGFß family such as bone morphogenetic proteins (BMP)5 influence multiple essential events during brain development, such as differentiation, proliferation, and migration (1–3). Stimulation of telencephalic neural stem cells by BMP4 induces differentiation into a variety of cellular fates, including neuronal, astrocytic and smooth muscle cells in vitro, and genetic studies have shown that BMP4 is essential for proper differentiation and regionalization of the telencephalic forebrain (1, 2, 4, 5). BMP4 mediates its effects through nuclear translocalization of Smad proteins such as Smad1 and Smad4 that can act directly as transcription factors and associate with a number of important cofactors, including TGIF, Sip1, and CBP/p300 (6, 7). BMP activity exert cross-talk with many signaling pathways, such as the membrane-bound receptor Notch, fibroblast growth factors (FGFs), and Wnt factors (8, 9), and it has been proposed that BMP molecules act in synergy with canonical Wnt signaling molecules, such as Wnt3a, to regulate telencephalic regionalization (1, 10). Less is known regarding downstream targets of BMP signaling that regulate the spatial and temporal context-specific differences in progenitor responsiveness to extracellular signaling factors. BMP activity is directly regulated by extracellular inhibitors such as noggin and chordin, and Wnt signaling activity is regulated at many levels by both extra- and intracellular mechanisms. An example of an intracellular inhibitor of Wnt signaling with reported effects on forebrain development is Idax. Idax binds directly to the Wnt signaling mediator Dishevelled (Dvl) and seems to compete at the binding site of Axin, resulting in an inhibition of Wnt activity (11). Idax belongs to the CXXC family of proteins of which other members, such as MBD1, are predominantly nuclear DNA-binding proteins with chromatin-modifying properties (12). Dvl is predominantly albeit not exclusively localized in the cytoplasm, and Idax appears to be an exception to other CXXC domain-containing proteins, being localized largely in the cytoplasm (11). It has been shown that some CXXC proteins are cancer-associated genes that can regulate the activity and function of important transcriptional complexes, such as Polycomb proteins (12). Idax is expressed in * This study was supported in part by grants from K&A Wallenberg Foundation, the Swedish Research Council (VR), the Swedish Cancer Society (CF), the Jeansson Foundation, the Åke Wiberg Foundation, the Åhle´n Foundation, the Swedish Medical Society, Karolinska Institutet, the Swedish Foundation for Strategic Research (SSF), and the Swedish Children’s Cancer Foundation (BCF) (to O. H.). The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. □S The on-line version of this article (available at http://www.jbc.org) contains supplemental Fig. S1 and Table S1. 1 Supported by Fundacio´n Ramo´n Areces, Spain. 2 Supported by VR-M, Sweden. 3 Supported by MSM0021622430 (Ministry of Education, Youth and Sports of the Czech Republic), KJB501630801, AVOZ50040507, AVOZ50040702 (Academy of Sciences of the Czech Republic), and EMBO Installation Grant. 4 To whom correspondence should be addressed. Tel.: 46-8-5248-7477; 46-76-118-7452; Fax: 46-8-341-960; E-mail: Ola.Hermanson@ki.se. 5 The abbreviations used are: BMP, bone morphogenetic protein; GFP, green fluorescent protein; FRET, fluorescence resonance energy transfer; DAPI, 4Ј,6-diamidino-2-phenylindole; ROI, region of interest; RT-qPCR, reverse transcriptase-quantitative PCR; NSC, neural stem cells. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 284, NO. 6, pp. 3672–3681, February 6, 2009 © 2009 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A. 3672 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284•NUMBER 6•FEBRUARY 6, 2009 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom http://www.jbc.org/cgi/content/full/M808119200/DC1 Supplemental Material can be found at: anterior brain structures and is required for appropriate forebrain development as loss of Idax by morpholinos in Xenopus tadpoles result in severe forebrain defects (13). In a systematic attempt to increase the understanding of the mechanisms underlying cell context-specific responses of telencephalic neural stem cells (NSCs) to BMP4 signaling during forebrain development, we have pursued gene expression profiling analysis using microarrays to identify putative direct and novel targets for BMP4 signaling in NSCs. By this approach, we identified CXXC5 as a direct BMP4 target in NSCs. In situ hybridization experiments revealed that CXXC5 expression localized selectively to the BMP4-rich dorsal telencephalon adjacent to the neighboring regions expressing Wnt3a. Although CXXC5 displayed predominantly nuclear localization, it co-localized and interacted with several Dvl isoforms in the cytoplasm. CXXC5 influenced the expression of the endogenous Wnt signaling target Axin2 as well as Wnt3a-induced activation of a TOPflash reporter. Based on these results, we propose that CXXC5 is a novel BMP-regulated modulator of Wnt signaling and targets thereof in telencephalic NSCs. EXPERIMENTAL PROCEDURES Cortical NSC Cultures—NSCs were obtained from the dissociated cerebral cortices of Sprague-Dawley rat E15.5 embryos and cultured as previously described (14–16). Briefly NSCs were cultured in serum-free DMEM:F12 media (Invitrogen) enriched with N2 supplement and grown on poly-L-ornithine/ fibronectin (Sigma)-coated cell culture dishes (Corning). Cells were maintained in a proliferative state using 10 ng/ml FGF2 (R&D Systems) and passaged once before nucleofections or twice before stimulations. After the second passage, cells were plated at 500–10,000 cells/cm2 and allowed to proliferate for 24 h prior to commencement of the experiment. To induce differentiation of NSCs, FGF2 was withdrawn from the cultures with or without the addition of other soluble factors. BMP4 and Wnt3a (R&D Systems) were both used at 10 ng/ml. Addition of soluble factors was carried out every 24 h, and media was changed every 48 h. For inhibition of protein synthesis, 10–20 ␮g/ml Cycloheximide (Sigma) was added to the media 15–30 min before addition of BMP4. All experiments in this study involving animals and stem cell isolation were approved by the ethics committee for animal research in Stockholm, Sweden. Gene Expression Profiling Using Microarrays—Rat neural stem cells were exposed to BMP4 or kept in FGF2 for 3h before the cells were lysed and RNA extracted (Qiagen). RNA from the time for stimulation start was extracted and used as control. The microarray was carried out by KI seq-express (former KIChip). Briefly, RNA was amplified (17) and labeled with Cy3 or Cy5 (PerkinElmer Life Sciences) and hybridized onto Agilent whole rat genome array (G4131A) (all kits and arrays from Agilent). The arrays were scanned at two different intensities, and the images were analyzed for background correction. Both BMP4 and FGF2 samples where co-hybridized with RNA from the starting point and a dye-swap was performed. The arrays were normalized and the differential gene expression analyzed using the R and bioconductor-based method LIMMA (18). RNA Isolation and Quantitative RT-PCR—RNA was isolated using RNeasy and contaminating DNA removed using RNasefree DNase set (Qiagen). First strand cDNA synthesis was obtained using High Capacity cDNA Reverse Transcription kit (Applied Biosystems). For quantitative PCR 0.5ϫ Platinum SYBR green mix (Invitrogen) was used and run on Applied Biosystems 7300 Real Time PCR system using the PCR-program recommended by Invitrogen for Platinum SYBR green. Primers used for the quantitative RT-PCR were designed to span an intron where possible. Sequences can be obtained upon request. To further exclude DNA contamination, ϪRT (no reverse transcriptase) samples were run as controls. Data were analyzed using the statistical programming language R. DNA Constructs—The open reading frame of rat CXXC5 was amplified using Taq-polymerase (Invitrogen) and gene-specific primers and cloned into pCR4 using TOPO TA kit according to manufacturers instructions (Invitrogen). Isolated clones were sequenced (KI seq-express) before subcloning into pcDNA3 and subsequent use in overexpression experiments. To obtain CXXC5-GFP and GFP-CXXC5 fusion proteins, the PCR product was purified using a PCR purification kit (Qiagen), cut with restriction enzymes (NEB), and ligated using Quick ligase (NEB) into the HindIII and EcoRI sites of pNGFP-EU or pCGFP-EU (19). Clones were sequenced and subsequently used for co-immunoprecipitation, FRET, and subcellular localization experiments. In vitro transcription/translation was performed using TNT (Promega) according to the supplier’s instructions. The BMP4 and Wnt3a constructs were kind gifts from Dr. John L. R. Rubenstein. In Situ Hybridization—Non-radioactive in situ hybridization on sections was performed as previously described (20). Heads (E10.5, E12.5, and E14.5) of mouse embryos were fixed in 4% paraformaldehyde (Sigma) for 2–5 h and then transferred to 30% sucrose at 4 °C overnight before frozen in Tissue-Tek (Sakura Finetek) on dry ice and kept in Ϫ70 °C until cryostatsectioned (Microm) onto SuperFrost Plus glasses (MenzelGmbH & Co KG) as 12-␮m sections. Prior to hybridization, sections were permeabilized with 1 ␮g/ml proteinase K (Invitrogen), acetylated with 13.5 ml/L triethanolamine (Sigma), 44 mM acetic anhydride (Sigma) and 1.7 mM HCl (Merck), and pre-hybridized with hybridization buffer (50% formamide (Fluka), 5ϫ SSC, 5ϫ Denhardt’s solution (Sigma), 250 ␮g/ml bakers’ yeast RNA (Sigma), 500 ␮g/ml herring sperm DNA (Invitrogen), 1 g/50 ml blocking reagent (Boehringer)) for 3–6 h. Hybridization was carried out at 70 °C overnight with 1–4 ␮l of DIG-labeled probe (0.1–0.2 ␮g/␮l) per 100 ␮l of hybridization buffer and glass. Post-hybridization washes were performed for 1–2 h at 70 °C in 0.2ϫ SSC. Sections were then blocked with 10% fetal calf serum for 2–4 h before being subjected to an anti-DIG antibody (1:5000, Roche) at 4 °C overnight. Finally, the slides were submerged into a 10% polyvinyl alcohol (Sigma) solution containing 1.72 ml/l NBT (Roche) and 1.72 ml/l B-CIP (Roche) chromogen components until sufficiently stained (3–72 h). DIG probes were produced according to the manufacturers’ recommendations using DIG RNA labeling mix (Roche). After in situ hybridization, the sections were analyzed and micrographs obtained with a Zeiss Axioskop 2 mot plus microscope. CXXC5, BMP4, and Wnt3a probes were used on Ͼ3 littermate brains per developmental stage. CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling FEBRUARY 6, 2009•VOLUME 284•NUMBER 6 JOURNAL OF BIOLOGICAL CHEMISTRY 3673 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom Immunoprecipitation, Immunoblotting, and Immunocytochemistry—COS7 and HEK-293 cells were transfected with GFP-CXXC5, CXXC5-GFP, or vector control only containing GFP individually or cotransfected with a full-length XDishevelled-Myc (21) using Lipofectamine 2000 (Invitrogen) according to the manufacturer’s instructions. Immunoprecipitation, immunoblotting, and culturing of COS7 and HEK-293 cells were performed essentially as previously described (22). Antibodies used for immunoprecipitation were anti-GFP (Fitzgerald Industries), rabbit anti-Myc (SCBT) and anti-Dvl3 (Cell Signaling Technology). Primary antibodies used for immunoblotting: anti-ß-catenin (BD Biosciences), anti-Dvl1 (SCBT), anti-Dvl2 (SCBT), anti-Dvl3 (SCBT), anti-GFP (Chemicon), mouse anti-Myc (SCBT), and anti-ß-actin (Abcam) (22). For immunocytochemistry, cultures were fixed using 10% formalin (Sigma) 24 h after transfections. Immunostaining was performed using standard protocols using chicken polyclonal anti-GFP (Chemicon, 1:1000) and mouse monoclonal anti-c-Myc (SCBT, 1:200) followed by appropriate species-specific Alexa-488 and Alexa-594-conjugated secondary antibodies (Molecular Probes, 1:500). Nuclei were visualized using Vectashield containing DAPI (Vector Laboratories, Inc.). Fluorescent and brightfield images were acquired using Zeiss Axioskop2/ MRm camera with Axiovision software. Images were assembled using Adobe Photoshop. FRET—To perform Fo¨rster (Fluorescent) Resonance Energy Transfer (FRET) measurements, we used a Zeiss LSM 5 Exciter inverted confocal scanning laser microscope using a ϫ63/1.2 NA objective as described previously (23). CXXC5-GFP-expressing NSCs were grown on prepared glass coverslips, and 24 h after nucleofection, cells were fixed using a 50% methanol/ 50% acetone fixing solution and then processed for immunocytochemistry. The coverslips were mounted on slides using the ProLong Antifade mounting media kit (Molecular Probes, Invitrogen) to enable perfect conservation of the fluorescent signal. A detailed description of the FRET technique can be found elsewhere (24, 25). The Fo¨rster constant, R0, for the donor-acceptor pair, GFP and Alexa-555, used in this study was 6.3 nm (Handbook of Probes, Invitrogen). FRET occurs when the fluorophores are separated by distances 0.5 R0 Ͻ r Ͻ 2 R0. Thus, it is possible to distinguish proteins that are spatially colocalized within a ϳ14-nm radius. To determine FRET, we quantified the quenching of donor fluorescence by performing acceptor photobleaching. NSCs expressing CXXC5-GFP and stained with Alexa555-labeled secondary goat anti-rabbit IgG antibody (Molecular Probes, Invitrogen; 1:100) to detect rabbit polyclonal antibody to Dvl2 (SCBT; 1:100) were excited with the 488 and 561 nm lasers. The acceptor, Alexa555, was then irreversibly photobleached in a selected adequate cytoplasmic region (ϳ5–10 ␮m2 ) by continuous excitation with the 561 nm laser for 10–20 s. Thereafter, the residual Alexa555 and GFP image was obtained, and a region of interest (ROI) was outlined in the photobleached area and processed using Zeiss LSM image Examiner. Ratios between GFP intensities in the ROI, before and after photobleaching, were calculated to quantify FRET. The FRET values presented are corrected for erroneous intensity changes for each cell in two cytoplasmic ROIs randomly selected outside the bleached area. 10–25 cells were measured for each experiment. Overexpression and siRNA-mediated Knockdown—NSCs were nucleofected according to the manufacturer’s instructions using program A-33 on the Nucleofector (Amaxa) as previously described (26). Experiments were commenced 24h after nucleofection, during which NSCs were maintained in FGF2. As the healthiness of the cells varied significantly when obtaining very high (Ͼ90%) or very low (Ͻ20%) knockdown of CXXC5 expression levels, experiments where a knockdown of CXXC5 mRNA showed between 40–70% efficiency (n ϭ 5) were chosen for detailed analysis. Transient transfection studies for luciferase assay in fibroblasts were performed essentially as previously described (27). Cells were plated in 24-well plates 24h before transfection and transfected with 0.6 ␮g of plasmid DNA/well complexed with 2 ␮l of Lipofectamine 2000 (Invitrogen). Typically, cells were transfected with 200 ng of the reporter construct and increasing doses of CXXC5 expression vector. After 5 h of incubation, the lipid/DNA mix was replaced with fresh 10% serum medium. Cells were serum-starved 24-h post-transfection and stimulated with vehicle or purified Wnt3a (R&D Systems). Luciferase activities were assayed using Dual-Luciferase Reporter Assay System (Promega), following the manufacturer’s protocol. Primary antibodies used for immunoblotting: anti-ß-catenin (BD Biosciences), anti-activated (dephosphorylated) ß-catenin (Upstate), anti-Dvl2 (SCBT), and anti-ß-actin (Abcam). Statistical Analysis—Statistical analysis was performed in Prism4 (GraphPad software). Initially variance analysis (ANOVA) was performed, followed by Student’s t test (unpaired). Significance were assumed at the level of p Ͻ 0.05 (*, p Ͻ 0.05; ***, p Ͻ 0.001, see figure legends). RESULTS BMP4 Rapidly Induces Significant Changes in Gene Expression in Neural Stem Cells—To elucidate novel downstream targets of BMP signaling in neural progenitors, we used a well-characterized rat embryonic telencephalic NSC preparation that has previously been used to study BMP signaling. During the continuous administration of FGF2 the NSCs remain undifferentiated with the capacity to differentiate into neuronal, astrocytic, oligodendrocyte, and mesenchymal cells (14–16, 26). Upon BMP4 stimulation these cells respond with increased Smad phosphorylation, subsequent nuclear translocation of Smads and increase in transcription of Smad target genes (5), and subsequent differentiation into astrocytic and mesenchymal cells (5, 28). After primary FGF2 expansion, one passage and continued FGF2 expansion, NSCs were treated with either FGF2 or BMP4 for 3 h. This time point was chosen due to the rapid induction (Ͻ30 min) of Smad-mediated transcription downstream of BMP4 stimulation (5). The gene expression profile was subsequently investigated using microarrays. This procedure identified 162 spots that were putatively (p value Ͻ 0.05) up- or downregulated more than 1.5-fold after 3 h (supplemental data). To initially validate the array, we picked Ϸ10% of the identified genes and performed quantitative RT-PCR (RT-qPCR) on the same samples that were used for the microarrays. This experiment showed that around 90% of the genes identified in the CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling 3674 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284•NUMBER 6•FEBRUARY 6, 2009 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom screen were up- or down-regulated, in approximate accordance with the chosen p value predicting 95% accuracy (see further below). The effects of BMP4 stimulation of NSCs are highly dependent on the microenvironment and cell culture conditions, such as plating density (4, 28). To analyze the reproducibility of the effects of BMP4 on the expression of the identified genes in NSCs, we therefore picked eight factors from our screen (Fig. 1A and see below), including known and potentially novel targets of BMP4, and repeated the experimental procedure four times. Six out of these eight mRNAs were reproducibly BMP4-regulated in NSCs: Gas1, Olig1, Id2, Idb4, CXXC5, and Hey1 (Fig. 1A), whereas two mRNAs, Sin3A and Skiip, were regulated only in occasional experiments (data not shown). CXXC5 Is a Novel Direct Target of BMP4 in NSCs—Id2 and Idb4 have previously been shown to be direct targets of BMP signaling. These factors interact with and down-regulate the activity of bHLH proteins, such as the oligodendrocyte-associated bHLH transcription factors Olig1 and Olig2 thus interfering with oligodendrocyte differentiation (29). In addition to this mechanism, we found that Olig1 expression was significantly down-regulated by BMP4 stimulation (Fig. 1A), suggesting that BMP4-mediated inhibition of oligodendrocyte differentiation may occur at several levels. Other reproducibly BMP4regulated mRNAs were the Notchassociated repressor Hey1 (8), and Gas1 recently linked to Sonic Hedgehog signaling (30–32) (Fig. 1A). Another significantly BMP4-regulated mRNA was the previously uncharacterized factor CXXC5 (Fig. 1A). CXXC5 expression levels were up-regulated by BMP4 approximately 15-fold in average after 3 h as assessed by RT-qPCR (Fig. 1, A and B). Time course experiments (0, 20, 40, 60, 90, 120, 150, 180 min, 4, 6, 10 h after initial BMP4 exposure) revealed that CXXC5 expression levels started to increase at 40 min, were high at 60 min and remained up-regulated at 10 h after BMP4 exposure (see further below and data not shown). FIGURE 1. BMP4 exposure results in rapid changes in gene expression in telencephalic NSCs. A, levels of mRNA expression relative to control mRNA (HPRT) of six putative target genes (Gas1, Olig1, Id2, Idb4, CXXC5, Hey1) in NSCs after BMP4 treatment and relative to unstimulated control as assessed by RT-qPCR. B, RT-qPCR assessment of the relative CXXC5 mRNA expression levels to HPRT after BMP4 stimulation with or without pretreatment with the protein synthesis inhibitor cycloheximide (CHX) compared with unstimulated control. FIGURE 2. CXXC5 mRNA is expressed at high levels in the dorsal telencephalon, the expression overlaps with BMP4 and is adjacent to Wnt3a expression. A–I, micrographs depicting in situ hybridization results of transverse sections of mouse brains at embryonic ages (E) 10.5 (A–C), 12.5 (D–F), and 14.5 (G–I) hybridized with probes for detection of CXXC5 (A, D, G), BMP4 (B, E, H), and Wnt3a (C, F, I). At E10.5, the expression patterns overlap significantly. At E12.5, CXXC5 and BMP4 mRNA expressions overlap significantly and are detected in the very close vicinity of Wnt3a labeling. At E14.5, very low levels of CXXC5 transcripts were detected, whereas BMP4 mRNA was found in and around the developing choroid plexus, and Wnt3a labeling was detected just lateral to the BMP4-expressing domain. Arrows mark regions of high specific labeling. CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling FEBRUARY 6, 2009•VOLUME 284•NUMBER 6 JOURNAL OF BIOLOGICAL CHEMISTRY 3675 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom To investigate whether CXXC5 could be a direct target of BMP4 signaling, we pretreated the NSCs with cycloheximide (CHX), an inhibitor of translation and thus de novo protein synthesis. We then evaluated the expression levels of CXXC5 3 h after subsequent FGF2 versus BMP4 stimulation. Pretreatment with CHX 15–30 min before BMP4 or FGF2 stimulation did not significantly inhibit BMP4-mediated up-regulation of CXXC5 (Fig. 1B), suggesting that the increased levels of CXXC5 mRNA could be the result of a direct effect of BMP4 signaling on CXXC5 gene expression. CXXC5 Is Highly Expressed in the Dorsal Telencephalon Where BMP4 Expression Is Enriched—To confirm the gene product of CXXC5, we performed in vitro transcription/translation by TNT using a sequenced construct harboring fulllength CXXC5. The protein product revealed a major band migrating at around 38 kDa (data not shown), in close range of the predicted protein product of the CXXC5 gene. To assess the gene expression pattern of CXXC5 in vivo in relation to BMP4 in neural progenitors, we performed in situ hybridization on forebrain sections from mice at embryonic age (E) 10.5, 12.5, and 14.5. At E10.5 and E12.5, CXXC5 expression was largely restricted to the dorsal pallium (Fig. 2, A and D). These regions are known to be rich in expression of BMP receptors as well as several BMP molecules, including BMP4 (2, 3). Indeed in situ hybridization experiments using a probe for BMP4 demonstrated a clear overlap in expression of CXXC5 and BMP4 in the dorsal pallium (Fig. 2, A, B, D, and E). Importantly, CXXC5 expression was prominent in and around the developing choroid plexus (Fig. 2D), a non-neuronal structure producing and secreting the cerebrospinal fluid. At E14.5 however, only low levels of CXXC5 expression were detected (Fig. 2G), whereas BMP4 transcripts displayed a high expression selectively in the ventricular zone at the region of the developing choroid plexus (Fig. 2H). In addition to BMP4, the signaling molecule Wnt3a has been shown to be required for proper development of the dorsal telencephalon, influencing the development of the hippocampal formation and pyramidal neurons therein (10). At E10.5, Wnt3a mRNA showed an overlapping expression with CXXC5 and BMP4 (Fig. 2, A–C), while at E12.5 Wnt3a transcripts showed a marked reduction in expression just at the border of CXXC5 and BMP4 expression (Fig. 2, D–F), possibly demarcating the regions of the developing hippocampal formation and the choroid plexus. At E14.5 a marked expression of the Wnt3a gene was detected just lateral to the BMP4 expression (Fig. 2, H and I). The expression pattern of CXXC5 thus suggests that this factor could play a role in telencephalic development and strengthens the association between CXXC5 and high BMP signaling activity. CXXC5 Interacts with the Wnt Signaling Mediator Dishevelled in Co-immunoprecipitation Experiments—CXXC5 has been predicted to be a paralogue of another CXXC domain containing factor named Idax (CXXC4) (11, 12, 33). Idax was identified in a screen for proteins interacting with the PDZ domain of the intracellular signaling factor Dishevelled (Dvl). Dvl is a predominantly cytoplasmic protein (34, 35) and acts as an intermediate of canonical Wnt signaling. Accordingly, Idax was found to inhibit canonical Wnt signaling as assessed by transcriptional assays (11). A model of the sequence in the domain of Idax that interacts with Dvl has been generated and this sequence shows 100% similarity to a sequence in CXXC5 (supplemental data). We, therefore, generated constructs where GFP was fused to the N or C terminus of CXXC5 to investigate whether Dvl interacts with CXXC5 in a similar manner as Idax. After co-transfection of the GFP fusion proteins with full-length Myc-Dvl in COS7 and HEK-293 cells, we immunoprecipitated whole cell protein extracts using Myc- or FIGURE3.CXXC5interactswithtransfectedandendogenousDvlproteins in co-immunoprecipitation experiments. A, both N- and C-terminally tagged full-length CXXC5 constructs interact with Myc-Dvl in co-immunoprecipitation (IP) experiments using GFP antibody for IP and Myc antibodies for blot. Control vector expressing only GFP does not show any interactions. Arrows point to CXXC5 (upper band) and GFP (lower band). An additional unspecific protein product is seen after IP with GFP-CXXC5 only. B, CXXC5GFP co-immunoprecipitates with Myc-Dvl also when using Myc antibody for IP whereas the GFP-expressing vector shows no interaction. C, COS7 cells were transfected with empty vector or CXXC5-GFP fusion protein. Cell lysates were immunoprecipitated with anti-GFP or anti-Dvl3 antibodies and analyzed for the presence of Dvl3 or CXXC5-GFP in the complex by Western blotting. IP, immunoprecipitation; TCL, total cell lysate, WB, Western blot. CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling 3676 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284•NUMBER 6•FEBRUARY 6, 2009 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom GFP-specific antibodies. This experiment revealed that both constructs of CXXC5 could be co-immunoprecipitated with Myc-Dvl in both cell types and with both antibodies, whereas the control vector expressing GFP alone did not interact with Dvl (Fig. 3, A and B). To investigate whether endogenous Dvl protein could interact with the CXXC5-GFP protein, we immunoprecipitated whole cell protein extracts of COS7 cells transfected with CXXC5-GFP using either the GFP antibody or antibodies against Dvl proteins, and examined putative co-immunoprecipitation by subsequent immunoblotting for the Dvl proteins or GFP, respectively. In this experiment, Dvl3, the most abundant Dvl isoform in COS7 cells, co-immunoprecipitated with CXXC5-GFP during both experimental conditions, whereas the vector only expressing GFP did not (Fig. 3C). These results suggest that CXXC5 interact with endogenous Dvl proteins, including Dvl3. CXXC5 Displays Both Nuclear and Cytoplasmic Subcellular Localization and Co-localizes with Dvl in the Cytoplasm—Although Idax has been demonstrated to be largely a cytoplasmic protein, in silico screens predicted CXXC5 to localize to the nucleus, despite the lack of classical nuclear localization signal motifs. It has further been reported that Dvl can display nuclear localization (34). To investigate whether the interaction of CXXC5 and Dvl occurred in the nucleus or in the cytoplasm, we investigated the subcellular localization of CXXC5 and Dvl. Transfection of the GFP-CXXC5 construct into COS7 cells revealed that CXXC5 located both to the nucleus and the cytoplasm (Fig. 4A). Nuclear GFP-CXXC5 could be detected in 90–95% of the transfected cells whereas clear cytoplasmic labeling was detected in ϳ30–40% of the transfected cells. The nuclear labeling was homogenous and associated with DNA as assessed by merging with the DAPI staining (Fig. 4D). The cytoplasmic labeling of CXXC5 displayed characteristics of large bodies in close proximity to the nucleus, localized just outside the nuclear membrane. Transfections of Myc-Dvl revealed a predominantly cytoplasmic localization (Fig. 4B). Co-transfection revealed that CXXC5 and Dvl indeed co-localized extensively in the cytoplasm (Fig. 4, C and D). The results showed some experimental variations but yielded an average of co-localization in around 40% of the co-transfected cells. FRET Experiments Show that CXXC5-GFP and Dvl2 Are Located in Close Proximity in Neural Stem Cells—All proteins of the Dishevelled family are expressed during the development of the nervous system, and Dvl2 is highly expressed in forebrain progenitors. In order to investigate the interaction between endogenous Dvl2 and CXXC5 specifically in NSCs, we performed Fo¨rster FRET experiments (23–25). For FRET quantification we used the acceptor photobleaching method on NSCs nucleofected with CXXC5-GFP. Endogenous Dvl2 was detected in NSCs by immunocytochemistry using polyclonal anti-Dvl2 IgG. CXXC5-GFP served as FRET donor. The primary antibody against Dvl2 was probed with an Alexa555-conjugated IgG secondary antibody, which served as the FRET acceptor (Dvl2Alexa555). In each cell, we measured the intensity of the CXXC5-GFP and Dvl2-Alexa555 in a subcellular ROI showing overlap of the two proteins, before and after irreversible and total acceptor photobleaching (Fig. 5A). Subsequently, two subcellular regions were randomly selected away from the photobleached region and the intensity of CXXC5-GFP was measured before and after photobleaching. These values were used to correct the value measured in the bleached region for erroneous intensity changes. Following acceptor photobleaching, the CXXC5-GFP fluorescence intensity in the ROI was significantly enhanced to 108.0 Ϯ 0.7% of the intensity before photobleaching (p Ͻ 0.001, n ϭ 23, Fig. 5B). These results implied that the donor and acceptor complexes, CXXC5-GFP and Dvl2-Alexa555, were in close proximity, separated by less than 14 nm, i.e. the maximal distance for FRET detection between GFP and Alexa555 (24). FIGURE 4. CXXC5 is localized both to the nucleus and the cytoplasm, and in the cytoplasm CXXC5 colocalizes with Dvl. A, micrograph depicting transfected GFP-CXXC5 in COS7 cells as detected by anti-GFP antibody (green). Note that CXXC5 localizes both to nuclear and cytoplasmic compartments. B, micrograph showing transfected Myc-Dvl in COS7 cells as revealed by immunocytochemistry (red). C, merged micrographs demonstrating the extent of subcellular co-localization of GFP-CXXC5 and Myc-Dvl in the cytoplasm (yellow). D, merged micrographs depicting GFP-CXXC5, Myc-Dvl, and nuclear staining using DAPI (blue). CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling FEBRUARY 6, 2009•VOLUME 284•NUMBER 6 JOURNAL OF BIOLOGICAL CHEMISTRY 3677 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom To ensure that the resulting increase in CXXC5-GFP fluorescence intensity was a specific consequence of the photobleaching of Dvl2-Alexa555 and not a result of laser excitation, we decreased the laser power during photobleaching to levels unable to fully bleach Dvl2-Alexa555. Dvl2-Alexa555 fluorescence intensity following the photobleaching protocol was reduced to 20–30% (20% laser power), 40–50% (10% laser power), or unchanged (0.1% laser power). In none of these cases was the CXXC5-GFP fluorescence intensity enhanced after photobleaching (20% laser power, 101.2 Ϯ 1.3, n ϭ 7; 10% laser power, 97.8 Ϯ 3.3, n ϭ 7; 0.1% laser power, 98.8 Ϯ 0.6, n ϭ 6). To further assure that the observed FRET between CXXC5 and Dvl2 uniquely reflected a property of this pair of proteins, we performed control experiments in NSCs using nucleofection and expression of either GFP alone or of another GFPtagged cytoplasmic protein, NFATc1. FRET analysis was performed using GFP/GFP-NFATc1 (donor) and the same Alexa555-labeled secondary antibody to detect the Dvl2 antibody (Dvl2-Alexa555, acceptor). No significant change in donor emission ratio before and after acceptor photobleaching was found for neither of these molecular pairs (n ϭ 11, data not shown). Taken together, these results indicate that CXXC5 and Dvl2 are located in close proximity (Ͻ14 nm) strengthening the suggestion that there is a specific physical association between CXXC5 and Dvl2 in NSCs. BMP4 Stimulation or Overexpression of CXXC5 Attenuates the Expression Levels of the Wnt3a Target Gene Axin2 in NSCs—To elucidate a functional role for the BMP4-induced increase in CXXC5 expression levels and the interaction between CXXC5 and Dvl, we next aimed at investigating the effects of CXXC5 overexpression on the response to Wnt signaling in the NSCs. In the developing nervous system, canonical Wnt signaling by Wnt3a has been show to exert cross-talk with BMP signaling in the neural crest and in the dorsal regions of the developing telencephalon (1, 9, 10, 36) where high levels of CXXC5 expression were found (Fig. 2, A and D). To first investigate the responsiveness of telencephalic NSCs to dorsally expressed signaling factors, we used RT-qPCR to estimate the endogenous levels of the bona fide Wnt target Axin2 in response to short-term exposure to recombinant BMP4 or Wnt3a. Axin2 gene expression is directly regulated by the Wnt signaling mediator ß-catenin and is commonly used as a measurement for canonical Wnt signaling activity (37–39). BMP4 treatment led to significantly reduced levels of Axin2 expression within 6 h, in accordance with an up-regulation of CXXC5 and thus inhibitory effect on Wnt signaling activity (Fig. 6A). In contrast, Wnt3a stimulation resulted in a significant increase (Ϸ4-fold) in Axin2 mRNA expression 4 h after stimulation compared with control (Fig. 6B, black bars). To study whether the Wnt3a response would be affected by increased levels of CXXC5 as assessed by Axin2 expression levels, we next overexpressed CXXC5 in the NSCs by nucleofection. Nucleofection efficiency was estimated to vary between 30–60% in individual experiments with the NSCs. Overexpression of CXXC5 resulted in a marked attenuation of the Wnt3amediated increase in Axin2 expression levels (Fig. 6B). A decrease in Axin2 expression levels was seen also in FGF2treated control cultures by CXXC5 overexpression (Fig. 6B). These results suggest that high levels of CXXC5 interfere with the response to Wnt3a stimulation in NSCs. CXXC5 Overexpression Attenuates Wnt3a-mediated Increase in TOPflash Reporter Activity—To see whether the effect of CXXC5 on Wnt signaling was selective for Axin2 expression in NSCs or a general effect on canonical Wnt signaling, we next investigated the effects of CXXC5 expression on Wnt3a mediated activation of a TOPflash reporter in HEK-293 cells, a comFIGURE 5. FRET experiments reveal that CXXC5-GFP and endogenous Dvl2 are localized in close proximity in neural stem cells. Acceptor pho- tobleachingFRETmeasurementswereperformedonNSCsnucleofectedwith CXXC5-GFP (donor) and in which Dvl2 was revealed by an Alexa555-coupled IgG secondary antibody (acceptor). A, expression of Dvl2-Alexa555 and CXXC5-GFP in NSCs before (left panel) and after (right panel) acceptor photobleaching. The bleached region is indicated by a white rectangle on the pictures. Scale bar: 5 ␮m. B, quantitative analysis of the FRET changes before and after photobleaching, with the number of cells analyzed in parentheses. ***, p Ͻ 0.001. FIGURE 6. BMP4 exposure or CXXC5 overexpression strongly reduces the expression of the canonical Wnt signaling target Axin2 in neural stem cells. A, mRNA expression levels of Axin2 relative to control mRNA (TBP) in NSC cultures as assessed by RT-qPCR after FGF2 versus BMP4 exposure for 6 h. The bars represent average levels after three independent experiments in triplicates. B, mRNA expression levels of Axin2 relative to control (HPRT) after nucleofections with either empty vector or expression vector for CXXC5 after 4 h. The Wnt3a-mediated increase in expression of Axin2 decreases significantly after CXXC5 overexpression, but the Axin2 expression remains higher than in control cultures. The bars represent an average of two independent experiments in triplicates. CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling 3678 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284•NUMBER 6•FEBRUARY 6, 2009 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom monly used assay for studies of canonical Wnt signaling (11). Control experiments confirmed that reporter activity was robustly activated by overexpression of the canonical Wnt signaling mediator ß-catenin (Fig. 7A). Transfections of increasing amounts of CXXC5 revealed that 100 ng of CXXC5 significantly inhibited the TOPflash reporter activity induced by 50 or 100 ng of recombinant Wnt3a (Fig. 7B). In addition, a decrease in Wnt3a-induced activated (dephosphorylated) ß-catenin was observed in CXXC5-transfected cells, although no significant changes in the total protein levels of ß-catenin or Dvl could be detected (Fig. 7, C and D). Hence these experiments strengthen the suggestion of a role for CXXC5 in modulating canonical Wnt signaling activity. siRNA Against CXXC5 Attenuates the BMP4 Response and Facilitates the Wnt3a Response in NSCs—The results from the overexpression experiments allowed us to speculate whether a reduction of the endogenous levels of CXXC5 should attenuate the BMP4-mediated decrease in Axin2 expression levels and facilitate Wnt3a signaling effects in the NSCs. To study the role for CXXC5 in more physiological cellular conditions, we developed an siRNA strategy that after nucleofection reduced the levels of CXXC5 with 40–70% (see “Experimental Procedures”) but did not reduce Idax levels significantly in control cultures (Fig. 8A and data not shown). We then used this siRNA against CXXC5, named sir-CXXC5, in comparison with an unrelated siRNA against cyan fluorescent protein (sir-CFP) to investigate putative changes in the NSCs response to BMP4 and Wnt3a as assessed by RT-qPCR measurement of Axin2 gene expression levels after 4h exposure. Interestingly, administration of sirCXXC5 24 h before exposure of BMP4 resulted in an almost complete abolishment of the BMP4-mediated reduction of Axin2 levels (Fig. 8B). Cultures treated with sir-CXXC5 displayed around 10% reduction of Axin2 expression levels after BMP4 exposure, compared with an average of 60% reduction in cells that were nucleofected with sir-CFP (Fig. 8B). As the efficiency of the siRNA was significantly lower than 100% in our experiments (40–70%), a complete reversal of the BMP4/ Wnt3a-mediated effects was not expected. Accordingly, sirCXXC5 facilitated the increase in Axin2 levels in Wnt3a stimulated cultures compared with those cultures that received control siRNA (Fig. 8B). It should be noted that the Wnt3amediated increase in Axin2 was lower in all siRNA-transfected cultures than in any of the other experimental series, which we speculate is due to the siRNA treatment per se (40). We conclude that CXXC5 is a mediator of BMP4-mediated modulation of Wnt signaling in NSCs. FIGURE 7. CXXC5 overexpression attenuates Wnt3a-mediated TOPflash reporter activity. A, relative luciferase activity of a TOPflash reporter construct in HEK-293 cells after co-transfections with empty (left bar) or ß-catenin (right bar) expression vectors. Bars illustrate the average from two independent experiments in triplicates. B, relative luciferase activity of the TOPflash reporter construct in HEK-293 cells after treatment with 50 (left bars) or 100 (right bars) ng of purified Wnt3a and co-transfections with 0, 50 or 100 ng of CXXC5-containing expression vector. Note the marked decrease in reporter activityafterco-transfectionwith100ngCXXC5construct.Barsrepresenttwo independent experiments in triplicates. C, immunoblotting of total cell extracts from a fraction of the HEK-293 cells used in the corresponding luciferase experiments shown in B. No significant changes in total protein levels wereseenafterimmunoblottingwithß-cateninorDvlantibodies.ß-actinwas used as loading control. D, immunoblotting of total cell extracts corresponding to C using an antibody against activated (dephosphorylated) ß-catenin (ABC) compared with loading control (ß-actin). The relative increase in ABC induced by Wnt3a is decreased in cells transfected with CXXC5. CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling FEBRUARY 6, 2009•VOLUME 284•NUMBER 6 JOURNAL OF BIOLOGICAL CHEMISTRY 3679 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom DISCUSSION BMP4 treatment induces differentiation of NSCs in vitro into astrocytic and mesenchymal cells (4, 5) and is required for the proper development of the dorsal-most telencephalic structures, such as the hippocampal formation and the non-neuronal choroid plexus (1–3). In addition, BMP4 is in coordination with Wnt3a and FGF8 required for the correct size and regionalization of the dorsal pallium (1, 10). Yet little is understood regarding the cellular properties that allow and restrict these multiple and complex outcomes downstream of BMP4. In this study, we have identified CXXC5 as a novel potential target for cross-talk between BMP4 and canonical Wnt3a signaling as we show that CXXC5 is induced by BMP4 stimulation, selectively expressed in the BMP4-rich dorsal telencephalon, interacting with the Wnt signaling intermediate Dvl, and modulates the response to Wnt3a exposure in NSCs. The telencephalic choroid plexus extends from the ventricular zone into the ventricles and produces and secretes the cerebrospinal fluid. Many non-neural cell types can be found in the choroid plexus, including epithelial and mesenchymal cells. Interestingly while Wnt3a expression is excluded from the developing choroid plexus, both BMP4 and CXXC5 expression levels are high in this region (Fig. 2 and Ref. 41). It appears that while Wnt3a is required for neuronal progression from these regions contributing to hippocampal structures, Wnt3a seems dispensable for choroid plexus development (1). A possible functional role for CXXC5 in forebrain development could therefore be to repress the effects of long range Wnt3a signaling to allow BMP4-dependent formation of the choroid plexus. Future studies using conditional gene deletion strategies specifically targeting CXXC5 in the developing dorsal telencephalon will be required to test this hypothesis. The Xenopus homologue of the closely related factor Idax (CXXC4), named Xidax, has been shown to disrupt the correct formation of anterior brain structures when overexpressed in the developing Xenopus forebrain (13). This is not surprising considering its interaction with Dvl and thus interference with canonical Wnt signaling (11). The expression pattern of Idax in mammalian brain is not yet reported, but we have noted that Idax is expressed in telencephalic NSCs6 suggesting that it may be expressed also in the dorsal telencephalon. Although the sequence in the C-terminal parts of Idax and CXXC5 are very similar, the N-terminal domain shows clear divergences. Further, whereas Idax subcellular localization is preferentially cytoplasmic (11), CXXC5 is predominantly nuclear. It will therefore be of interest to investigate whether Idax and CXXC5 work synergistically, complementary, or antagonistically. It has been demonstrated that the CXXC domain in the founding members of the CXXC family, the MBD proteins, participate in DNA and protein interactions (12). In addition, several of the other CXXC domain containing proteins have been shown to participate in chromatin modifying mechanisms and regulate and interfere with the activity of chromatin modifying factors, such as Polycomb proteins (12). It is therefore possible that some of the effects of CXXC5 are due to additional function/s in the nucleus, and ongoing studies are focusing on elucidating a potential role for nuclear CXXC5. In summary, our results suggest that CXXC5 is a BMP4regulated modulator of Wnt signaling essential for proper response of telencephalic stem cells to canonical Wnt3a signaling. We speculate that the functional role of CXXC5 in telencephalic development may be to contribute to the establishment of the demarcation between BMP4 and Wnt3a signaling regions during the development of hippocampal and choroid plexus formations. Acknowledgments—We thank Per Uhle´n for advice on FRET, Lars Bjo¨rklund and the members of the Hermanson laboratory for discussions, John L. R. Rubenstein for constructs, and Christer Ho¨o¨g, Claes Wahlestedt, Ole Isacson, and Jamie Timmons for support. REFERENCES 1. Shimogori, T., Banuchi, V., Ng, H. Y., Strauss, J. B., and Grove, E. A. (2004) Development 131, 5639–5647 2. Panchision, D. M., Pickel, J. M., Studer, L., Lee, S. H., Turner, P. A., Hazel, T. G., and McKay, R. D. (2001) Genes Dev. 15, 2094–2110 3. Hebert, J. M., Mishina, Y., and McConnell, S. K. (2002) Neuron 35, 1029–1041 4. Tsai, R. Y., and McKay, R. D. (2000) J. Neurosci. 20, 3725–3735 5. Rajan, P., Panchision, D. M., Newell, L. F., and McKay, R. D. (2003) J. Cell Biol. 161, 911–921 6. van Grunsven, L. A., Verstappen, G., Huylebroeck, D., and Verschueren, K. (2005) Cytokine Growth Factor Rev. 16, 495–512 6 T. Andersson and O. Hermanson, unpublished observations. FIGURE 8. siRNA against CXXC5 modulates BMP4- and Wnt3a-mediated effects on endogenous Axin2 levels in neural stem cells. A, mRNA expression of CXXC5 relative to control (TBP) in NSC cultures receiving either a control siRNA (sir-CFP, see “Results”) or an siRNA directed against CXXC5 (sir-CXXC5) after exposure to FGF2, BMP4, or Wnt3a for 4 h. NSC cultures that have received sir-CXXC5 show significantly lower levels of CXXC5 mRNA. B, mRNA expression levels of Axin2 relative to control (TBP) in NSC cultures as assessed by RT-qPCR after transfections with either control siRNA (sir-CFP) or siRNA against CXXC5 (sir-CXXC5) after 4 h of exposure to FGF2, BMP4 or Wnt3a. Notably, NSCs that have received sir-CXXC5 show almost no down-regulation of Axin2 after BMP4 exposure. *, p Ͻ 0.05 compared with basal FGF2-treated control, n ϭ 6. CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling 3680 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284•NUMBER 6•FEBRUARY 6, 2009 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom 7. Xu, L. (2006) Biochim. Biophys. Acta 1759, 503–513 8. Dahlqvist, C., Blokzijl, A., Chapman, G., Falk, A., Dannaeus, K., Ibanez, C. F., and Lendahl, U. (2003) Development 130, 6089–6099 9. Kleber, M., Lee, H. Y., Wurdak, H., Buchstaller, J., Riccomagno, M. M., Ittner, L. M., Suter, U., Epstein, D. J., and Sommer, L. (2005) J. Cell Biol. 169, 309–320 10. Sur, M., and Rubenstein, J. L. (2005) Science 310, 805–810 11. Hino, S., Kishida, S., Michiue, T., Fukui, A., Sakamoto, I., Takada, S., Asashima, M., and Kikuchi, A. (2001) Mol. Cell. Biol. 21, 330–342 12. Katoh, M. (2004) Int. J. Oncol. 25, 1193–1199 13. Michiue, T., Fukui, A., Yukita, A., Sakurai, K., Danno, H., Kikuchi, A., and Asashima, M. (2004) Dev. Dyn. 230, 79–90 14. Johe, K. K., Hazel, T. G., Muller, T., Dugich-Djordjevic, M. M., and McKay, R. D. (1996) Genes Dev. 10, 3129–3140 15. Hermanson, O., Jepsen, K., and Rosenfeld, M. G. (2002) Nature 419, 934–939 16. Brunkhorst, A., Karlen, M., Shi, J., Mikolajczyk, M., Nelson, M. A., Metsis, M., and Hermanson, O. (2005) Mol Cell Neurosci 29, 250–258 17. Marko, N. F., Frank, B., Quackenbush, J., and Lee, N. H. (2005) BMC Genomics 6, 27 18. Smyth, G. K. (2004) Stat. Appl. Genet. Mol. Biol. 3, Article 3 19. Kawate, T., Allerson, C. R., and Wolfe, J. L. (2005) Org. Lett 7, 3865–3868 20. Schaeren-Wiemers, N., and Gerfin-Moser, A. (1993) Histochemistry 100, 431–440 21. Ossipova, O., Dhawan, S., Sokol, S., and Green, J. B. (2005) Dev. Cell 8, 829–841 22. Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007) J. Cell Sci. 120, 586–595 23. Miyakawa-Naito, A., Uhlen, P., Lal, M., Aizman, O., Mikoshiba, K., Brismar, H., Zelenin, S., and Aperia, A. (2003) J. Biol. Chem. 278, 50355–50361 24. Uhlen, P. (2003) Ann. N. Y. Acad. Sci. 986, 514–518 25. Wouters, F. S., Verveer, P. J., and Bastiaens, P. I. (2001) Trends Cell Biol. 11, 203–211 26. Ilkhanizadeh, S., Teixeira, A. I., and Hermanson, O. (2007) Biomaterials 28, 3936–3943 27. Sacchetti, P., Dwornik, H., Formstecher, P., Rachez, C., and Lefebvre, P. (2002) J. Biol. Chem. 277, 35088–35096 28. Sailer, M. H., Hazel, T. G., Panchision, D. M., Hoeppner, D. J., Schwab, M. E., and McKay, R. D. (2005) J. Cell Sci. 118, 5849–5860 29. Samanta, J., and Kessler, J. A. (2004) Development 131, 4131–4142 30. Allen, B. L., Tenzen, T., and McMahon, A. P. (2007) Genes Dev. 21, 1244–1257 31. Lee, C. S., Buttitta, L., and Fan, C. M. (2001) Proc. Natl. Acad. Sci. U. S. A. 98, 11347–11352 32. Martinelli, D. C., and Fan, C. M. (2007) Genes Dev. 21, 1231–1243 33. Biemar, F., Nix, D. A., Piel, J., Peterson, B., Ronshaugen, M., Sementchenko, V., Bell, I., Manak, J. R., and Levine, M. S. (2006) Proc. Natl. Acad. Sci. U. S. A. 103, 12763–12768 34. Itoh, K., Brott, B. K., Bae, G. U., Ratcliffe, M. J., and Sokol, S. Y. (2005) J. Biol. 4, 3 35. Leonard, J. D., and Ettensohn, C. A. (2007) Dev. Biol. 306, 50–65 36. Ille, F., Atanasoski, S., Falk, S., Ittner, L. M., Marki, D., Buchmann-Moller, S., Wurdak, H., Suter, U., Taketo, M. M., and Sommer, L. (2007) Dev. Biol. 304, 394–408 37. Yan, D., Wiesmann, M., Rohan, M., Chan, V., Jefferson, A. B., Guo, L., Sakamoto, D., Caothien, R. H., Fuller, J. H., Reinhard, C., Garcia, P. D., Randazzo, F. M., Escobedo, J., Fantl, W. J., and Williams, L. T. (2001) Proc. Natl. Acad. Sci. U. S. A. 98, 14973–14978 38. Lustig, B., Jerchow, B., Sachs, M., Weiler, S., Pietsch, T., Karsten, U., van de Wetering, M., Clevers, H., Schlag, P. M., Birchmeier, W., and Behrens, J. (2002) Mol. Cell. Biol. 22, 1184–1193 39. Jho, E. H., Zhang, T., Domon, C., Joo, C. K., Freund, J. N., and Costantini, F. (2002) Mol. Cell. Biol. 22, 1172–1183 40. Jepsen, K., Solum, D., Zhou, T., McEvilly, R. J., Kim, H. J., Glass, C. K., Hermanson, O., and Rosenfeld, M. G. (2007) Nature 450, 415–419 41. Furuta, Y., Piston, D. W., and Hogan, B. L. (1997) Development 124, 2203–2212 CXXC5 Is a Neural BMP4-induced Modulator of Wnt Signaling FEBRUARY 6, 2009•VOLUME 284•NUMBER 6 JOURNAL OF BIOLOGICAL CHEMISTRY 3681 atKarolinskainstitutetlibraryonFebruary3,2009www.jbc.orgDownloadedfrom Vítězslav Bryja, 2014    Attachments      #9      Witte F1 , Bernatik O1 , Kirchner K, Masek J, Mahl A, Krejci P, Mundlos S, Schambony A,  Bryja V*, Stricker S.* (2010): Negative regulation of Wnt signaling mediated by CK1‐ phosphorylated Dishevelled via Ror2. FASEB J. 24(7):2417‐26.  1  equal contribution, * corresponding authors      Impact factor (2010): 6.515  Times cited (without autocitations, WoS, Feb 21st 2014): 14  Significance:  The  first  study  showing  physical  interaction  of  Ror2,  a  Wnt  receptor  dedicated  to  non‐canonical  Wnt  signaling,  and  Dishevelled.  The  interaction  is  controlled  by  CK1  and  requires  C‐terminus  of  Dishevelled.  The  first  study  that  pointed towards an important regulatory role of Dishevelled C‐terminus.  Contibution of the author/author´s team: Biochemical description of the interaction,  functional in vitro data.                The FASEB Journal • Research Communication Negative regulation of Wnt signaling mediated by CK1phosphorylated Dishevelled via Ror2 Florian Witte,*,†,‡,1 Ondrej Bernatik,§,ʈ,1 Katharina Kirchner,¶ Jan Masek,§,ʈ Annika Mahl,*,†,‡ Pavel Krejci,§,ʈ Stefan Mundlos,*,† Alexandra Schambony,¶ Vitezslav Bryja,§,ʈ,2 and Sigmar Stricker*,†,2 *Development and Disease Group, Max Planck-Institute for Molecular Genetics, Berlin, Germany; † Institute for Medical Genetics, University Medicine Charite´, Berlin, Germany; ‡ Institut fu¨r Chemie/ Biochemie, Freie Universita¨t Berlin, Berlin, Germany; § Institute of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republic; ʈ Department of Cytokinetics, Institute of Biophysics, Academy of Sciences of the Czech Republic, Brno, Czech Republic; and ¶ Developmental Biology Unit, Biology Department, University of Erlangen-Nuremberg, Erlangen, Germany ABSTRACT Dishevelled (Dvl) is a multifunctional effector of different Wnt cascades. Both canonical Wnt3a and noncanonical Wnt5a stimulate casein-kinase-1 (CK1) -mediated phosphorylation of Dvl, visualized as electrophoretic mobility shift [phosphorylated and shifted Dvl (ps-Dvl)]. However, the role of this phosphorylation remains obscure. Here we report the functional interaction of ps-Dvl with the receptor tyrosine kinase Ror2, which is an alternative Wnt receptor and is able to inhibit canonical Wnt signaling. We demonstrate interaction between Ror2 and ps-Dvl at the cell membrane after Wnt3a or Wnt5a stimulus dependent on CK1. Ps-Dvl interacts with the C-terminal proline-serine-threonine-rich domain of Ror2, which is required for efficient inhibition of canonical Wnt signaling. We further show that the Dvl C terminus, which seems to be exposed in ps-Dvl and efficiently binds Ror2, is an intrinsic negative regulator of the canonical Wnt pathway downstream of ␤-catenin. The Dvl C terminus is necessary and sufficient to inhibit canonical Wnt/␤-catenin signaling, which is dependent on the presence of Ror2. Furthermore, both the Dvl C terminus and CK1␧ can inhibit the Wnt5a/Ror2/ATF2 pathway in mammalian cells and Xenopus explant cultures. This suggests that phosphorylation of Dvl triggers negative feedback regulation for different branches of Wnt signaling in a Ror2-dependent manner.—Witte, F., Bernatik, O., Kirchner, K., Masek, J., Mahl, A., Krejci, P., Mundlos, S., Schambony, A., Bryja, V., Stricker, S. Negative regulation of Wnt signaling mediated by CK1phosphorylated Dishevelled via Ror2. FASEB J. 24, 2417–2426 (2010). www.fasebj.org Key Words: Wnt/␤-catenin signaling ⅐ casein kinase ⅐ Wnt5a ⅐ Xenopus The wnt family of secreted signaling molecules, consisting of 19 members in mammals, plays multiple roles in embryonic development, adult homeostasis, and disease. In humans, deregulation of Wnt cascades is responsible for human diseases, including cancer, bone density defects, and limb malformations (1). Classically, Wnt ligands are divided into factors that can activate the so-called canonical branch, acting via stabilization of ␤-catenin or factors that activate several alternative pathways that have been collectively termed noncanonical cascades. Wnt molecules activating the canonical pathway, e.g., Wnt3a, bind surface receptors of the Frizzled and lipoprotein receptor-related protein (Lrp) families and induce the dispersal of a protein complex that targets ␤-catenin for proteasomal destruction. This leads to accumulation of ␤-catenin levels in the cell and nuclear translocation of ␤-catenin, where it interacts with transcription factors of the Tcf/Lef family to activate transcription (1, 2). Wnt5a, the founding member of the “noncanonical” Wnt class, was shown to activate or inhibit canonical Wnt signaling dependent on the presence of the coreceptor Ror2 (3). In addition, Wnt5a can signal to noncanonical pathways via Ror2. In Xenopus gastrulation, Wnt5a via Ror2 triggers a novel noncanonical pathway implicating PI3-kinase, Cdc42, JNK, and ATF2/c-Jun (hereafter referred to as Wnt5a/Ror2/ATF2 pathway) to regulate the expression of paraxial protocadherin (XPAPC) and coordinate convergent extension movements (4). Binding of Wnt5a to Ror2 results in dimerization and phosphorylation of Ror2 (5, 6). Ror2 is a receptor tyrosine kinase (RTK) consisting of conserved extracellular immunoglobulin (IG), cysteine-rich (CRD), and kringle (KR) domains, a cytoplasmic kinase domain, and, as a unique feature among RTKs, distally located proline- and serine/threoninerich (PST-rich) domains (7). A critical component of several Wnt-signaling branches is the cytoplasmic protein Dishevelled (Dvl). Dvl proteins carry an N-terminal DIX (Dishevelled/ 1 These authors contributed equally to this work. 2 Correspondence: S.S., Development and Disease Group, Max Planck Institute for Molecular Genetics, Ihnestr. 73, 14195 Berlin, Germany. E-mail: strick_s@molgen.mpg.de; V.B., Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotla´rˇska´ 2, 61137 Brno, Czech Republic. E-mail: bryja@sci.muni.cz doi: 10.1096/fj.09-150615 24170892-6638/10/0024-2417 © FASEB Axin) domain, a central PDZ (PSD-95; DLG, ZO1) domain, and a DEP (Dishevelled, EGL-10, Pleckstrin) domain, each with a variety of interaction partners. The C terminus of the Dvl proteins comprises ϳ220 aa [hereafter termed Dvl C terminus (Dvl-CT)], a domain for which no function has been described so far. Depending on Wnt stimulation and receptor context, Dvl can inhibit the ␤-catenin destruction complex, thereby stabilizing ␤-catenin and activating canonical Wnt signaling, or Dvl can regulate some of the noncanonical branches of Wnt signaling (8, 9). In response to Wnt ligands, several kinases, including casein kinases (CKs) CK1␦, CK1ε, and CK2, PKC, MAK, and PAR1, have been reported to phosphorylate Dvl, but the relevance of individual Dvl phosphorylation remains unclear (10–17). In this study, we report the interaction of Ror2 specifically with the C terminus of CK1-phosphorylated Dishevelled (ps-Dvl) and reveal that the Dvl-CT is an intrinsic negative regulator of the canonical Wnt-signaling cascade depending on the presence of Ror2. Inhibition of canonical Wnt signaling is not mediated via the Wnt5a/Ror2/ATF2 cascade, but CK1 and Dvl-CT also antagonize this pathway. These findings indicate a yet unappreciated role for CK1-phosphorylated Dvl, which mediates negative feedback control of canonical as well as noncanonical Wnt cascades. MATERIALS AND METHODS Constructs and antibodies Full-length (FL) and truncated mouse Ror2 constructs were described previously (18). We thank Yasuhiro Minami (Kobe University, Kobe, Japan) for Ror2–13S/T and Ror2–5YF constructs (19). Human Ror2 constructs were described previously (20). Dvl2-HA was kindly provided by Ulrich Stelzl (Max Planck-Institute for Molecular Genetics, Berlin, Germany), Dvl3-Flag, XWnt-8, XWnt5a, XDsh, and ␤-catenin constructs were provided by Randall T. Moon [Howard Hughes Medical Institute (HHMI), Seattle, WA, USA], and pCS2-Xror2 was provided by Maranori Taira (Tokyo University, Tokyo, Japan). We thank Shin-ichi Yanagawa (Kyoto University, Kyoto, Japan), Robert J. Lefkowitz (HHMI, Durham, NC, USA), and Jonathan M. Graff (University of Texas, Dallas, TX, USA) for expression vectors encoding Dvl2-Myc, CK1ε, and CK1ε (KϾR). CK2 constructs were kindly provided by Isabel Dominguez (Boston University, Boston, MA, USA). PAR1 expression constructs (13) were obtained from Olga Ossipova and Sergei Sokol (Mount Sinai School of Medicine, New York, NY, USA). The following antibodies were used in immunoprecipitation: Western blot or immunofluorescence: anti-Ror2 (R&D Systems, Minneapolis, MN, USA); anti-Dvl2 (sc-13974), antiDvl3 (sc-8027), anti-CK1ε (sc-25423), anti-JNK, anti-pJNK (sc-6254), anti-AKT1,2,3, anti-Myc (sc-40 for WB, sc-789 for IP) (Santa Cruz Biotechnology, Santa Cruz, CA, USA); antiAKTpS473 (ab18206), anti-pATF2 (ab 4734), anti-HA for IP (ab9110) (Abcam, Cambridge, UK); anti-HA for Western blot (H6908–2), anti-␤-actin (A5441), anti-Flag (M2 for WB, F7425 for IP) (Sigma-Aldrich, St. Louis, MO, USA); anti-CK2␣ (BD Biosciences, Franklin Lakes, NJ, USA), and anti-FL-ATF2 (Cell Signaling, Beverly, MA, USA). Secondary antibodies for immunofluorescence: AlexaFluor 568 and 488 (Invitrogen Molecular Probes, Carlsbad, CA, USA). Cell culture and immunofluorescence Hek293T, Cos-7, and Cos-1 cells were grown in 5–10% fetal calf serum (FCS, heat inactivated) according to standard protocols. For immunofluorescence, cells were transfected with either ExGen500 (Fermentas, Burlington, ON, Canada) or Polyfect (Qiagen, Hilden, Germany), with constructs as indicated. For immunofluorescence, 24 h after transfection, cells were starved 2 h in medium containing 1% FCS followed by stimulation with 200 ng/ml recombinant Wnt5a or Wnt3a (R&D Systems). D4476 (100 ␮M; Calbiochem, San Diego, CA, USA) was added together with Fugene (1 ␮l/200 ml; Roche, Mannheim, Germany) to enhance penetration. Cells were fixed with ice-cold methanol, and immunodetection was performed according to standard procedures. Pictures were taken with a Zeiss Axiovert 200M microscope using ApoTome optics (Zeiss, Jena, Germany) and processed with Axiovision software (Zeiss). For quantification, 100 cells in each of 3–5 independent experiments were photographed and analyzed for yellow membrane staining. For Wnt5a/Ror2 pathway analysis, Cos-1 cells were serumstarved for 16 h and stimulated with 100 ng/ml recombinant Wnt5a (R&D Systems) in SF-DMEM for the indicated time periods. Cell extracts were prepared using NOP buffer, protein content was quantified by BCA assay (Pierce, Rockford, IL, USA), and equal amounts were subjected to SDSPAGE and transferred to PVDF membranes. For CK1 inhibition, cells were pretreated with D4476/Fugene for 30 min before stimulation with Wnt5a. Coimmunoprecipitation and Western blotting Coimmunoprecipitation was performed in Cos-7 cells transfected with the indicated plasmids using polyethyleneimine as described previously (21). Immunoprecipitation and Western blotting were performed essentially as described previously (22). When required, Western blots were quantified by densitometry using ImageJ 1.42 software (U.S. National Institutes of Health, Bethesda, MD, USA). Dual luciferase assay HEK293 cells were transfected with mixture of SuperTopFlash (100 ng), pRL-TK (50 ng, internal control), and other plasmids as indicated, in a 24-well plate using polyethyleneimine, as described previously (21). Cells were lysed 24 h after transfection and further processed using the Dual-Luciferase® Reporter 1000 Assay System (Promega, Madison, WI, USA) according to the manufacturerЈs instructions and measured on a luminometer. Confirmation of the expression levels of the proteins overexpressed in the reporter assays was assessed in parallel experiments by Western blotting and detection with appropriate antibodies (see Supplemental Fig. 5). siRNA treatment HEK293 cells were transfected with siRNA using retrofection according to the manufacturerЈs instructions (Ambion, Austin, TX, USA). We mixed siRNAs (0.55 ␮l of 20 ␮M siRNA) with lipofectamine 2000 (1.45 ␮l) and OptiMEM (48 ␮l) and incubated them at room temperature for 20 min. The transfection mixture was added to a 24-well plate and mixed with a suspension of freshly trypsinized HEK293 cells (30,000 cells/well in 300 ␮l of complete medium). This procedure results in a final siRNA concentration of 30 nM. Transfection was terminated after 5 h by changing medium. After 24 h, 2418 Vol. 24 July 2010 WITTE ET AL.The FASEB Journal ⅐ www.fasebj.org cells were transfected with appropriate plasmids according to the transfection scheme. At 24 h posttransfection (48 h after siRNA transfection), cells were collected for further analysis. siRNAs against Ror2 were purchased from Ambion (I, catalog no. s9758; II, s9759; III, s9760). Silencer® Negative Control siRNA (4635; Ambion) was used as negative control. The efficiency of the silencing was assessed by Western blotting. Injection and analysis of Xenopus laevis embryos Capped mRNAs were transcribed from linearized DNA templates (pCS2-XDsh, psp64T-XWnt-8, psp64T-␤-catenin, pcDNA-Ror2Flag, pcDNA-Ror2⌬745Flag, pcDNA-Dvl3, pcDNA-Dvl3⌬C, pcDNADvl3CT, pCS2-CK1ε, pCS2-XWnt5a, pCS2-XRor2) using mMessage mMachine Kits (Ambion). Eggs from human chorionic gonadotropin-treated females were fertilized by standard methods and staged according to Nieuwkoop and Faber (23). Morpholino and RNA were injected into both cells of 2-cell-stage embryos or the marginal zone of the ventral or dorsal blastomeres of 4-cell-stage embryos in a total volume of 4 nl. If not mentioned otherwise in the text, the following amounts were injected: XWnt-8, 8 pg; ␤-catenin, 100 pg; XDsh and Dvl3-C-term, 500 pg; Ror2 and Ror2⌬745, 200 pg; XWnt5a 100 pg, XRor2, 3 pg; CK 1ε, 5 pg. For axis duplication assays, embryos were cultivated till stage 24 and scored for secondary body axes. For real-time RT-PCR, Animal Cap explants were prepared at stage 8 and cultivated for 4 h in Barth’s solution in the absence or presence of 10 ␮M D4476. Real-time RT-PCR was carried out as described previously (4). In situ hybridization Two-cell-stage embryos were injected in one blastomere with 80 pg ␤-galactosidase RNA and MO or RNA as indicated. After fixation, the embryos were stained for ␤-galactosidase activity. Whole-mount in situ hybridizations were carried out by using the Digoxygenin/Alkaline Phosphatase detection system (Roche) as described previously (24). RESULTS AND DISCUSSION Ror2 interacts via its PST-rich domain with the Dvl-CT Ror2 was previously implicated in canonical as well as noncanonical Wnt signaling (3, 4, 25–27). We therefore tested whether Dvl, a protein that participates in both signaling cascades, could interact with Ror2. Using coimmunoprecipitation, we found that mouse Ror2 can interact with mouse Dvl2 (Fig. 1A and Supplemental Fig. 1). This interaction was abolished by deletion of C-terminal parts of Ror2 (Ror2⌬745 and Ror2⌬469), both representing mutations found in human brachydactyly type 1 (BDB1) (28), but was not dependent on the 2 extracellular domains for ligand and coreceptor interaction (Ror2⌬CRD/KR) (Fig. 1A and Supplemental Fig. 1). Thus, it is likely that the interaction is direct and not mediated by secondary interaction of Ror2 with Frizzled proteins, which have been shown to bind each other via their CRDs (27). Both BDB1 mutants of Ror2 cannot interact with Dvl, indicating that this interaction could be part of the molecular pathogenesis of BDB1. To map the interaction domain on Dvl, we used Cand N-terminally truncated forms of human Dvl3 (29) (Fig. 1B). HDvl3 also interacted with the Cterminal PST-rich domain of mRor2 (Supplemental Fig. 1C). Mouse Ror2 was coimmunoprecipitated with all Dvl3 mutants to a variable degree. The weakest interaction was observed between Ror2 and FL Dvl3. The most efficient interaction was found between the C terminus of Dvl3 (albeit very weakly expressed) and Ror2 (Fig. 1B and Supplemental Fig. 1, quantification in Supplemental Table 1). The truncations containing either the DIX or the DEP domains but not the PDZ domain bound with intermediate efficiency. The most probable explanation of these results is that Dvl coimmunoprecipitates with Ror2 via 2 binding sites. A high-affinity binding site localizes to the C-terminal domain of Dvl, whereas other low-affinity binding sites appear to exist outside the C terminus. Notably, the DIX, PDZ, and DEP domains of Dvl appear to be dispensable for highaffinity binding to Ror2. Figure 1. Interaction of Ror2 and CK1ε-phosphorylated ps-Dvl. A, B) mRor2 (A) and hDvl3 (B) constructs used. Interaction is indicated by ϩ or Ϫ. Ror2 interacts via its C-terminal proline/ serine/threonine-rich domain with Dvl2. A Ror2 mutant that lacks the extracellular kringle and cystein-rich domains (Ror2⌬CRD/KR) still interacts with Dvl2. Ror2 predominantly interacts with the C terminus of Dvl3. Note that interaction with full-length Dvl is weak. CRD, cysteine-rich domain; IG, immunoglobulin domain; KR, kringle domain; P, proline-rich domain; S/T, serine/threonine-rich domain; TK, tyrosine kinase domain; TM, transmembrane domain. C) Ror2 specifically interacts with Dvl phosphorylated by CK1ε. Phosphorylation of Dvl by CKIε results in a visible size shift of the protein. Addition of CKIε increases the pulldown efficiency of Dvl3 after Ror2 immunoprecipitation and vice versa. Note that specifically the upper, phosphorylated band of Dvl3 is enriched in the Ror2 pulldown experiment. CK1ε is also coprecipitated with either Ror2-HA or Dvl3-Flag. D) CK1ε but not CK2␣, CK2␤, or Par1 increases interaction of Ror2 and Dvl3. E) Phosphorylation of Dvl by CK1ε increases binding of full-length Dvl3 (construct 1) to Ror2, while binding of constructs 4 and 7 remains unaffected. 2419INTERACTION OF CK1-PHOSPHORYLATED DVL WITH Ror2 Ror2 specifically interacts with CK1-phosphorylated Dvl On Wnt stimulus, Dvl proteins become phosphorylated, which can be seen as a shift of the apparent molecular weight (ps-Dvl). One major Dvl-specific kinase that stimulates the formation of ps-Dvl is CK1ε (11, 17, 22, 30, 31). To examine the influence of Dvl phosphorylation status on the interaction with Ror2, we coexpressed Ror2, Dvl3, and CK1ε in Cos-7 cells and performed coimmunoprecipitation for Ror2 or Dvl3. Without CK1ε, Ror2-immunoprecipitation resulted in a weak pulldown of Dvl3 (Fig. 1C, black arrowhead). Coexpression of CK1ε strongly increased the interaction between Ror2 and Dvl3. Interestingly, ps-Dvl3 was the predominant form found in the Ror2 pulldown (Fig. 1C, white arrowhead). Vice versa, immunoprecipitation of Dvl3 yielded a weak pulldown of Ror2 and was boosted 3 times by the coexpression of CK1ε (Fig. 1C, values for input were compared with values for pulldown; see Supplemental Table 1 for densitrometric data). Thus, these results provide evidence that Ror2 preferentially binds to CK1ε-phosphorylated ps-Dvl. We observed the increase in Ror2/Dvl3 interaction only with CK1ε but not with other kinases known to phosphorylate Dvl, such as CK2 or PAR1 (Fig. 1D), indicating that this is specific for CK1ε. To further elucidate the putative mechanism of CK1-stimulated interaction between Ror2 and Dvl, we used full-length Dvl3 and 2 complementary truncated versions, the Dvl3-CT and Dvl3⌬C (constructs 1, 4, and 7 in Fig. 1B). In the absence of CK1ε, Ror2 showed weak interaction with wild-type (wt) Dvl3 and Dvl3⌬C and more efficient binding with Dvl3-CT, as demonstrated above. Coexpression of CK1ε resulted in a highly enhanced interaction with full-length Dvl3 (4.2ϫ increase; pulldown relative to input), but had no effects on the interaction with Dvl3-CT and Dvl3⌬C (Fig. 1E). These results support the possibility that facilitated interaction between Ror2 and full-length Dvl in the presence of CK1ε could be due to phosphorylation-induced conformational changes that result in a better accessibility of the Dvl-CT domain. Alternatively, CK1ε -mediated phosphorylation of serine/threonine residues in Dvl3 might be directly required for the high-affinity interaction with Ror2; however, the observation that CK1ε coexpression had no effect on Dvl3-CT and Dvl3⌬C argue against this. Therefore, our data suggest that CK1ε phosphorylation-induced conformational changes expose the C-terminal binding interface of Dvl, leading to strong interaction between ps-Dvl and Ror2. Wnt stimulus induces Ror2/ps-Dvl interaction at the membrane dependent of CK1 Phosphorylation of Dvl by CK1ε is seen in response to Wnt5a (noncanonical) as well as Wnt3a (canonical) signaling (22, 32, 33), As shown in Fig. 2A, the interaction of endogenous Ror2 with Dvl3-Flag was markedly enhanced by addition of recombinant Wnt5a. This demonstrates that Dvl3 can more efficiently bind to Ror2 not only after overexpression of CK1ε but also after Wnt5a stimulus, which is one physiological activator of CK1ε activity toward Dvl (22). To identify where in the cell the Wnt5a-dependent interaction between Ror2-Dvl is taking place, we transfected untagged human Ror2 (20) into Cos-1 cells and monitored the subcellular localization of endogenous Dvl2. Dvl proteins locate in small cytoplasmic puncta and were shown to be recruited to the membrane in noncanonical PCP signaling (8, 34), as well as during canonical Wnt signaling in Lrp6 signalosomes (35, 36). Without stimulation, Ror2 localized to the membrane, and endogenous Dvl2 showed a granular cytoplasmic distribution (Fig. 2B). After 30 min of stimulation with recombinant Wnt5a, a robust signal for Dvl2 was present at the membrane that colocalized with Ror2 in a punctate pattern (Fig. 2C; quantifications shown in Fig. 2E). This association of Dvl2 with Ror2 at the membrane was also clearly detectable after Wnt3a stimulus (Fig. 2D). Using the intracellular truncated Ror2 reflecting the human BDB1 mutation Ror2W749X, we were not able to detect Wnt5a- or Wnt3ainduced colocalization of Dvl2 and Ror2 at the membrane (Fig. 2F and Supplemental Fig. 2). Taken together, these results for the first time visualize an interaction of full-length Ror2 with Dvl2 at the cell membrane induced by either Wnt5a or Wnt3a. Based on the coimmunoprecipitations, we assume that the fraction of Dvl2 that colocalizes with Ror2 at the cell membrane represents ps-Dvl2. Indeed, it was possible to induce such colocalization by cotransfection of CK1ε (Fig. 2G). The involvement of endogenous CK1 in the Wnt5a-induced colocalization of Dvl2 and Ror2 was verified by the use of the CK1specific inhibitor D4476, which blocks both ps-Dvl2 formation (22) and double-positive Dvl2/Ror2 punctae formation at the membrane (Fig. 2H). Similar results were obtained after Wnt3a stimulus (Supplemental Fig. 2). This indicated that the punctae represented an interaction of Ror2 with ps-Dvl phosphorylated by CK1ε; however, effects of other CK1 isoforms cannot be ruled out. CK1␧ kinase activity phosphorylating Dvl, but not Ror2, is required for the interaction of Dvl and Ror2 at the cell membrane A kinase-deficient variant of CK1ε (CK1ε-KD) was not able to increase coprecipitation of Dvl with Ror2 (Fig. 1D). We thus tested whether the kinase activity of CK1ε is also required for recruitment of Dvl to the cell membrane by Ror2. Although expression of CK1ε induced a strong overlap between Dvl2 and Ror2 at the membrane, CK1ε-KD was not able to do so (Fig. 3A, B). Notably, CK1ε-KD appeared to function as a dominant-negative in this context, because it interfered with the colocalization of Dvl2 and Ror2 induced by either Wnt3a or Wnt5a (Fig. 3A, B), further emphasizing the importance of CK1ε for Wnt-induced Dvl/ Ror2 interaction. CK1ε interacts with both Ror2 and Dvl (17, 19); hence it is conceivable that the interaction between Ror2 and Dvl is mediated by CK1ε. However, the 2420 Vol. 24 July 2010 WITTE ET AL.The FASEB Journal ⅐ www.fasebj.org CK1ε-KD isoform interacts with both Ror2 and Dvl with comparable efficiency as wt CK1ε (19, 22). Because CK1ε-KD neither increases Dvl pulldown after precipitation of Ror2 (Fig. 1D) nor induces Dvl2/Ror2 colocalization at the cell membrane, this appears unlikely. CK1ε has been previously described as a priming kinase for Ror2 tyrosine kinase activity, phosphorylating Ror2 on serine/threonine residues in the PST domain (19). To exclude the possibility that phosphorylation of Ror2 by CK1ε might be required for the interaction with ps-Dvl, we performed pulldown assays with a Ror2 mutant, where 13 serine and threonine residues in the PST domain have been changed to alanine (Ror2–13S/T), which prevents phosphorylation by CK1ε (19). Following CK1ε overexpression, this mutant is coimmunoprecipitated with Dvl2-Myc with the same efficiency as wt Ror2 (Fig. 3C). Kani et al. (19) postulated that Ror2 is in a complex with Frizzled proteins at the cell membrane, which mediate activation of CK1ε. So far it is unclear whether binding of the Wnt ligand to Ror2 is also involved in Figure 2. Wnt ligands induce Ror2-Dvl interaction at the cell membrane via a CK1-dependent mechanism. A) IP of endogenous Ror2 from Cos-7 cells results in efficient pulldown of Dvl3-Flag only after treatment with Wnt5a. B–D) Wnt5a and Wnt3a lead to recruitment of endogenous Dvl2 to the cell membrane in a punctate pattern overlapping with Ror2 expression in Cos-1 cells. Untreated cells (B) and cells treated with Wnt5a (C) or Wnt3a (D) were transfected with hRor2 and immunolabeled with anti-Ror2 and anti-Dvl2 antibodies. E) Quantification of Dvl/Ror2 membrane staining; percentage of cells showing overlapping surface staining for Ror2 and Dvl2. Data are derived from 3 independent experiments. Error bars ϭ se. F) Truncated BDB1 mutant of human Ror2 (Ror2-W749X) is unable to recruit Dvl to the membrane. G, H) CK1ε overexpression induces association of Ror2 and Dvl2 (G), while CK1 inhibitor D4476 inhibits positive effects of Wnt5a on Ror2 and Dvl2 association at the cell membrane (H). 2421INTERACTION OF CK1-PHOSPHORYLATED DVL WITH Ror2 this process. As shown above, Dvl interacts with Ror2⌬CRD/KR, which lacks the extracellular ligand binding domains. In Cos-7 cells we also observed Ror2⌬CRD and Dvl2 colocalization at the membrane after Wnt5a and Wnt3a stimulus (Supplemental Fig. 3). This indicates that the Ror2/ps-Dvl association is independent of Wnt binding to Ror2, supporting the notion that activation of CK1ε is mediated by Wnt/ Frizzled signaling. Taken together it is likely that Ror2 is located at the cell membrane in association with Wnt/ Frizzled/CK1ε complexes and can bind ps-Dvl emanating from Wnt/Frizzled signaling. Ror2 inhibits canonical Wnt signaling in vitro and in vivo dependent on its C terminus It is known that Ror2 can inhibit canonical Wnt signaling (3, 37, 38), although the mechanism of this inhibition remains somewhat elusive. We thus analyzed at which level Ror2 can inhibit the canonical pathway using TOPflash assays in HEK 293 cells. TOPflash activity induced by overexpression of Dvl2 was further potentiated by coexpression of CK1ε. Full-length Ror2 was able to inhibit reporter activity induced by Dvl2 ϩ CK1ε or Dvl3 ϩ CK1ε by ϳ70% (Fig. 4A and Supplemental Fig. 4). This was not increased by addition of recombinant Wnt5a (data not shown), indicating that overexpression of Ror2 alone was sufficient to activate downstream signaling. The truncated Ror2 mutant lacking the PST-rich domain, Ror2⌬745, showed a strongly reduced inhibitory potential (Fig. 4A and Supplemental Fig. 4). The Ror2–13ST/A mutant, which cannot be phosphorylated by CK1ε, and the 5YF mutant, which shows no autophosphorylation on CK1ε stimulus (19), both exhibited an inhibitory potential comparable to wt Ror2 in TOPflash assays (Fig. 4A and Supplemental Fig. 4), indicating that phosphorylation and activation of Ror2 by CK1ε is not required for the inhibition of canonical Wnt signaling. Stimulation of TCF/LEF-reporter activity with constitutively active (ca-) ␤-catenin obtained similar results (Fig. 4B), indicating that in mammalian cells Ror2 can inhibit Wnt/␤-catenin signaling downstream of ␤-catenin. We then wanted to test whether Ror2 is able to antagonize canonical Wnt-signaling in vivo. Axis duplication assays in Xenopus embryos are a well-established readout for canonical Wnt signaling. XWnt8a mRNA injection on the ventral side of the embryo robustly induced secondary axes in Xenopus embryos, which was inhibited by coinjection of mRor2-wt mRNA, confirming that Ror2 acts as an antagonist of canonical Wnt signaling in this assay (Fig. 4C). XDvl induced secondary axes to a much lower percentage, which was still inhibited by coinjection of mRor2 to a significant extent. Ror2⌬745, in contrast to wt mRor2, was not able to bind XDvl in coimmunoprecipitation experiments (Supplemental Fig. 6), and concordantly had no significant effect on axis induction by XWnt8a or XDvl (Fig. 4C). Ror2 was unable to significantly inhibit secondary axes induced by injection of X␤-catenin in this assay, although wt Ror2 (and not Ror2⌬745) induced a slight decrease, which suggests that a similar mechanism might act in both Xenopus and mammalian cells. Both data sets congruently show that the PST domain of Ror2 is essential for the efficient inhibition of canonical Wnt signaling, supporting the assumption that the interaction of ps-Dvl with the Ror2-PST might be involved in antagonizing canonical Wnt signaling. Dvl-CT is a negative regulator of canonical Wnt signaling dependent on Ror2 We then tested the role of the Dvl-CT in the canonical Wnt pathway using the TOPflash assay. To this end we stimulated TOPflash activity with either Dvl3 Figure 3. CK1ε kinase activity is required for Dvl2/Ror2 interaction. A) Immunolabeling for transfected Ror2 and endogenous Dvl2 on Cos-1 cells. Transfection of CK1ε but not of CK1ε-KD induces Dvl2/Ror2 colocalization at the membrane. CK1ε-KD also interferes with Wnt3a- or Wnt5a-induced Dvl2/Ror2 colocalization. B) Quantification of Dvl/Ror2 membrane staining; percentage of cells showing overlapping surface staining for Ror2 and Dvl2. Data are derived from 3 independent experiments. Error bars ϭ se. C) Equal interaction of Dvl2 with wt-Ror2 or a Ror2 mutant that cannot be phosphorylated by CK1ε (Ror2–13S/T). Cells were transfected with indicated constructs; lysates were subjected to immunoprecipitation against Dvl2-Myc. 2422 Vol. 24 July 2010 WITTE ET AL.The FASEB Journal ⅐ www.fasebj.org or Dvl3⌬C in the presence or absence of CK1ε. Both Dvl3 constructs alone induced only a weak TOPflash signal (Fig. 5A). Coexpression of CK1ε together with Dvl3 strongly promoted TOPflash activity, and, interestingly, CK1ε boosted the activity of Dvl3⌬C much more efficiently than it did for wt Dvl3 (Fig. 5A). Next, we tested whether the Dvl-CT alone, which is able to bind Ror2 (Fig. 1B, E), is sufficient to inhibit the canonical Wnt pathway. In TOPflash assays, the coexpression of Dvl3-CT inhibited reporter activation induced by Dvl3/CK1ε in a dose-dependent manner (Fig. 5B). This indicates that in full-length Dvl, the C terminus, exposed by CK1ε-mediated phosphorylation, may act as a domain mediating a negative feedback on canonical signaling, potentially by its binding to Ror2. In the next step we wanted to know whether Ror2 is necessary for the inhibitory effect elicited by the Dvl-CT. We tested 3 different siRNAs directed against human ROR2. All 3 siRNAs efficiently reduced the expression of ROR2 in HEK293 cells (Supplemental Fig. 7A). Using all 3 siRNAs, knockdown of Ror2 alone increased reporter activity in HEK293 cells stimulated with Dvl3 ϩ CK1ε significantly (Fig. 5C, Supplemental Fig. 7B), demonstrating that endogenous Ror2 is an intrinsic negative regulator of the canonical pathway. Moreover, knockdown of Ror2 completely abolished the negative regulatory potential of Dvl-CT (Fig. 5C). Notably, Dvl-CT was also able to inhibit TOPflash activation stimulated by Dvl⌬C ϩ CK1ε (Fig. 5D) or ca-␤-catenin (Fig. 5E). Again, the inhibitory effects of Dvl-CT on TCF/LEF reporter activity driven by ca-␤-catenin were diminished by Ror2 knockdown (Fig. 5F). These results suggest that the inhibitory effect of Dvl-CT is not mediated via direct competition with FL Dvl for downstream effectors, but rather that Dvl-CT, mimicking the exposed C terminus in ps-Dvl, triggers a Ror2-dependent pathway, which inhibits canonical Wnt signaling downstream of ␤-catenin in mammalian cells. CK1/ps-Dvl negatively regulates the Wnt5a/Ror2/ ATF2 pathway Ror2 was described as an inhibitor of canonical Wnt/␤catenin signaling (3, 24, 37, 38), but it can also activate alternative Wnt cascades. In Xenopus, XWnt5a triggers a noncanonical cascade via XRor2, utilizing Akt, JNK, and ATF2 (hereafter termed Wnt5a/Ror2/ATF2 pathway), leading to the expression of XPAPC (4). In parallel, XWnt5a also leads to the production of ps-Dvl (39). To test the role of the ps-Dvl/Ror2 interaction for the Wnt5a/Ror2/ATF2 pathway, we first used Xenopus Animal Cap explants and investigated the influence of CK1ε and hDvl3 or Dvl3-CT (which is supposed to mimic the effect of ps-Dvl) on XPAPC expression. XPAPC is not expressed in Animal Caps endogenously, but its expression is induced by overexpression of XWnt5a (4) and more robustly by coinjection of XWnt5a and low amounts of XRor2 RNA. CK1ε overexpression did not induce XPAPC expression in AC explants, but significantly inhibited XPAPC upregulation in response to XWnt5a/XRor2 (Fig. 6A). Treatment with the CK1 inhibitor D4476 had no effect on basal or XWnt5a/XRor2-induced XPAPC levels. Concordantly, coinjection of hDvl3 ϩ CK1ε completely blocked XPAPC up-regulation. Expression of hDvl3-CT alone abolished XPAPC up-regulation. This was not significantly affected by D4476 or coinjection of CK1ε, although in the latter we observed slightly elevated XPAPC levels (Fig. 6A). In accordance with the assumption that Dvl-CT mimics the effects of ps-Dvl, overexpression of Figure 4. Inhibition of canonical Wnt signaling by Ror2 in vitro and in vivo. A, B) TOPflash reporter assays in HEK293 cells stimulated with Dvl2 ϩ CK1ε (A) or ␤-catenin (B). In mammalian cells, Ror2 inhibits canonical Wnt signaling induced by either Dvl2 ϩ CK1ε or ␤-catenin; C terminus of Ror2 is required for this inhibition. Ror2 mutants defective for serine/threonine or tyrosine phosphorylation (Ror2– 13S/T, Ror2–5YF) show inhibitory activity comparable to wt-Ror2. C) Top panel: Ror2, but not C-terminal truncated Ror2⌬745, significantly inhibits secondary axes induced by XWnt-8 or XDvl, but not axis induction by X␤-catenin. Bottom panel: embryos injected with either XWnt-8 alone or XWnt-8 and Ror2 or Ror2⌬745. Chart shows averages of Ն3 independent experiments. N, total numbers of embryos analyzed. Error bars ϭ sem. *P Ͻ 0.05, **P Ͻ 0.01, ***P Ͻ 0.001; Student’s t test. 2423INTERACTION OF CK1-PHOSPHORYLATED DVL WITH Ror2 hDvl3-CT also abolished XWnt5a/XRor2-induced phosphorylation of endogenous ATF2 in Xenopus embryos (Fig. 6B). These effects on XPAPC gene expression were confirmed by XPAPC in situ hybridization of singleside injected embryos (Fig. 6C, D). Knockdown of XWnt5a down-regulated XPAPC, as shown previously (4). Similar down-regulation was observed after CK1ε overexpression, while expression of a dominantnegative (dn-) CK1ε up-regulated XPAPC on the injected side in 32% of the embryos (Fig. 6C, D). With this construct, however, we observed also reduced (18%) or mislocalized (25%) expression, indicating not unexpectedly that dnCK1ε does not act specifically on the regulation of XPAPC expression but also affects other processes. Overexpression of hDvl3CT induced strong gastrulation defects. The embryos failed to close the blastopore and showed necrotic patches in the organizer region (Fig. 6C). We also observed severely reduced XPAPC expression in these embryos; however, with the disruption of development we are unable to clearly attribute this phenotype to a pathway-specific inhibitory effect. We next analyzed Cos-1 cells, which express endogenous Ror2, and found that Wnt5a stimulation induced a rapid, transient increase of pAKT (a PI3K target), pJNK, and pATF2 (Fig. 6E). This is reminiscent of the Wnt5a/Ror2/ATF2 pathway that we previously defined in Xenopus (4) and indicates that recombinant Wnt5a stimulates the same Ror2/ATF2 signaling cascade also in mammalian cells. Interestingly, the phosphorylation of AKT, JNK, and ATF2 already occurs after 15 min of Wnt5a stimulation (Fig. 6E). An increase of endogenous ps-Dvl was observed in our assay between 30 and 60 min of Wnt5a stimulation and persisted until 120 min of stimulation (Fig. 6F), comparable to previous observations (22). Conversely, pAKT, pJNK, and pATF2 decreased back to basal levels in the same time interval between 30 and 60 min (Fig. 6E). We thus hypothesized that the interaction of ps-Dvl with Ror2 might represent a negative regulatory mechanism of Wnt5a/Ror2/ATF2 signaling, as suggested by the timely correlation of ps-Dvl appearance and pATF2 decrease. In this line, inhibition of CK1 (i.e., inhibition of ps-Dvl formation) should relieve this inhibition. Indeed, stimulation of Cos-1 cells with Wnt5a and concomitant inhibition of CK1 with D4476 resulted in a prolonged activation of the pathway, because pAKT and pATF2 could be detected even after 60 and 120 min of stimulation in the presence of D4476 (Fig. 6F). In these samples the formation of ps-Dvl was not detectable even after 120 min of stimulation, supporting the hypothesis that the activation of CK1 and the formation of ps-Dvl could represent a negative autoregulatory loop. Our data provide the first evidence that the appearance of ps-Dvl, most likely emanating from Wnt/Frizzled signaling, might be a general negative feedback Figure 5. Dvl-CT is an intrinsic negative regulator of Wnt/␤-catenin signaling dependent on Ror2. A) TOPflash assay stimulated with Dvl3 and CK1ε. CK1ε increases reporter activity stimulated by wt Dvl3 and strongly increases activity induced by Dvl lacking the C terminus, Dvl3⌬C. Dvl-CT has no stimulating activity alone or in combination with CK1ε. B–F) Isolated C terminus of Dvl3 (Dvl3-CT) negatively regulates canonical Wnt signaling induced by Dvl3 ϩ CK1ε (B), Dvl3⌬C ϩ CK1ε (D), or ␤-catenin (E) in a dose-dependent manner (0.2– 0.8 ␮g); siRNA-mediated knockdown of Ror2 increases TOPflash activation by Dvl3 (C) or ␤-catenin (F). Inhibitory effect of Dvl3-CT is abrogated by siRor2, demonstrating that the inhibitory function of the Dvl-CT is dependent on the presence of Ror2. Charts show combined data from Ն3 independent experiments. Error bars ϭ sem. *P Ͻ 0.05, **P Ͻ 0.01, ***P Ͻ 0.001 vs. Dvl3 ϩ CK1ε (B), Dvl3⌬C ϩ CK1ε (D), ␤-catenin (E); Student’s t test. 2424 Vol. 24 July 2010 WITTE ET AL.The FASEB Journal ⅐ www.fasebj.org mechanism for different branches of Wnt signaling acting via interaction with Ror2. The authors thank Kathrin Seidel for excellent technical assistance. This work was funded by grants from the Deutsche Forschungsgemeinschaft (DFG) to S.S. and S.M. (SFB 577) and to A.S. (SCHA965/2–3 and 6–1). V.B. is supported by MSM0021622430 (Ministry of Education, Youth and Sports of the Czech Republic), KJB501630801, AVOZ50040507, AVOZ50040702 (Academy of Sciences of the Czech Republic), 204/09/H058, 204/09/0498, 204/ 09/J030 (Czech Science Foundation), and an EMBO installation grant. A.M. is supported by the Berlin-Brandenburg School for Regenerative Therapies (BSRT). Figure 6. CK1ε and Dvl-CT antagonize the Wnt5a/Ror2/ATF2 pathway. A) Relative expression of the XWnt5a/XRor2 target XPAPC in Xenopus Animal Cap explants injected with indicated constructs. Chart shows averages of Ն3 independent experiments. Error bars ϭ se. *P Ͻ 0.05 vs. XWnt5a/XRor2; t test. B) Dvl3-CT is able to inhibit ATF2 phosphorylation in Xenopus embryos. XWnt5a and XWnt5a ϩ XRor2 induce phosphorylation of ATF2, which is prevented by coinjection of Dvl3-CT. Morpholinos against XWnt5a or XRor2 serve as controls. C) In situ hybridization for XPAPC on Xenopus embryos injected in one blastomere at the 2-cell stage with mRNA for LacZ and constructs as indicated. Injected side is identified by LacZ staining and oriented to the left. D) Quantification of XPAPC expression determined by ISH. N, numbers of embryos analyzed. E) Stimulation of Cos-1 cells with recombinant Wnt5a leads to rapid activation of PI3-kinase after 15 min (monitored by AKT phosphorylation), phosphorylation of JNK and ATF2 indicating activation of the Wnt5a/Ror2/ATF2 pathway. Note that pAKT, pJNK, and pATF2 decrease back to basal levels between 30 and 60 min. F) Phosphorylation of Dvl2 and Dvl3 appears between 30 and 60 min of Wnt5a treatment, which is prevented by addition of CK1 inhibitor D4476. Treatment with D4476 results in prolonged activation of the Wnt5a/Ror2/ATF2 pathway monitored by AKT and ATF2 phosphorylation. 2425INTERACTION OF CK1-PHOSPHORYLATED DVL WITH Ror2 REFERENCES 1. MacDonald, B. T., Tamai, K., and He, X. (2009) Wnt/beta-catenin signaling: components, mechanisms, and diseases. Dev. Cell 17, 9–26 2. Angers, S., and Moon, R. T. (2009) Proximal events in Wnt signal transduction. Nat. Rev. Mol. Cell. Biol. 10, 468–477 3. 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 4. Schambony, A., and Wedlich, D. (2007) Wnt-5A/Ror2 regulate expression of XPAPC through an alternative noncanonical signaling pathway. Dev. Cell 12, 779–792 5. Yamamoto, H., Yoo, S. K., Nishita, M., Kikuchi, A., and Minami, Y. (2007) Wnt5a modulates glycogen synthase kinase 3 to induce phosphorylation of receptor tyrosine kinase Ror2. Genes Cells 12, 1215–1223 6. Liu, Y., Ross, J. F., Bodine, P. V., and Billiard, J. (2007) Homodimerization of Ror2 tyrosine kinase receptor induces 14–3-3(beta) phosphorylation and promotes osteoblast differentiation and bone formation. Mol. Endocrinol. 21, 3050–3061 7. Masiakowski, P., and Carroll, R. D. (1992) A novel family of cell surface receptors with tyrosine kinase-like domain. J. Biol. Chem. 267, 26181–26190 8. Wallingford, J. B., and Habas, R. (2005) The developmental biology of Dishevelled: an enigmatic protein governing cell fate and cell polarity. Development 132, 4421–4436 9. Semenov, M. V., Habas, R., Macdonald, B. T., and He, X. (2007) SnapShot: noncanonical Wnt signaling pathways. Cell 131, 1378 10. Chen, W., ten Berge, D., Brown, J., Ahn, S., Hu, L. A., Miller, W. E., Caron, M. G., Barak, L. S., Nusse, R., and Lefkowitz, R. J. (2003) Dishevelled 2 recruits beta-arrestin 2 to mediate Wnt5Astimulated endocytosis of Frizzled 4. Science 301, 1391–1394 11. Cong, F., Schweizer, L., and Varmus, H. (2004) Casein kinase Iepsilon modulates the signaling specificities of dishevelled. Mol. Cell. Biol. 24, 2000–2011 12. Kibardin, A., Ossipova, O., and Sokol, S. Y. (2006) Metastasisassociated kinase modulates Wnt signaling to regulate brain patterning and morphogenesis. Development 133, 2845–2854 13. Ossipova, O., Dhawan, S., Sokol, S., and Green, J. B. (2005) Distinct PAR-1 proteins function in different branches of Wnt signaling during vertebrate development. Dev. Cell 8, 829–841 14. Sun, T. Q., Lu, B., Feng, J. J., Reinhard, C., Jan, Y. N., Fantl, W. J., and Williams, L. T. (2001) PAR-1 is a Dishevelledassociated kinase and a positive regulator of Wnt signalling. Nat. Cell Biol. 3, 628–636 15. Willert, K., Brink, M., Wodarz, A., Varmus, H., and Nusse, R. (1997) Casein kinase 2 associates with and phosphorylates dishevelled. EMBO J. 16, 3089–3096 16. Rothbacher, U., Laurent, M. N., Deardorff, M. A., Klein, P. S., Cho, K. W., and Fraser, S. E. (2000) Dishevelled phosphorylation, subcellular localization and multimerization regulate its role in early embryogenesis. EMBO J. 19, 1010–1022 17. Peters, J. M., McKay, R. M., McKay, J. P., and Graff, J. M. (1999) Casein kinase I transduces Wnt signals. Nature 401, 345–350 18. Sammar, M., Stricker, S., Schwabe, G. C., Sieber, C., Hartung, A., Hanke, M., Oishi, I., Pohl, J., Minami, Y., Sebald, W., Mundlos, S., and Knaus, P. (2004) Modulation of GDF5/BRI-b signalling through interaction with the tyrosine kinase receptor Ror2. Genes Cells 9, 1227–1238 19. Kani, S., Oishi, I., Yamamoto, H., Yoda, A., Suzuki, H., Nomachi, A., Iozumi, K., Nishita, M., Kikuchi, A., Takumi, T., and Minami, Y. (2004) The receptor tyrosine kinase Ror2 associates with and is activated by casein kinase Iepsilon. J. Biol. Chem. 279, 50102–50109 20. Schwarzer, W., Witte, F., Rajab, A., Mundlos, S., and Stricker, S. (2009) A gradient of ROR2 protein stability and membrane localization confers brachydactyly type B or Robinow syndrome phenotypes. Hum. Mol. Genet. 18, 4013–4021 21. Bryja, V., Gradl, D., Schambony, A., Arenas, E., and Schulte, G. (2007) Beta-arrestin is a necessary component of Wnt/betacatenin signaling in vitro and in vivo. Proc. Natl. Acad. Sci. U. S. A. 104, 6690–6695 22. Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007) Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J. Cell Sci. 120, 586–595 23. Nieuwkoop, P. D., and Faber, J. (1975) Normal Table of Xenopus laevis (Daudin), Elsevier, Amsterdam 24. Hollemann, T., Panitz, F., and Pieler, T. (1999) In situ hybridization techniques with Xenopus embryos. In A Comparative Methods Approach to the Study of Oocytes and Embryos (Richter, J. D., ed) pp. 279–290, Oxford University Press, Oxford 25. Billiard, J., Way, D. S., Seestaller-Wehr, L. M., Moran, R. A., Mangine, A., and Bodine, P. V. (2005) The orphan receptor tyrosine kinase Ror2 modulates canonical Wnt signaling in osteoblastic cells. Mol. Endocrinol. 19, 90–101 26. Hikasa, H., Shibata, M., Hiratani, I., and Taira, M. (2002) The Xenopus receptor tyrosine kinase Xror2 modulates morphogenetic movements of the axial mesoderm and neuroectoderm via Wnt signaling. Development 129, 5227–5239 27. Oishi, I., Suzuki, H., Onishi, N., Takada, R., Kani, S., Ohkawara, B., Koshida, I., Suzuki, K., Yamada, G., Schwabe, G. C., Mundlos, S., Shibuya, H., Takada, S., and Minami, Y. (2003) The receptor tyrosine kinase Ror2 is involved in non-canonical Wnt5a/JNK signalling pathway. Genes Cells 8, 645–654 28. Schwabe, G. C., Tinschert, S., Buschow, C., Meinecke, P., Wolff, G., Gillessen-Kaesbach, G., Oldridge, M., Wilkie, A. O., Komec, R., and Mundlos, S. (2000) Distinct mutations in the receptor tyrosine kinase gene ROR2 cause brachydactyly type B. Am. J. Hum. Genet. 67, 822–831 29. Angers, S., Thorpe, C. J., Biechele, T. L., Goldenberg, S. J., Zheng, N., MacCoss, M. J., and Moon, R. T. (2006) The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wntbeta-catenin pathway by targeting Dishevelled for degradation. Nat. Cell Biol. 8, 348–357 30. Kishida, M., Hino, S., Michiue, T., Yamamoto, H., Kishida, S., Fukui, A., Asashima, M., and Kikuchi, A. (2001) Synergistic activation of the Wnt signaling pathway by Dvl and casein kinase Iepsilon. J. Biol. Chem. 276, 33147–33155 31. Swiatek, W., Tsai, I. C., Klimowski, L., Pepler, A., Barnette, J., Yost, H. J., and Virshup, D. M. (2004) Regulation of casein kinase I epsilon activity by Wnt signaling. J. Biol. Chem. 279, 13011–13017 32. Schulte, G., Bryja, V., Rawal, N., Castelo-Branco, G., Sousa, K. M., and Arenas, E. (2005) Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation. J. Neurochem. 92, 1550–1553 33. Gonzalez-Sancho, J. M., Brennan, K. R., Castelo-Soccio, L. A., and Brown, A. M. (2004) Wnt proteins induce dishevelled phosphorylation via an LRP5/6- independent mechanism, irrespective of their ability to stabilize beta-catenin. Mol. Cell. Biol. 24, 4757–4768 34. Wallingford, J. B., Fraser, S. E., and Harland, R. M. (2002) Convergent extension: the molecular control of polarized cell movement during embryonic development. Dev. Cell 2, 695–706 35. Bilic, J., Huang, Y. L., Davidson, G., Zimmermann, T., Cruciat, C. M., Bienz, M., and Niehrs, C. (2007) Wnt induces LRP6 signalosomes and promotes dishevelled-dependent LRP6 phosphorylation. Science 316, 1619–1622 36. Schwarz-Romond, T., Fiedler, M., Shibata, N., Butler, P. J., Kikuchi, A., Higuchi, Y., and Bienz, M. (2007) The DIX domain of Dishevelled confers Wnt signaling by dynamic polymerization. Nat. Struct. Mol. Biol. 14, 484–492 37. Green, J. L., Inoue, T., and Sternberg, P. W. (2007) The C. elegans ROR receptor tyrosine kinase, CAM-1, non-autonomously inhibits the Wnt pathway. Development 134, 4053–4062 38. Winkel, A., Stricker, S., Tylzanowski, P., Seiffart, V., Mundlos, S., Gross, G., and Hoffmann, A. (2008) Wnt-ligand-dependent interaction of TAK1 (TGF-beta-activated kinase-1) with the receptor tyrosine kinase Ror2 modulates canonical Wnt-signalling. Cell. Signal. 20, 2134–2144 39. Bryja, V., Schambony, A., Cajanek, L., Dominguez, I., Arenas, E., and Schulte, G. (2008) Beta-arrestin and casein kinase 1/2 define distinct branches of non-canonical WNT signalling pathways. EMBO Rep. 9, 1244–1250 Received for publication November 18, 2009. Accepted for publication February 4, 2010. 2426 Vol. 24 July 2010 WITTE ET AL.The FASEB Journal ⅐ www.fasebj.org Supplemental Data Supplemental Figure 1. Western blots for schematic depictions of Ror2 / Dvl interaction (Fig. 1A, B). Cells were co-transfected with indicated constructs and pulldown was performed against Ror2-Flag (A) or Dvl3-Flag (B). (C) Interaction of Dvl3 with Ror2. Cells were co-transfected with indicated constructs and pulldown was performed against Dvl3-Flag. Dvl3 interacts with full length Ror2 but not with Ror2∆745 lacking the PST-rich domain. Supplemental Figure 2. Wnt3a-induced Dvl2 / Ror2 colocalization at the cell membrane is dependent on the C-terminus of Ror2 and CK1 activity. Upper panel: cells have been transfected with the BDB1 mutant form Ror2W749X and stimulated with Wnt3a for 30 min. Lower panel: cells have been transfected with wt Ror2, treated with CK1 inhibitor D4476 and stimulated with Wnt3a for 30 min. In both cases, no co-localization of Ror2 and endogenous Dvl2 at the cell membrane is observed. Supplemental Figure 3. Recruitment of Dvl to the cell membrane by Ror2 is not dependent on the ligand-binding cysteine-rich (CRD) domain of Ror2. A mutant form of human Ror2 that lacks the Wnt-binding CRD (Ror2∆CRD) shows colocalization with endogenous Dvl2 at the cell membrane after stimulus with Wnt3a (upper panel) or Wnt5a (lower panel). Supplemental Figure 4. Inhibition of TOPFLASH activity induced by Dvl2 + CK1e by wt and mutant forms of Ror2. Error bars depict standard errors. Statistical significance determined by Student’s t-test. One asterisk: p<0,05; two asterisks: p<0,01; three asterisks: p<0,001. Supplemental Figure 5. Western blot analysis of cell transfections used for TOPFLASH reporter assays. (A) Cells were transfected with plasmids encoding Dvl2-Myc, CK1ε and Ror2-FLAG mutants as indicated in the identical amounts as in the Fig. S4, and the expression levels were analyzed by Western blotting against Flag and CK1ε. The blot demonstrates equal expression of individual Ror2 mutants. (B) Cells were transfected with plasmids encoding Dvl3-Flag, Dvl3∆C-Flag and CK1ε in identical amounts as in the Fig. 5A. The kinase-dead variant CK1ε-KD was used as control. The blot demonstrates equal expression of Dvl3-Flag and Dvl3∆C-Flag, and the kinase activity of CK1ε towards both constructs, visible as phosphorylation-dependent mobility shift. Both pictures originate from the same blot membrane. (C) Cells were transfected with plasmids encoding Dvl3-Flag, CK1ε and Dvl3-C-terminus (Dvl3-CT) in identical amounts as in the Fig. 5B, and analyzed by Western blotting. The blot demonstrates activity of CK1ε visible as Dvl shift, increasing doses of Dvl3-CT and no effect of Dvl3-CT on the levels of Dvl3-Flag. (D) Cells were transfected with the plasmids encoding HA-S33A-beta-catenin and increasing doses of FlagDvl3-CT, in the amounts identical to Fig. 5E and analyzed by Western blotting. Supplemental Figure 6. Interaction of mouse Ror2 (mRor2), but not the truncated mRor2∆745 lacking the PST domain, with Xenopus Dvl (XDvl). Supplemental Figure 7. Validation of Ror2 siRNAs. (A) Western blot from HEK293 cells transfected with control or 3 human-Ror2-specific siRNAs probed for endogenous Ror2. (B) TOPFLASH assays stimulated with Dvl3 + CK1ε. All three hRor2-siRNAs enhanced reporter activation. Supplemental Table 1. Densitrometric quantification of western blot band intensities from Figures 1 C, D, E. and Supplemental Figure 1B. See arrows for pulldown differences mentioned in the manuscript text. Vítězslav Bryja, 2014    Attachments      #10      Foldynova‐Trantirkova  S1 ,  Sekyrova  P1 ,  Tmejova  K,  Brumovska  E,  Bernatik  O,  Blankenfeldt  W,  Krejci  P,  Kozubik  A,  Dolezal  T,  Trantirek  L*,  Bryja  V*  (2010):  Breast  cancer‐specific  mutations  in  CK1epsilon  inhibit  Wnt/beta‐catenin  and  activate  the  Wnt/Rac1/JNK  and  NFAT  pathways  to  decrease  cell  adhesion  and  promote  cell  migration. Breast Cancer Res. 12(3):R30.  1  equal contribution, * corresponding authors      Impact factor (2010): 5.785  Times cited (without autocitations, WoS, Feb 21st 2014): 13  Significance: This work demonstrates that CK1 inhibition leading to the activation of  Rac1 (desribed by us in the work #6) has clinical consequences. Specifically, we  show that CK1 mutations identified in breast cancer inactivate CK1, are dominant  negative, activate non‐canonical Wnt/Dvl/Rac1 pathway and inhibit canonical Wnt  signaling. As such they promote Rac1‐driven migration of breast cancer cells.  Contibution  of the author/author´s team:  Experimental design, functional testing of  CK1epsilon mutants in biochemical and biological experiments.                Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Open AccessRESEARCH ARTICLE © 2010 Foldynová-Trantírková et al.; licensee BioMed Central Ltd. This is an open access article distributed under the terms of the Creative Commons Attribution License http://creativecommons.org/licenses/by/2.0, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Research article Breast cancer-specific mutations in CK1ε inhibit Wnt/β-catenin and activate the Wnt/Rac1/JNK and NFAT pathways to decrease cell adhesion and promote cell migration Silvie Foldynová-Trantírková†1, Petra Sekyrová†1, Kateřina Tmejová2,3, Eva Brumovská1, Ondřej Bernatík2, Wulf Blankenfeldt4, Pavel Krejčí2,3, Alois Kozubík2,3, Tomáš Doležal1, Lukáš Trantírek*1,5 and Vítězslav Bryja*2,3 Abstract Introduction: Breast cancer is one of the most common types of cancer in women. One of the genes that were found mutated in breast cancer is casein kinase 1 epsilon (CK1ε). Because CK1ε is a crucial regulator of the Wnt signaling cascades, we determined how these CK1ε mutations interfere with the Wnt pathway and affect the behavior of epithelial breast cancer cell lines. Methods: We performed in silico modeling of various mutations and analyzed the kinase activity of the CK1ε mutants both in vitro and in vivo. Furthermore, we used reporter and small GTPase assays to identify how mutation of CK1ε affects different branches of the Wnt signaling pathway. Based on these results, we employed cell adhesion and cell migration assays in MCF7 cells to demonstrate a crucial role for CK1ε in these processes. Results: In silico modeling and in vivo data showed that autophosphorylation at Thr 44, a site adjacent to the breast cancer point mutations in the N-terminal lobe of human CK1ε, is involved in positive regulation of the CK1ε activity. Our data further demonstrate that, in mammalian cells, mutated forms of CK1ε failed to affect the intracellular localization and phosphorylation of Dvl2; we were able to demonstrate that CK1ε mutants were unable to enhance Dvl-induced TCF/LEF-mediated transcription, that CK1ε mutants acted as loss-of-function in the Wnt/β-catenin pathway, and that CK1ε mutants activated the noncanonical Wnt/Rac-1 and NFAT pathways, similar to pharmacological inhibitors of CK1. In line with these findings, inhibition of CK1 promoted cell migration as well as decreased cell adhesion and E-cadherin expression in the breast cancer-derived cell line MCF7. Conclusions: In summary, these data suggest that the mutations of CK1ε found in breast cancer can suppress Wnt/βcatenin as well as promote the Wnt/Rac-1/JNK and Wnt/NFAT pathways, thus contributing to breast cancer development via effects on cell adhesion and migration. In terms of molecular mechanism, our data indicate that the breast cancer point mutations in the N-terminal lobe of CK1ε, which are correlated with decreased phosphorylation activities of mutated forms of CK1ε both in vitro and in vivo, interfere with positive autophosphorylation at Thr 44. Introduction Mammary carcinomas are one of the most common neoplasias in women. Several improvements in understanding the molecular pathology of breast cancer have been achieved in the past decade. In most cases, however, the molecular mechanisms underlying this malignancy are still unknown. Sequencing of mammary carcinoma samples by Fuja and colleagues revealed that the casein kinase 1 epsilon (CK1ε) gene was mutated in this disease; CK1ε was found to be mutated within its N-terminal region with approximately 15% incidence [1]. CK1ε is a Ser/Thr kinase with * Correspondence: L.Trantirek@uu.nl, bryja@sci.muni.cz 1 Biology Centre AS CR, v.v.i. AND University of South Bohemia, Branisovska 31, 37005 Ceske Budejovice, Czech Republic 2 Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlarska 2, CZ-61137 Brno, Czech Republic † Contributed equally Full list of author information is available at the end of the article Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 2 of 14 many known functions and substrates. CK1ε phosphorylates several regulators of crucial processes, such as cell proliferation, differentiation, migration, and circadian rhythms. The key known targets of CK1ε involve p53, key components of the circadian clock, the Wnt signaling pathway, and cell division machinery (for a review, see [2]). In the original sequencing study, 19 nonsynonymous mutations were identified in the CK1ε gene in ductal carcinoma samples [1]. The identified mutations were shown to have a significant association with the loss of heterozygosity and decreased staining of CK1ε in the tumor sections. Some of the mutations in CK1ε were found repeatedly in several patients, such as L39Q (detected five times), L49Q (detected three times), and S101R (detected twice) [1]. These observations suggest that these mutations may affect CK1ε function and may be favored during the microevolution of the tumor, and thus may contribute to tumor progression. Importantly, nothing is known about how these mutations affect the kinase activity and signaling potential of CK1ε and the behavior of mammary cells. In the present study, we characterized three CK1ε mutants that were previously identified in mammary carcinoma. We demonstrated that these CK1ε mutants had limited kinase activity and failed to phosphorylate the physiological targets of CK1ε in vitro and in vivo. The analyzed mutations acted as loss of function in the Wnt/β-catenin pathway and promoted the alternative Wnt/Rac1 pathway, which in turn decreased cell adhesion and promoted cell migra- tion. Materials and methods Plasmids ORFs of the wild-type (WT), full-length human CKIε cDNA (residues 1 to 416), two mutants mimicking either nonphosphorylatable Thr 44 (Thr44Ala) or constitutively phosphorylated Thr 44 (Thr44Asp), and three mutated versions (P3, P4, and P6) were cloned into pcDNA3. The truncated versions of CKIεΔC (residues 1 to 315) were cloned into pHAK-B3. Plasmids encoding mDvl2-Myc [3] and human Dvl3-Flag [4] have been previously described. Details and bacterial overexpression vectors are presented in Additional file 1. Structural modeling The three-dimensional model for CK1ε was obtained via template-based homology modeling using the program PHYRE [5]. The mutated sites and kinase-specific functional domains were mapped onto a three-dimensional model of CK1ε using the program CHIMERA [6]. The kinase-specific functional domains in CK1ε were predicted using the NCBI Conserved Domain Database [7]. Predictions of changes in protein stability upon point mutations were conducted using CUPSTAT [8]. (Auto)Phosphorylation sites were predicted using GPS v. 2.1 [9]. Western blot analysis, immunoprecipitation, and small GTPase activity assays Western blot analysis, immunoprecipitation, and small GTPase activity assays were performed as previously described [10,11]. The antibodies used for the western blot analysis were as follows: mouse anti-Flag (M2; Sigma, Schnelldorf, Germany), goat anti-CK1e (sc6471; Santa Cruz Biotechnology, Heidelberg, Germany), mouse anti-Myc (sc-40) and anti-actin (sc-1615-R) (both Santa Cruz Biotechnology), anti-HA (HA.11; Nordic Biosite, Täby, Sweden), mouse anti-Rac (05-389; Upstate, Waltham, Massachusetts, USA), and mouse anti-Cdc42 (610928, 1:1,000; BD Biosciences, San Jose, California, USA). The antibodies used for immunoprecipitation were as follows: anti-CK1ε (sc-6471; Santa Cruz Biotechnology), anti-MYC (C3965; Sigma), and anti-FLAG (F1804; Sigma). Immunohistochemistry Transfected cells were grown on glass coverslips, washed with PBS, fixed for 15 minutes in 4% paraformaldehyde, washed with PBS, and blocked in 1.5% BSA, 0.1% Triton X-100 in PBS for 1 hour. After overnight incubation at 4°C with the primary antibody, cells were washed with PBS, 0.1% Triton X-100, incubated for 2 hours at room temperature with secondary antibody, washed with PBS, 0.1% Triton X-100, and counterstained with 4 μg/ml 4'',6 Diamidino - 2 - phenylindole, dihydrochloride. Phalloidin-Alexa488 (1:66; Molecular Probes, Carlsbad, CA, USA) was added in the last 30 minutes of incubation with the secondary antibody. Samples were analyzed with a FV1000 confocal microscope (Olympus). The following antibodies were used: mouse anti-Myc (1:150) and goat anti-CK1ε (1:100) (both from Santa Cruz Biotechnology), anti-goat-Cy5 (1:500; Molecular Probes), and antimouse-Alexa488 (1:1,000; Molecular Probes). Real-time impedance measurements AceaE-plates®96 were used for non-invasive real-time measurements with an xCELLigence RTCA SP system and RTCA software version 1.1 (both Roche Applied Science, Indianapolis, IN, USA). First, a background measurement was performed using 100 μl complete cultivation media with or without CK1 inhibitors incubated for 30 minutes in the incubator (37°C, 5% CO2). MCF7 cells were trypsinized, quantified, and seeded (15,000 cells/well) in an additional 100 μl cultivation media. The impedance was then monitored continually for a period of 20 hours. Data are presented as a cell index. In parallel to xCELLigence measurements, cells were seeded (15,000 cells/well) in 24-well plates, and the Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 3 of 14 effects of IC261 on cell number and cell viability were determined after 20 hours. The cell number was measured using a Coulter Counter (Immunotech a. s., A Beckman Coulter Company, Praque, Czech Republic). Cell viability was determined using eosin staining. Hanging drop assay MCF7 cells (4 × 105 cells/ml) with or without CK1ε inhibitors were seeded in 30 μl drops on the inner side of a 10 cm plate lid. The lid of the plate was then turned upside down and placed on top of the plate filled with 10 ml PBS. MCF7 cells under the force of gravity aggregated at the bottom of the hanging drop. After 24 hours, cell clusters were photographed, collected, and resuspended by pipetting up and down seven times, and single cells released from the clusters were counted. Luciferase assays HEK293 and MCF7 cells were transfected in a 24-well plate format with the indicated combinations of plasmids using polyethylene-imine (HEK293) or FuGENE (MCF7). The amounts used per well were as follows: vectors encoding Dvl2, Dvl3, and CK1ε, 300 ng; pRLtkLuc (Renilla Promega, Madison, WI, USA), 50 ng; and firefly luciferase reporters, 200 ng. The following luciferase reporters were used: pSuperTopFlash [12], p-E-cadherinLuc [13], pNFAT-Luc, and pAP1-Luc (Stratagene, La Jolla, CA, USA). The total amount of plasmid DNA per well was kept constant in all conditions. Twenty-four hours post transfection, cells were harvested and processed using the Dual Luciferase kit (Promega, Madison, WI, USA) according to the manufacturer's protocol. To normalize for the efficiency of transfection, firefly luciferase values were normalized to Renilla luciferase in each well. Each data point was run in duplicate, and at least three independent experiments were performed. Datasets from each experiment were normalized to the control, and all individually normalized experiments were statistically analyzed by analysis of variance and Tukey's multiple comparison post tests (GraphPad Prism software, La Jolla, CA, USA The graphs indicate the mean ± standard deviations from at least three independent replicates. Transwell assay MCF7 cells (5 × 104 cells) were seeded into the upper chamber of Transwell plates (Transwell 96-well plate, 5.0 μm polycarbonate membrane; Corning (Corning, MY, USA)). The inhibitors (D4476 or IC261; Calbiochem, Darmstadt, Germany) were added into the bottom chamber at the indicated concentrations. Identical amounts of dimethylsulfoxide were used as a control. After 24 hours, cells that migrated through the filter into the bottom chamber were counted with a Coulter Counter. The number of migrating cells was normalized to control conditions and expressed as a migration index. Results CK1ε mutants can bind but fail to phosphorylate Dishevelled L39Q is the most frequently occurring mutation found in patients with breast cancer. In these patients, this mutation occurs alone (Patient P6) or in combination with other mutations - L39Q and S101R (Patient P3), or L39Q, L49Q, and N78T (Patient P4) (Figure 1a). Leu 39 and Leu 49, as well as Ser 101 (occupied exclusively by Ser or Thr), are absolutely conserved among CK1 paralogs and CK1ε orthologs (data not shown). One of the best defined functions of CK1ε is the Wntinduced phosphorylation of Dvl, a crucial component of the Wnt signaling cascades [11,14-20]. In the first step, we analyzed the capacity of individual CK1ε mutants to phosphorylate Dvl in mammalian HEK293 cells. As shown in Figure 1b, WT CK1ε promoted the formation of phosphorylated and shifted (PS) Dvl2, whereas mutant forms did not. A partial shift was observed with the P6 mutant, but the P3/P4 mutants were indistinguishable when compared with the control-transfected sample. Deletion of the C-terminus of CK1e, which inhibits CK1e kinase activity when phosphorylated [20], did not affect these results (Figure 1c). Importantly, using immunoprecipitation we were able to detect the presence of overexpressed Dvl2 or Dvl3, and both WT and mutated CK1e kinases in one protein complex (Figure 2a and Additional file 2), suggesting that the CK1e mutants still possess the ability to bind Dvl. Phosphorylation of Dvl by CK1ε leads not only to the formation of PS-Dvl but also to changes in the intracellular distribution of Dvl [11,19]. Dvl is usually present in dynamic multiprotein aggregates [21,22] called Dvl dots. Based on the cellular context and activity of CK1e, which promotes the dissolution of Dvl aggregates, Dvl is usually present either in dots or in an even distribution (see Figure 2b for COS7 cells; see Additional file 2 for HEK293 cells). WT CK1ε strongly promoted an even localization of Dvl2-Myc (Figure 2c, d and Additional file 2). In contrast, all of the analyzed mutants (P3, P4, and P6) promoted the formation/maintenance of Dvl dots in COS7 cells and co-localized with Dvl in these multiprotein complexes (Figure 2c, d), similar to earlier observations with dominant negative CK1ε or CK1 inhibitors [11,19]. All CK1ε proteins were evenly distributed in the absence of Dvl (Figure 2c, e). These data together show that despite the fact that CK1ε mutants bind and co-localize with Dvl, they cannot phosphorylate Dvl or efficiently promote its even localization. Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 4 of 14 In silico modeling of ductal carcinoma-specific CK1ε mutations To understand how individual mutations are spatially related to functionally conserved regions in CK1ε, we developed three-dimensional models for individual CK1ε mutants. In the single-point mutant (P6), the mutated site (Leu 39) directly adjoins conserved residues that participate in ATP binding (Figure 3a). The CK1ε mutant P3 contains two mutations, L39Q and S101R. These mutated sites are distant from each other not only in sequence but also spatially (Figure 3b). While Leu 39 is located in the βstrand-rich N-terminal lobe, Ser 101 is located at the Cterminal end of helix I in the primarily helical C-terminal subdomain. The S101R mutation is predicted to destabilize and alter local protein structure, thus affecting the structural integrity of the four-helix bundle, which forms the structurally conserved core of the C-terminal subdomain [23]. The last analyzed mutant, P4, contains three point mutations: L39Q, L49Q, and N78T. Although the individually mutated sites are distant in primary sequence, when mapped onto the three-dimensional structure of CK1ε they cluster into a very small area with a radius <10 Å that is located between strands four and five as well as into αhelix B of the N-terminal lobe (Figure 3c). Interestingly, all of these mutations surround a conserved predicted autophosphorylation site at Thr 44 in the N-terminal catalytic domain. This site is localized within a loop region between the terminus of the fourth β-strand (only five residues upstream from the mutation site L39Q) and αhelix B (five residues downstream from the second mutation site L49Q) (Figure 3d). To assess kinase activities of the individual CK1ε mutants, we attempted to heterologously express individual mutants of human CK1ε as affinity tagged recombinant proteins lacking the autoinhibitory domain in Escherichia coli. Despite significant efforts, we were unable to obtain soluble overexpression for all of the constructs when we expressed them with the same affinity tag. Soluble expressions were feasible for WT, P3, P4, and P6, however, when the recombinant proteins were tagged to His6, to maltose binding protein, to maltose binding protein, and to Small Ubiquitin-like Modifier (SUMO), respectively. The individual recombinant fusion proteins were assayed for their ability to phosphorylate Dvl in vitro. Although the data from this in vitro phosphorylation assay were in qualitative agreement with our in vivo data (see above), we considered these in vitro data inconclusive since the presence of multiple different tags has made direct interpretation impossible (see Additional files 1 and 3). CK1ε mutants act as loss-of-function in the Wnt/β-catenin pathway To test the role of mutated CK1ε proteins in the canonical Wnt pathway, we induced Wnt/β-catenin signaling via the overexpression of several of the components of this pathway - such as Dvl2, Dvl3, β-catenin, Wnt3a, and the Lrp6 co-receptor - and analyzed TCF/LEF-driven transcription using the TopFlash reporter system [24]. As shown in Figure 4, Dvl2-Myc weakly activated the TopFlash reporter in HEK293 and MCF7 cells, but the signal Figure 1 Casein kinase 1 epsilon mutant phosphorylation of Dishevelled. All three mutants of casein kinase 1 epsilon (CK1ε) bind Dvl but are unable to cause the typical phosphorylation-dependent shift of the Dvl protein. (a) Schematic representation of the CK1ε mutants used in this study. (b) HEK293 cells were transfected with a plasmid encoding Dvl2-Myc and the CK1ε variants. CK1ε phosphorylates Dvl proteins, which subsequently exhibit a mobility shift in western blot analysis. The P6 mutant maintains residual activity, while the P3 and P4 mutants are unable to promote a Dvl2-Myc shift. (c) Truncated versions of CK1ε that lack the C-terminal autoinhibitory domain do not differ from full-length proteins in their ability to shift the Dvl2 protein. B. WT P4P3 P6 CK1 C CTRL C. Dvl2Myc HA-CK1 actin WT P4P3 P6 CTRL CK1 CK1 Dvl2Myc actin Patient/mutation # Changes predicted by mutations P3 L39Q,S101R P4 L39Q, L49Q, N78T P6 L39Q A. Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 5 of 14 Figure 2 Interaction of casein kinase 1epsilonmutants andDishevelled2. Casein kinase 1 epsilon (CK1ε) mutants associate with Dvl2, co-localize with Dvl2, and promote the punctate cytoplasmic localization of Dvl2. (a) HEK293 cell lysates transfected with either wild-type (WT) CK1ε or the P3, P4, and P6 mutants together with Dvl2-Myc were lysed and immunoprecipitated using an anti-Myc antibody. WT CK1ε and all of the CK1ε mutants efficiently bind the Dvl2 protein. (b) Dvl2-Myc protein localization in transfected COS7 cells was observed by confocal microscopy using an anti-Myc antibody. Dvl2 is either found in cytoplasmic inclusions or evenly dispersed within the cytoplasm. (c) WT CK1ε co-transfected with Dvl2-Myc into COS7 cells dissolves most of the Dvl2 punctae, resulting in the predominance of evenly distributed Dvl2 protein. In contrast, the P3, P4, and P6 mutants enhance the formation of Dvl2 clusters with a punctate appearance in the cytoplasm. A localization pattern of CK1ε was observed with an anti-CK1e antibody. All CK1ε forms show even cytoplasmic distribution when transfected alone. (d), (e) Cells with an even or punctate distribution of Dvl2 and CK1e were counted. For each condition, at least 150 cells were examined. 2 Dvl2Myc+WT Dvl2-Myc CK1¡ merge Dvl2-Myc A. C. D. D vl2-M ycD vl2-M yc /W TD vl2-M yc /P3D vl2-M yc /P4D vl2-M yc /P6 0.0 0.2 0.4 0.6 0.8 1.0 even punctate Dvl2-Myclocalization W T P3 P4 P6 0.0 0.2 0.4 0.6 0.8 1.0 even punctate CK1!localization CK1¡ P4 P6 P3 WT Dvl2Myc+P3Dvl2Myc+P4Dvl2Myc+P6 E. Dvl2Myc B. WT P4P3 P6 Dvl2-Myc CK1¡ CK1¡ Dvl2-Myc TCL IP: Myc WTCK1¡ + ++ +-+ Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 6 of 14 was boosted when WT CK1e was co-expressed. The CK1e mutants P3 and P4 failed to synergize with Dvl2, and P6 promoted Dvl2-driven TopFlash only moderately (Figure 4a, b). Very similar results have been obtained with Dvl3-Flag (Additional file 4). Importantly, the effects of CK1e were Dvl dependent because the overexpression of any form of CK1e had only negligible effects (Additional file 4). The inhibitory effects of the CK1e mutants were found at the level of Dvl because the activation of TopFlash by constitutively active (S33A)-β-catenin could not be significantly modulated by the overexpression of any CK1e (Figure 4c). Cocultivation with fibroblasts producing Wnt3a or overexpression of crucial Wnt co-receptor Lrp6 efficiently induced TCF/LEF-dependent transcription. Co-expression of WT CK1e promoted TopFlash even further, whereas the expression of the P3, P4, and P6 CK1e mutants were able to reduce the Wnt3a/Lrp6-induced signal. This effect was obvious in Wnt-3a-stimulated cells and became statistically significant in cells overexpressing Lrp6 (Figure 4d, e). These analyses demonstrate that the P3, P4, and P6 mutants of CK1e are dysfunctional in the Wnt/β-catenin pathway and act upstream of β- catenin. Our structural modeling (Figure 3) suggested that CK1e mutations found in breast cancer may affect the predicted autophosphorylation site at Thr 44. To test the function of Thr 44 we generated two mutants mimicking Figure 3 Homology model of the catalytic domain of casein kinase 1 epsilon mutants. Three-dimensional representation of a homology model of the catalytic domain of casein kinase 1 epsilon (CK1ε) (residues 1 to 295). Phylogenetically conserved functional sites across the serine/threonine kinase superfamily are indicated: pink, substrate binding pocket; yellow, ATP binding site; red, catalytic loop; blue, activation loop. The individual point mutations that were found in patients (a) P6 (L39Q), (b) P3 (L39Q, S101R), and (c) P4 (L39Q, L49Q, N78T) are indicated in orange. (d) A mutation cluster region in P4. The mutations and ATP binding region are indicated in orange and yellow, respectively. A blue sphere indicates a predicted autophosphorylation site (Thr 44). P4 P3 Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 7 of 14 Figure 4 Casein kinase 1 epsilon mutants downregulate Wnt/β-catenin signaling at the level of the Dishevelled protein. HEK293 or MCF7 cells were transfected with the indicated plasmids, and after 24 hours TCF/LEF-driven transcription was measured using the SuperTopFlash dual luciferase reporter assay. (a) Dvl2 enhances TCF/LEF-dependent transcription when co-transfected with wild-type (WT) casein kinase 1 epsilon (CK1ε) in HEK293 cells. In contrast, all CK1ε mutants are unable to promote Dvl-mediated TopFlash. (b) Similar results to those obtained for HEK293 cells have been obtained for the breast cancer cell line MCF7. (c) β-Catenin induces the TCF/LEF transcriptional response in the cell by recruiting transcriptional cofactors directly to the TCF/LEF complex. Neither WT CK1ε nor the CK1ε mutants could influence the high TopFlash signal caused by constitutively active β-catenin, confirming that CK1ε and its mutants operate upstream of β-catenin (HEK293 cells). (d) The Wnt signaling pathway was induced by seeding B1A-wnt3a cells (fibroblasts that constitutively produce the Wnt3a protein) over transfected HEK293 cells. Wnt3a induces the Wnt/β-catenin/ TCF/LEF pathway in control cells transfected with empty plasmid. The Wnt3-induced signal is enhanced by overexpression of WT CK1ε but is strongly depleted by overexpression of the CK1ε mutants. (e) Constitutively activation of the co-receptor LRP6 in transfected HEK293 cells confirmed the regulatory function of CK1ε downstream of the Wnt-receptor complex. WT CK1ε transduces the signal from the activated receptor towards TCF/LEF-driven transcription, while mutant CK1ε proteins abolish downstream signaling. Neither WT nor mutant CK1ε are able to markedly activate the TCF/LEF signal when Wnt stimulation is blocked by the mutant LRP6. (f) Nonphosphorylatable (T44A) and phospho-mimicking (T44D) mutants of CK1ε were co-expressed with Dvl2-Myc in HEK293 cells. The effects on TCF/LEF-dependent transcription show that T44 D behaves as overactive CK1ε. (a) to (f) Data represent the mean ± standard deviation from normalized values. *P < 0.05, ***P < 0.001; n.s., not significant; one-way analysis of variance, Tukey post-test, n ≥ 3. Dvl2-Myc (MCF-7) control D vl2 D vl2+W T D vl2+P3 D vl2+P4 D vl2+P6 0 5 10 TopFlash(foldchange) Dvl2-Myc (HEK293) 0 1 2 3 4 5 6 7 8TopFlash(foldchange) D. B. C. `-catenin control -catenin ` -catenin+ W T-catenin+P3-catenin+P4-catenin+P6 0.0 0.5 1.0 1.5 TopFlash(foldchange) Wnt3a - + - + - + - + - + 0.0 0.5 1.0 1.5 pCDNA3 WT P3 P4 P6 Wnt3a TopFlash(foldchange) E. Lrp6 Lrp6 WT Lrp6 MUT 0.0 0.5 1.0 1.5 2.0 no CK1" WT P6 P3 P4 TopFlash(foldchange) *** A. ***n.s. *** n.s. Dvl2 Myc/CK1-T44 control D vl2 D vl2+W T D vl2+T44A D vl2+T44D 0 1 2 3 ** TopFlash(foldchange) F. control D vl2 D vl2+W T D vl2+P3 D vl2+P4 D vl2+P6 ` ` ` ` Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 8 of 14 either nonphosphorylatable Thr 44 by replacing Thr 44 for alanine (T44A) or constitutively phosphorylated Thr 44 by replacing Thr 44 for aspartic acid (T44D). As shown in Figure 4f, T44D-CK1e is a more potent activator and T44A a less potent activator of TCF/LEF-driven transcription than WT CK1e when co-expressed with Dvl2-Myc. This finding confirms the positive functional role of Thr 44 phosphorylation in the modulation of CK1e activity, and explains how L39Q mutation contributes to the diminishing of CK1e activity in the Wnt/βcatenin pathway. CK1ε mutants activate small GTPase Rac-1 and AP1-driven transcription Our laboratory and others have previously shown that CK1e can act as a switch that promotes Wnt signaling pathways that are dependent on PS-Dvl (such as the Wnt/ β-catenin or PS-Dvl-dependent noncanonical pathways) and blocks the Wnt/Rac1/JNK pathway [10,19]. We hypothesized that the CK1e mutants may mimic CK1e inhibition. To test this prediction, we tested the effects of WT and P3 CK1e on Rac1 activation. As shown in Figure 5a, WT CK1e slightly downregulated Rac1 activity, whereas expression of the P3 mutant or inhibition of CK1 by D4476 promoted this activity. We were not able to detect any effects of CK1e on the activity of another small GTPase, Cdc42 (Additional file 5), and we also failed to detect active RhoA in HEK293 cells (data not shown). Importantly, the CK1e mutants also promoted Dvl2Myc-induced AP1-mediated transcription, which is downstream of the Rac1/JNK pathway (Figure 5b, c). Similar results were obtained with Dvl3-Flag (Additional file 4). We also tested the effects of the CK1e mutants on the noncanonical Wnt/Ca2+ pathway. As shown in Figure 5d, WT CK1e did not affect the transcriptional activity of NFAT, a transcription factor that is activated by calcium waves. The activity of the NFAT reporter was promoted by the CK1e mutants, however, suggesting that mutant forms of CK1e can promote the Wnt/Ca2+ pathway. In summary, our data demonstrate that mutant CK1e, which is present in ductal carcinoma, can act as an inhibitor of the Wnt/β-catenin and an activator of the noncanonical Wnt/Rac1/JNK and Wnt/Ca2+ pathways. CK1ε inhibition decreases cell adhesion and promotes migration in MCF7 cells To date, whether mutations in CK1e that are associated with breast cancer have any effect on the behavior of breast adenocarcinoma cells is not clear. Our results show that mutated forms of CK1e behaved identically to CK1 inhibition in all tested assays - compare Figure 2 and [11], and compare Figure 5 and [10]. Based on these analyses, one could expect that the presence of CK1e mutants Figure 5 Casein kinase 1 epsilon mutants act as loss of function in the Wnt/β-catenin pathway. Casein kinase 1 epsilon (CK1ε) mutants activate small GTPase Rac1 and the transcriptional activity of AP1 and NFAT. (a) Lysates from HEK293 cells transfected with Rac1-Myc were subjected to pull-down of the active GTP-Rac1 form with agarose-PAK beads. HEK293 cells were either co-transfected with wild-type (WT) or mutant CK1ε or treated with 100 μM D4476 inhibitor for 4 hours prior to lysis. The P3 mutant promotes Rac1 activation, as a higher amount of GTP-Rac1 was pulled from the lysate (as compared with WT CK1ε). Consistently, D4476 inhibits the function of WT CK1ε and elevates the level of GTP-Rac1 in cells. Rac1 protein was detected with an anti-Myc antibody. Western blots were quantified by densitometry, and the GTP-Rac1/Rac1 ratios are indicated by the numbers below the panel. After probing, membranes were stained with amidoblack to confirm equal protein loading. (b) HEK293 cells or (c) MCF7 cells were transfected with an AP1-luciferase reporter, Dvl2-Myc, and CK1ε as indicated. Cells were lysed, and luciferase activity was analyzed the next day. (d) HEK293 cells were transfected with a NFAT-luciferase reporter and CK1ε as indicated. Cells were lysed, and luciferase activity was analyzed the next day. (b) to (d) Data represent the mean ± standard deviation. *P < 0.05, **P < 0.01; one-way analysis of variance, Tukey post-test, n ≥ 3. A. B. D. AP1-Luc (HEK 293) pcD N A D vl2 D vl2+W T D vl2+P3 D vl2+P4 D vl2+P6 0 1 2 3 Foldchange * ** *** NFAT reporter C ontrol W T P3 P4 P6 0 1 2 3 4 5 Foldchange * * * P3 WT- D4476 total Rac1 GTP-Rac1 WTCK1¡ - ++-- Rac1-Myc + ++++ - - - 1.0 0.9 2.6 1.7 2.5 GST-PAK total protein C. * ** AP1-Luc (MCF-7) pC D N A 3 D vl2 D vl2+W T D vl2+P3 D vl2+P4 D vl2+P6 0.0 0.5 1.0 1.5 2.0 Foldchange Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 9 of 14 is functionally equivalent to the pharmacological inhibition of CK1e. Taking advantage of this finding, we tested the effects of the CK1 inhibitors D4476 and IC261 on the cell adhesion of MCF7 epithelial breast cancer cells. The MCF7 cell line retains several characteristics of differentiated mammary epithelium, including the ability to process estradiol via cytoplasmic estrogen receptors and the capability of forming domes [25], and is suitable for analyzing cellular changes during tumor progression. D4476 is the most specific inhibitor of CK1 known to date [26,27], whereas IC261 specifically inhibits the CK1δ and CK1ε isoforms at doses <10 μM [27,28]. MCF7 cells efficiently aggregated in hanging drops and usually formed one cell cluster per drop (Figure 6a) after overnight aggregation. Inhibition of CK1 interfered with aggregate formation and caused the MCF7 cells to become more loosely attached to each other (Figure 6a, c). To quantify the level of adhesion, we mechanically resuspended the cell clusters and counted the number of single cells, which was increased by CK1 inhibition (Figure 6b, d). Similar results were obtained when we downregulated the levels of CK1δ and CK1ε in MCF7 cells via siRNAmediated knockdown (Additional file 6). Next, we seeded the MCF7 cells in the presence of 5 μM IC261 and compared the morphology of the adherent cells after 20 hours. Control cells efficiently adhered to the plastic surface, spread out, and formed cytoskeletal actin networks that were visible as stress fibers and cytoplasmic protrusions (Figure 6e, f). In contrast, IC261-treated cells attached only loosely to the surface and exhibited a predominantly rounded morphology with no detectable stress fibers and protrusions (Figure 6e, f). To quantify the intensity of cell adhesion to the cell surface, we employed direct measurements of impedance caused by cell adhesion using the xCELLigence System (Roche Applied Science). This technology measures changes in the electrical conductivity produced by cells, which attach to golden electrodes that cover the surface of the well. The changes correspond to the area of cell that is in direct contact with the bottom of the well. These parameters allow for quantification of adhesion intensity (termed cell index) given the prerequisite that cell number and cell viability are identical between samples. CK1 inhibition dramatically decreased the cell index (Figure 6g). Because 2.5 μM IC261 did not affect the cell number (Figure 6h) or cell viability (89% vs. 87.6% for dimethylsulfoxide vs. 2.5 μM IC261), we interpreted changes in the cell index as decreased adhesion to the dish surface. In summary, these results suggest inhibition of CK1δ/ε prevents the formation of cell-cell and cell-surface contacts in epithelial MCF7 cells. The decrease in cell adhesion that was induced by CK1 inhibition and observed in different experimental setups (Figure 6) might be relevant to cell behavior associated with tumor progression, such as epithelial-mesenchymal transition or increased cell migration. This view is supported by our analysis of the E-cadherin promoter (Figure 7a). E-cadherin mediates intracellular contacts and is commonly used as a marker of epithelial fate, whose downregulation contributes to a shift towards mesenchymal fate [29]. Overexpression of WT CK1e in MCF7 cells slightly increased the activity of the E-cadherin promoter. In contrast, mutant forms of CK1e decreased reporter activity, and the effects of the P6 mutant were statistically significant from WT (Figure 7a). Similar results were obtained in HEK293 cells (Additional file 7). These data suggest that mutation of CK1e contributes to decreased adhesion through the regulation of E-cadherin expres- sion. Treatment with the CK1e inhibitor IC261 decreased the expression of endogenous E-cadherin in a dosedependent manner in MCF7 cells (Figure 7b). These changes in cell adhesion translate into changes of the migratory capacity of MCF7 cells in Transwell assays. As shown in Figure 7c, the dose-dependent inhibition of CK1δ/ε by IC261 promoted the migration of MCF7 cells. Similar effects were obtained with 10 μM D4476, which also increased the migration of MCF7 cells (Figure 7d). These results demonstrate that inhibition of CK1e or analogically mutants of CK1e can decrease cell adhesion and promote cell migration via mechanisms involving the activation of Wnt/Rac1 and Wnt/Ca2+ and the repression of E-cadherin expression. Discussion CK1ε is a crucial regulator of both the canonical Wnt/βcatenin pathway [14,15,17] and noncanonical Wnt pathways [10,11,30-32]. Wnt ligands from both classes which either activate β-catenin or do not - activate CK1e, which in turn phosphorylates Dvl [11,20]. Phosphorylated PS-Dvl is thought to mediate downstream signaling, which results in the stabilization and nuclear translocation of β-catenin or in the activation of some noncanonical pathways [11,32]. Our results demonstrate that the three CK1ε mutants analyzed in this study that were identified in samples of breast cancer efficiently bind, but fail to phosphorylate, Dvl and act as loss of function in the Wnt/β-catenin pathway. We demonstrate that cells expressing these mutants are biased towards Rac1/JNK and NFAT activation at the expense of the Wnt/β-catenin pathway. Importantly, phosphorylation of Dvl by CK1e was shown to be inhibitory for the noncanonical Wnt/Dvl/Rac1/JNK pathway [10,19]. CK1e can thus act as a switch that directs signaling away from the Wnt/Rac1/JNK branch of noncanonical signaling towards the Wnt/β-catenin and PS-Dvldependent noncanonical pathways. Wnt/Ca2+, the other Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 10 of 14 Figure 6 Casein kinase 1 epsilon inhibition decreases the cell adhesion of MCF7 breast epithelial cells. (a) to (d) MCF7 cells grown in hanging drop culture tend to form compact cell clusters under control conditions (dimethylsulfoxide (DMSO)). With increasing amounts of the casein kinase 1 epsilon (CK1ε) inhibitors - (a), (b) D4476, and (c), (d) IC261 - the large single-cluster formation is disrupted, and cells form smaller and disintegrated aggregates. After resuspending these aggregates by pipetting up and down seven times, a significantly higher number of single cells was observed to be released from clusters in conditions treated with (b) D4476 and (d) IC261, as normalized to control. ***P < 0.001, Student's t test, n = 3. (e), (f) MCF7 cells were seeded in the presence or absence of 5 μM IC261. Cell morphology was analyzed after 20 hours in (e) phase contrast or (f)after staining of actin filaments (phalloidin, green) and cell nuclei (Hoechst, blue). (g), (h) Dynamics of MCF7 cell adhesion was monitored for 20 hours using the xCELLigence System (Roche Applied Science). Cell adhesion is expressed as the cell index for two control wells (DMSO) and two experimental wells (2.5 μM IC261). (h) Treatment with IC261 did not affect the cell number (compared with control) during 20 hours of stimulation. DMSO D4476 (25 μM) D4476 (50 μM) DMSO IC261 (0.6 μM) IC261 (2.5 μM) A. B. D. E. DMSO IC261 (5 μM) F.DMSO IC261 (5 μM) 0 0,2 0,4 0,6 0,8 1 1,2 0 5 10 15 20 Time (hours) Cellindex DMSO I DMSO II IC261 (2.5 μM) - I IC261 (2.5 μM) - II H. D M SO IC 261 0 2000 4000 6000 8000 Numberofcells G. C. D M SO D 4476 10 +M D 4476 25 0.0 0.5 1.0 1.5 2.0 2.5 Numberofsinglecells *** D M SO IC 261 2.5 0 1 2 3 4 5 Numberofsinglecells +M +M Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 11 of 14 branch of noncanonical Wnt signaling, was shown to be antagonized by CK1 [33]. Our results confirm this observation and show that the CK1e mutants tested here increase the transcriptional activity of NFAT. These effects are not promoted by co-expression of Dvl (data not shown), suggesting that CK1e can repress NFAT directly. Indeed, phosphorylation by CK1e was shown to promote the cytoplasmic localization of NFAT and reduce its transcriptional activity in other experimental systems [34]. Together, our results provide the first evidence that the switch function of CK1ε, which was predicted based on Xenopus and cell culture experiments, is physiologically relevant and may contribute to cancer progression in ductal carcinoma. The observation that ductal carcinoma-specific mutants of CK1ε promote the Wnt/Rac1/JNK and NFAT pathways and, on the other hand, inhibit the Wnt/βcatenin pathway are in good agreement with some clinical observations. The expression of β-catenin in breast cancer is usually low and was shown by several laboratories to correlate with poor prognosis [35-38]. Moreover, β-catenin is not localized within the cell nucleus in breast tumors [35-38], suggesting that Wnt/β-catenin signaling is in the off state in most breast cancers. Important to note, however, is the fact that there are also reports supporting a positive role for Wnt/β-catenin signaling in breast cancer (for a review, see [39]). Wnt itself was first identified as an oncogene that is activated by the insertion of the mouse mammary tumor virus; and the mouse mammary tumor virus-Wnt-1 transgenic mouse is a well-established model for studies of the genetic basis of breast cancer (for a review, see [40]). Moreover, increased nuclear β-catenin levels, which correlate with cyclin D1 levels and a poor prognosis, were found in a subset of patients [41]. In contrast to other cancers, such as colon or skin cancer, the key components of the Wnt/β-catenin pathway, such as axin, adenomatous polyposis coli, or β-catenin are mutated in only a small portion of cases (for a review, see [39]). The levels of Wnt pathway components, which are common to both the canonical and noncanonical pathways, are altered more often; Dvl1 is upregulated in 50% of ductal breast cancer cases [42], and sFRP1 - a soluble Wnt antagonist that can block both Wnt/β-catenin and noncanonical pathways - is repressed in more than 80% of breast carcinomas [43]. These observations suggest that modulators of Wnt signaling, both at the extracellular level (sFRP1) and intracellular level (Dvl and CK1ε), will have a critical role in the biological outcome. Figure 7 Blocking casein kinase 1 epsilon function decreases E-cadherin expression and promotes cell migration. (a) Wild-type (WT) and mutant casein kinase 1 epsilon (CK1ε) were expressed in MCF7 cells together with the reporter encoding E-cadherin-promoter coupled to luciferase. Cells were lysed and the activity of luciferase was analyzed next day. (b) MCF7 cells were treated after seeding out by increasing concentrations of IC261 (2.5 μM, 5 μM and 10 μM) for 48 hours and the expression of E-cadherin was analyzed by western blotting. Empty arrowhead, unspecific band, which confirms equal protein loading. (c), (d) MCF7 cells were seeded out in the upper part of the Transwell migration chamber and were stimulated with indicated doses of (c) IC261 and (d) D4476. The number of migrating cells in the bottom chamber was counted next day, and is expressed as the migration index. Data represent mean ± standard deviation. *P< 0.05; one-way analysis of variance; (a), (c)Tukey post test, (d) Student's ttest; n = 3. DMSO, dimethylsulfoxide. A. D. Transwell assay D M SOD 4476 (10 μM ) 0.0 0.5 1.0 1.5 2.0 2.5 Migrationindex Transwell assay D M SO M ) IC 261 (1 M ) IC 261 (2.5 M ) IC 261 (10 0 1 2 3 Migrationindex C.E-cadherin reporter C TR L W T P3 P4 P6 0.0 0.5 1.0 1.5 Foldchange * * * B. DMSO E-cadherin IC261 μ μ μ Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 12 of 14 CK1ε has been shown to be highly expressed in noninvading ductal carcinoma in situ [1], which is also highly positive for β-catenin staining [36]. Based on our findings, we can speculate that the high activity of CK1ε in ductal carcinoma in situ reduces Rac-1/JNK and NFAT activity, keeping the cells tightly attached and blocking tumor invasion. CK1ε function can be compromised by somatic mutations [1], which dramatically affect CK1ε kinase activity (present study). The functional importance of this event supports the fact that samples with mutations in CK1ε show an increased frequency of loss of heterozygosity at the CSNK1ε locus [1]. Lack of CK1ε activity/expression leads to the activation of the Rac1/ JNK/AP1 and NFAT pathways, which mediates the invasion of breast cancer cells and correlates with increased aggressiveness of the breast cancer [33,44-47]. Based on our data and other published information, we propose that mutation of CK1ε might be important for the transition between ductal carcinoma in situ and invasive carci- noma. The effect of mutation or lack of CK1ε on Wnt signaling remains to be tested. In this context, it is especially important to determine how the status of CK1ε affects Wnt5a. Although Wnt5a is implicated in breast cancer pathology, its functional mechanism remains unclear. First, Wnt5a expression was shown to positively correlate with disease-free survival [48], and Wnt5a blocks breast cancer cell invasion [49-51]. On the other hand, Wnt5a is frequently upregulated in breast tumors in comparison with surrounding tissues [52] and was shown to promote the invasion of MCF7 cells via the JNK pathway [47]. We can speculate in cells with mutated CK1ε that Wnt5a will predominantly signal via the Rac-1/JNK and NFAT pathways, thus promoting breast cancer cell invasion and tumor aggressivity [44,45]. In contrast, in cells with abundant and intact CK1ε, Wnt5a principally stimulates other noncanonical signaling pathways that involve CK1, such as the Wnt/CK1ε/Rap1 [32] and Wnt/Yes-Cdc42-CK1α [33] pathways, and promotes cell adhesion [33]. In summary, our data demonstrate that the CK1ε mutants found in breast cancer act as loss of function, suppress jWnt/β-catenin, and promote Wnt/Rac-1-mediated and NFAT-mediated pathways. Furthermore, we show that inhibition of CK1ε reduces the intracellular adhesion and increases the migration of breast cancer cells. Our findings show that CK1ε has the potential to act as a tumor suppressor in breast cancer via its negative effects on the Wnt/Rac1/JNK and NFAT pathways. These results demonstrate for the first time how mutations in CK1ε affect cell behavior and may provide a general paradigm for consequences of CK1ε alteration in cancer. Conclusions In the present study, we functionally analyzed mutations in CK1ε that are frequently found in breast cancer. Our data demonstrate that breast cancer-specific mutants of CK1ε act as loss of function in the Wnt/β-catenin pathway but activate the Wnt/Rac1/JNK and Wnt/Ca2+ pathways. Physiological consequences of these signaling events in MCF7 breast cells include increased migratory capacity and decreased E-cadherin expression and cell adhesion. Additional material Additional file 1 Supplementary materials and methods. Materials and methods for construction of the recombinant protein expression constructs, for recombinant protein overexpression and purification, for in vitro kinase assays, for construction of vectors for mammalian expression, and for siRNA-mediated knockdown of CK1ε. Notes: pOPIN vectors [54] were used for bacterial overexpression. The GST-bPDZ construct was prepared as described previously [55]. siRNA-mediated knockdown performed as described previously [11]. Additional file 2 Interaction between casein kinase 1 epsilon mutants and Dishevelled. (a) HEK293 cell lysates transfected with either WT CK1ε or the P3, P4, and P6 mutants together with Dvl3-Flag were lysed and immunoprecipitated using an anti-Flag antibody. WT CK1ε and all of the CK1ε mutants efficiently bind the Dvl3 protein. (b) HEK293 cells were transfected with WT CK1ε or P3, P4 and P6 mutants together with Dvl2-Myc. Dvl2 protein localization in transfected HEK293 cells was observed by confocal microscopy using anti-Myc antibody. Dvl2 is either found in cytoplasmic inclusions or evenly dispersed within cytoplasm. Wt CK1ε co-transfected with Dvl2-Myc dissolves most of Dvl2 punctae, resulting in predominance of evenly distributed Dvl2 protein. In contrast, P3, P4 and P6 mutants are not able to promote even localization to the extent of WT CK1ε. The graphs indicate localization patterns (%) in 150 cells. Additional file 3 Recombinant CK1εΔC mutants exhibit different kinase activity. (a) His6-WT CK1εΔC phosphorylates the α and β isoforms of its natural substrate casein (1 to 3), while the kinase activity of the mutants maltose binding protein (MBP)-P3ΔC (L39Q, S101R) and MBP-P4ΔC (L39Q, L49Q, N78T) is strongly reduced (4 to 9). A single mutation in SUMOP6ΔC (L39Q) results in partial kinase activity of the CK1ε enzyme (10 to 12). (b) The bPDZ domain of Dvl is phosphorylated by individual recombinant kinases similarly as casein. (c) The CK1ε target sequence in Dvl2, which corresponds to residues 145 to 168 of hDvl1 (ENLEPETETESVVSLRRERPRRR), was prepared as a synthetic peptide. This sequence was phosphorylated by His6-WT CK1εΔC and partially phosphorylated by the SUMO-P6ΔC mutant. The MBP-P3ΔC and MBP-P4ΔC mutants were unable to phosphorylate this peptide. Additional file 4 Reporter assays with Dishevelled 3 protein. TopFlash and AP1 luciferase assays with Dvl3 protein confirm results obtained with Dvl2. Graphs indicate the mean ± standard deviation from three independent replicates. (a) Co-expression of Dvl3-Flag and WT CK1ε in HEK293 potentiates Wnt/β-catenin signaling, while CK1ε mutants have opposite effects and downregulate TCF/LEF mediated luciferase transcription. (b) Coexpression of Dvl3-Flag and WT CK1ε in MCF7 potentiates Wnt/β-catenin signaling, while CK1ε mutants are unable to do so, similarly to the situation in HEK293 cells. (c) CK1ε forms without Dvl overexpression do not elevate TCF/LEF-dependent transcription as compared with control empty plasmid. (d) Dvl3-Flag and WT CK1ε transfected in HEK293 cells decrease JNK/ AP1 signaling. In contrast, each mutant CK1ε together with Dvl3-Flag induces transcription from AP1 luciferase reporter. Additional file 5 Casein kinase 1 epsilon does not activate Cdc42. HEK293 cells were either transfected with CK1e forms or treated with 100 μM D4476 inhibitor 4 hours prior to lysis. Lysates from HEK293 cells were subjected to pull-down of active GTP-Cdc42 form with agarose-GST-WASP beads, which specifically interact only with the activated form of Cdc42. Amount of Cdc42 in pull-down (GTP-Cdc42) and in the original lysate (Cdc42) were detected by Cdc42 specific antibody using western blotting. Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 13 of 14 Abbreviations BSA: bovine serum albumin; CKIε: casein kinase 1 epsilon; Dvl: Dishevelled; GST: glutathione S-transferase; ORF: open reading frame; PBS: phosphate-buffered saline; PS-Dvl: phosphorylated and shifted Dishevelled; siRNA: small interfering RNA; WT: wild type. Competing interests The authors declare that they have no competing interests. Authors' contributions SF-T, PS, KT, OB, and VB carried out the molecular genetic and biochemical studies. SF-T, EB, WB, and LT participated in cloning, plasmid production and protein purification. PK, AK, and TD designed and coordinated the study. LT and VB designed and coordinated the study, and drafted the manuscript. All authors read and approved the final manuscript. Acknowledgements Molecular graphics images were produced using the UCSF Chimera package from the Resource for Biocomputing, Visualization, and Informatics at the University of California, San Francisco (supported by NIH P41 RR-01081). The authors thank Dr Christina Modak for providing cDNA templates for CK1ε and P3, P4 and P6 mutants. They also thank Dr Marek Mlodzik for GST-bPDZ constructs, Dr S Yanagawa (Kyoto University, Japan) for expression vectors encoding Dvl2-Myc, Dr Jan Kitajewski (Columbia University) for HA-Wnt3a-B1A cells, Dr Mikhail Semenov and Dr Xe He (Harvard Medical School) for vectors encoding mLrp6, Dr Saitoh (University of Tokyo, Japan) for E-cadherin reporter, and Dr Randy Moon (University of Washington, Seattle) for SuperTopFlash, ca-βcatenin and hDvl3 constructs. Plasmids for the overexpression of recombinant CK1ε were generated at the Dortmund Protein Facility [53]. The project was supported by the Grant Agency of the Czech Republic (301/07/0814 to LT and TD, 204/09/0498 and 204/09/H058 to VB, 301/09/0587 to PK), the Academy of Sciences of the Czech Republic (KJB501630801, AVOZ50040507, AVOZ50040702), the Ministry of Education, Youth and Sports (MSM0021622430) and by an EMBO Installation Grant to VB. Author Details 1Biology Centre AS CR, v.v.i. AND University of South Bohemia, Branisovska 31, 37005 Ceske Budejovice, Czech Republic, 2Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlarska 2, CZ-61137 Brno, Czech Republic, 3Department of Cytokinetics, Institute of Biophysics AS CR, Kralovopolska 135, 60200 Brno, Czech Republic, 4Max-Planck-Institute of Molecular Physiology, Otto-Hahn-Straße 11, 44227 Dortmund, Germany and 5Current address: Department of Chemistry, Faculty of Science, Utrecht University, Padualaan 8, NL-3584 CH Utrecht, The Netherlands References 1. Fuja TJ, Lin F, Osann KE, Bryant PJ: Somatic mutations and altered expression of the candidate tumor suppressors CSNK1 epsilon, DLG1, and EDD/hHYD in mammary ductal carcinoma. Cancer Res 2004, 64:942-951. 2. Knippschild U, Gocht A, Wolff S, Huber N, Lohler J, Stoter M: The casein kinase 1 family: participation in multiple cellular processes in eukaryotes. Cell Signal 2005, 17:675-689. 3. Lee JS, Ishimoto A, Yanagawa S: Characterization of mouse dishevelled (Dvl) proteins in Wnt/Wingless signaling pathway. J Biol Chem 1999, 274:21464-21470. 4. Angers S, Thorpe CJ, Biechele TL, Goldenberg SJ, Zheng N, MacCoss MJ, Moon RT: The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wnt-β-catenin pathway by targeting Dishevelled for degradation. Nat Cell Biol 2006, 8:348-357. 5. Kelley LA, Sternberg MJ: Protein structure prediction on the Web: a case study using the Phyre server. Nat Protoc 2009, 4:363-371. 6. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin TE: UCSF Chimera - a visualization system for exploratory research and analysis. J Comput Chem 2004, 25:1605-1612. 7. Marchler-Bauer A, Bryant SH: CD-Search: protein domain annotations on the fly. Nucleic Acids Res 2004, 32:W327-W331. 8. Parthiban V, Gromiha MM, Schomburg D: CUPSAT: prediction of protein stability upon point mutations. Nucleic Acids Res 2006, 34:W239-W242. 9. Xue Y, Ren J, Gao X, Jin C, Wen L, Yao X: GPS 2.0, a tool to predict kinasespecific phosphorylation sites in hierarchy. Mol Cell Proteomics 2008, 7:1598-1608. 10. Bryja V, Schambony A, Cajanek L, Dominguez I, Arenas E, Schulte G: Betaarrestin and casein kinase 1/2 define distinct branches of noncanonical WNT signalling pathways. EMBO Rep 2008, 9:1244-1250. 11. Bryja V, Schulte G, Rawal N, Grahn A, Arenas E: Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J Cell Sci 2007, 120:586-595. 12. Veeman MT, Slusarski DC, Kaykas A, Louie SH, Moon RT: Zebrafish prickle, a modulator of noncanonical Wnt/Fz signaling, regulates gastrulation movements. Curr Biol 2003, 13:680-685. 13. Horiguchi K, Shirakihara T, Nakano A, Imamura T, Miyazono K, Saitoh M: Role of Ras signaling in the induction of snail by transforming growth factor-beta. J Biol Chem 2009, 284:245-253. 14. Peters JM, McKay RM, McKay JP, Graff JM: Casein kinase I transduces Wnt signals. Nature 1999, 401:345-350. 15. Sakanaka C, Leong P, Xu L, Harrison SD, Williams LT: Casein kinase iepsilon in the wnt pathway: regulation of β-catenin function. ProcNatl Acad Sci USA 1999, 96:12548-12552. 16. McKay RM, Peters JM, Graff JM: The casein kinase I family in Wnt signaling. Dev Biol 2001, 235:388-396. 17. Kishida M, Hino S, Michiue T, Yamamoto H, Kishida S, Fukui A, Asashima M, Kikuchi A: Synergistic activation of the Wnt signaling pathway by Dvl and casein kinase Iepsilon. J Biol Chem 2001, 276:33147-33155. 18. Gonzalez-Sancho JM, Brennan KR, Castelo-Soccio LA, Brown AM: Wnt proteins induce dishevelled phosphorylation via an LRP5/6independent mechanism, irrespective of their ability to stabilize βcatenin. Mol Cell Biol 2004, 24:4757-4768. 19. Cong F, Schweizer L, Varmus H: Casein kinase Iepsilon modulates the signaling specificities of dishevelled. Mol Cell Biol 2004, 24:2000-2011. 20. Swiatek W, Tsai IC, Klimowski L, Pepler A, Barnette J, Yost HJ, Virshup DM: Regulation of casein kinase I epsilon activity by Wnt signaling. J Biol Chem 2004, 279:13011-13017. 21. Schwarz-Romond T, Merrifield C, Nichols BJ, Bienz M: The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J Cell Sci 2005, 118:5269-5277. 22. Smalley MJ, Signoret N, Robertson D, Tilley A, Hann A, Ewan K, Ding Y, Paterson H, Dale TC: Dishevelled (Dvl-2) activates canonical Wnt signalling in the absence of cytoplasmic puncta. J Cell Sci 2005, 118:5279-5289. 23. Scheeff ED, Bourne PE: Structural evolution of the protein kinase-like superfamily. PLoS Comput Biol 2005, 1:e49. 24. Korinek V, Barker N, Morin PJ, van Wichen D, de Weger R, Kinzler KW, Vogelstein B, Clevers H: Constitutive transcriptional activation by a betacatenin-Tcf complex in APC-/- colon carcinoma. Science 1997, 275:1784-1787. 25. Dickson RB, Bates SE, McManaway ME, Lippman ME: Characterization of estrogen responsive transforming activity in human breast cancer cell lines. Cancer Res 1986, 46:1707-1713. Additional file 6 siRNA-mediated knockdown of casein kinases decreases cell adhesion. MCF7 cells were transfected with either control siRNA or mixture of siRNAs targeted against CK1δ and CK1ε, and were subjected to the hanging drop assay next day. Cells were photographed 24 hours after seeding; cell clusters with typical morphology are presented. Knockdown of CK1δ and CK1ε decreases cell adhesion, which leads to the formation of looser cell aggregates. The efficiency of knockdown of CK1ε has been determined by western blotting, actin used as a loading control. Additional file 7 The effects of casein kinase 1 epsilon mutants on Ecadherin expression in HEK293 cells. WT and mutant CK1ε were expressed in HEK cells together with the reporter encoding E-cadherin-promoter coupled to luciferase. Cells were lysed and the activity of firefly luciferase, which reflects the activity of E-cadherin promoter, was analyzed next day. Renilla luciferase was used as an internal control. All results were normalized to Renilla and to the control transfection. Graph shows mean ± standard error of the mean from three independent experiments. Received: 6 November 2009 Revised: 4 April 2010 Accepted: 27 May 2010 Published: 27 May 2010 This article is available from: http://breast-cancer-research.com/content/12/3/R30© 2010 Foldynová-Trantírková et al.; licensee BioMed Central Ltd.This is an open access article distributed under the terms of the Creative Commons Attribution License http://creativecommons.org/licenses/by/2.0, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.Breast Cancer Research 2010, 12:R30 Foldynová-Trantírková et al. Breast Cancer Research 2010, 12:R30 http://breast-cancer-research.com/content/12/3/R30 Page 14 of 14 26. Rena G, Bain J, Elliott M, Cohen P: D4476, a cell-permeant inhibitor of CK1, suppresses the site-specific phosphorylation and nuclear exclusion of FOXO1a. EMBO Rep 2004, 5:60-65. 27. Bain J, Plater L, Elliott M, Shpiro N, Hastie CJ, McLauchlan H, Klevernic I, Arthur JS, Alessi DR, Cohen P: The selectivity of protein kinase inhibitors: a further update. Biochem J 2007, 408:297-315. 28. Mashhoon N, DeMaggio AJ, Tereshko V, Bergmeier SC, Egli M, Hoekstra MF, Kuret J: Crystal structure of a conformation-selective casein kinase- 1 inhibitor. J Biol Chem 2000, 275:20052-20060. 29. Huber MA, Kraut N, Beug H: Molecular requirements for epithelialmesenchymal transition during tumor progression. Curr Opin Cell Biol 2005, 17:548-558. 30. Klein TJ, Jenny A, Djiane A, Mlodzik M: CKIepsilon/discs overgrown promotes both Wnt-Fz/β-catenin and Fz/PCP signaling in Drosophila. Curr Biol 2006, 16:1337-1343. 31. Strutt H, Price MA, Strutt D: Planar polarity is positively regulated by casein kinase Iepsilon in Drosophila. Curr Biol 2006, 16:1329-1336. 32. Tsai IC, Amack JD, Gao ZH, Band V, Yost HJ, Virshup DM: A Wnt-CKIvarεRap1 pathway regulates gastrulation by modulating SIPA1L1, a Rap GTPase activating protein. Dev Cell 2007, 12:335-347. 33. Dejmek J, Safholm A, Kamp Nielsen C, Andersson T, Leandersson K: Wnt- 5a/Ca2+-induced NFAT activity is counteracted by Wnt-5a/Yes-Cdc42casein kinase 1α signaling in human mammary epithelial cells. Mol Cell Biol 2006, 26:6024-6036. 34. Okamura H, Garcia-Rodriguez C, Martinson H, Qin J, Virshup DM, Rao A: A conserved docking motif for CK1 binding controls the nuclear localization of NFAT1. Mol Cell Biol 2004, 24:4184-4195. 35. Dolled-Filhart M, McCabe A, Giltnane J, Cregger M, Camp RL, Rimm DL: Quantitative in situ analysis of β-catenin expression in breast cancer shows decreased expression is associated with poor outcome. Cancer Res 2006, 66:5487-5494. 36. Bankfalvi A, Terpe HJ, Breukelmann D, Bier B, Rempe D, Pschadka G, Krech R, Lelle RJ, Boecker W: Immunophenotypic and prognostic analysis of Ecadherin and β-catenin expression during breast carcinogenesis and tumour progression: a comparative study with CD44. Histopathology 1999, 34:25-34. 37. Bukholm IK, Nesland JM, Karesen R, Jacobsen U, Borresen-Dale AL: Ecadherin and α-, β-, and γ-catenin protein expression in relation to metastasis in human breast carcinoma. J Pathol 1998, 185:262-266. 38. Nakopoulou L, Gakiopoulou H, Keramopoulos A, Giannopoulou I, Athanassiadou P, Mavrommatis J, Davaris PS: c-met tyrosine kinase receptor expression is associated with abnormal β-catenin expression and favourable prognostic factors in invasive breast carcinoma. Histopathology 2000, 36:313-325. 39. Cowin P, Rowlands TM, Hatsell SJ: Cadherins and catenins in breast cancer. Curr Opin Cell Biol 2005, 17:499-508. 40. Li Y, Hively WP, Varmus HE: Use of MMTV-Wnt-1 transgenic mice for studying the genetic basis of breast cancer. Oncogene 2000, 19:1002-1009. 41. Lin SY, Xia W, Wang JC, Kwong KY, Spohn B, Wen Y, Pestell RG, Hung MC: Beta-catenin, a novel prognostic marker for breast cancer: its roles in cyclin D1 expression and cancer progression. Proc Natl Acad Sci USA 2000, 97:4262-4266. 42. Nagahata T, Shimada T, Harada A, Nagai H, Onda M, Yokoyama S, Shiba T, Jin E, Kawanami O, Emi M: Amplification, up-regulation and overexpression of DVL-1, the human counterpart of the Drosophila disheveled gene, in primary breast cancers. Cancer Sci 2003, 94:515-518. 43. Ugolini F, Charafe-Jauffret E, Bardou VJ, Geneix J, Adelaide J, Labat-Moleur F, Penault-Llorca F, Longy M, Jacquemier J, Birnbaum D, Pebusque MJ: WNT pathway and mammary carcinogenesis: loss of expression of candidate tumor suppressor gene SFRP1 in most invasive carcinomas except of the medullary type. Oncogene 2001, 20:5810-5817. 44. Yiu GK, Toker A: NFAT induces breast cancer cell invasion by promoting the induction of cyclooxygenase-2. J Biol Chem 2006, 281:12210-12217. 45. Jauliac S, Lopez-Rodriguez C, Shaw LM, Brown LF, Rao A, Toker A: The role of NFAT transcription factors in integrin-mediated carcinoma invasion. Nat Cell Biol 2002, 4:540-544. 46. Schnelzer A, Prechtel D, Knaus U, Dehne K, Gerhard M, Graeff H, Harbeck N, Schmitt M, Lengyel E: Rac1 in human breast cancer: overexpression, mutation analysis, and characterization of a new isoform, Rac1b. Oncogene 2000, 19:3013-3020. 47. Pukrop T, Klemm F, Hagemann T, Gradl D, Schulz M, Siemes S, Trumper L, Binder C: Wnt 5a signaling is critical for macrophage-induced invasion of breast cancer cell lines. Proc Natl Acad Sci USA 2006, 103:5454-5459. 48. Jonsson M, Dejmek J, Bendahl PO, Andersson T: Loss of Wnt-5a protein is associated with early relapse in invasive ductal breast carcinomas. Cancer Res 2002, 62:409-416. 49. Safholm A, Leandersson K, Dejmek J, Nielsen CK, Villoutreix BO, Andersson T: A formylated hexapeptide ligand mimics the ability of Wnt-5a to impair migration of human breast epithelial cells. J Biol Chem 2006, 281:2740-2749. 50. Safholm A, Tuomela J, Rosenkvist J, Dejmek J, Harkonen P, Andersson T: The Wnt-5a-derived hexapeptide Foxy-5 inhibits breast cancer metastasis in vivo by targeting cell motility. Clin Cancer Res 2008, 14:6556-6563. 51. Medrek C, Landberg G, Andersson T, Leandersson K: Wnt-5a-CKIα signaling promotes β-catenin/E-cadherin complex formation and intercellular adhesion in human breast epithelial cells. J Biol Chem 2009, 284:10968-10979. 52. Lejeune S, Huguet EL, Hamby A, Poulsom R, Harris AL: Wnt5a cloning, expression, and up-regulation in human primary breast cancers. Clin Cancer Res 1995, 1:215-222. 53. Dortmund Protein Facility [http://www.mpi-dortmund.mpg.de/misc/ dpf/] 54. Berrow NS, Alderton D, Sainsbury S, Nettleship J, Assenberg R, Rahman N, Stuart DI, Owens RJ: A versatile ligation-independent cloning method suitable for high-throughput expression screening applications. Nucleic Acids Res 2007, 35:e45. 55. Jenny A, Reynolds-Kenneally J, Das G, Burnett M, Mlodzik M: Diego and Prickle regulate Frizzled planar cell polarity signalling by competing for Dishevelled binding. Nat Cell Biol 2005, 7:691-697. doi: 10.1186/bcr2581 Cite this article as: Foldynová-Trantírková et al., Breast cancer-specific mutations in CK1? inhibit Wnt/?-catenin and activate the Wnt/Rac1/JNK and NFAT pathways to decrease cell adhesion and promote cell migration Breast Cancer Research 2010, 12:R30 1 SUPPLEMENTARY INFOMATION Breast cancer-specific mutations in CK1ε inhibit Wnt/β-catenin and activate the Wnt/Rac1/JNK and NFAT pathways to decrease cell adhesion and promote cell migration. Silvie Foldynová-Trantírková, Petra Sekyrová, Kateřina Tmejová, Eva Brumovská, Ondřej Bernatík, Wulf Blankenfeldt, Pavel Krejčí, Alois Kozubík, Tomáš Doležal, Lukáš Trantírek, Vítězslav Bryja Supplementary – FIGURE CAPTION Supplementary Figure 1 A. HEK 293 cell lysates transfected with either WT CK1ε or the P3, P4, and P6 mutants together with Dvl3-Flag were lysed and immunoprecipitated using an anti-Flag antibody. WT CK1ε and all of the CK1ε mutants efficiently bind the Dvl3 protein B. HEK 293 cells were transfected with wt CK1ε or P3, P4 and P6 mutants together with Dvl2-Myc. Dvl2 protein localization in transfected HEK293 cells was observed by confocal microscopy using anti-Myc antibody. Dvl2 is either found in cytoplasmic inclusions or evenly dispersed within cytoplasm. Wt CK1ε co-transfected with Dvl2Myc dissolves most of Dvl2 punctae resulting in predominance of evenly distributed Dvl2 protein. In contrast, P3, P4 and P6 mutants are not able to promote even localization to the extent of WT CK1ε. The graphs indicate localization patterns (%) in 150 cells. Supplementary Figure 2 TopFlash and AP-1 luciferase assays with Dvl3 protein confirm results obtained with Dvl2. Graphs indicate mean+SD from three independent replicates. A. Coexpression of Dvl3-Flag and wt CK1ε in HEK293 potentiates Wnt/β-catenin signaling, while CK1ε mutants have opposite effects and downregulate TCF/LEF mediated luciferase transcription. 2 B. Coexpression of Dvl3-Flag and wt CK1ε in MCF-7 potentiates Wnt/β-catenin signaling, while CK1ε mutants are unable to do so, similarly to the situation in HEK293 cells. C. CK1ε forms without Dvl overexpression do not elevate TCF/LEF dependent transcription as compared to control empty plasmid. D. Dvl3-Flag and wt CK1ε transfected in HEK 293 cells decrease JNK/AP-1 signaling. In contrast, each mutant CK1ε together with Dvl3-Flag induces transcription from AP- 1 luciferase reporter. Supplementary Figure 3 HEK 293 were either transfected with CK1ε forms or treated with 100 µM D4476 inhibitor 4 hours prior to lysis. Lysates from HEK 293 cells were subjected to pull-down of active GTP-Cdc42 form with agarose-GST-WASP beads, which specifically interact only with the activated form of Cdc42. Amount of Cdc42 in pulldown (GTP-Cdc42) and in the original lysate (Cdc42) were detected by Cdc42 specific antibody using Western blotting. Supplementary Figure 4 MCF-7 cells were transfected with either control siRNA or mixture of siRNAs targeted against CK1δ and CK1ε, and subjected to the hanging drop assay next day. Cells were photographed 24 hours after seeding, cell clusters with typical morphology are presented. Knockdown of CK1δ and CK1ε decreases cell adhesion, which leads to the formation of looser cell aggregates. The efficiency of knockdown of CK1ε has been determined by Western blotting, actin is used as a loading control. Supplementary Figure 5 WT and mutant CK1ε were expressed in HEK cells together with the reporter encoding Ecadherin-promoter coupled to luciferase. Cells were lysed and the activity of firefly luciferase, which reflects the activity of E-cadherin promoter was analyzed next day. Renilla luciferase was used as an internal control. All results were normalized to Renilla and to the control transfection. Graph shows mean+SEM from three independent experiments. 3 4 5 Supplementary information - Material and Methods Construction of the expression constructs, protein over-expression and purification The CKIε DNAs corresponding to truncated forms of CKIε lacking auto-inhibition domain in a forms of wild type and P3, P4, and P6 mutants were generated from full-lengths constructs, courtesy of Dr. C. Modak, by PCR. The constructs of CK1ε catalytical domain corresponding to wild-type, P3, P4, and P6 mutants were introduced via homologous recombination into four different bacterial expression vectors, pOPINE, pOPINF, pOPINM, and pOPINS [1], producing C-terminally His6-, N-terminally His6-, N-terminally maltose binging protein-, and N-terminally SUMO-tagged fusion proteins, respectively. All PCR products were sequenced in their entirety. The vectors were introduced into the (DE3)RIL E.coli over-expression strain. Subsequently, the transfected bacteria were screened for soluble expression of the individual proteins under three different growth and induction conditions. Soluble over-expression of CKIεΔC wild-type was observed only for C-terminally His6-tagged protein. Soluble P3ΔC and P4ΔC mutants were obtained in a form of N-terminally MBP-tagged fusion proteins. P6ΔC mutants provided soluble over-expression as N-terminally SUMO-tagged fusion protein. The CKIe gene constructs were expressed in E. coli BL21(DE3) Codon Plus RIL cells (Novagen). An overnight culture from a single colony was diluted (1:100) into fresh Terrific Broth (TB) auto-induction medium supplemented with 50 mg l−1 ampicillin and 34 mg l−1 chloramphenicol. The cells were grown at 310 K in the TB medium with vigorous shaking for 4 hours. Then, the temperature was set to 298 K and auto-induced gene expression was allowed for 24 hours. The cells were harvested by centrifugation for 15 min at 6000g, resuspended in 50 mM Na2HPO4 pH 7.5 300 mM NaCl supplemented with 2 mM βmercaptoethanol and then lysed with a fluidizer. The lysate was then centrifuged at 150 000g (Beckmann Optima L-­70K ultracentrifuge; Ti-45 rotor) for 30 min and the resulting supernatant filtered, adsorbed and eluted from Ni–NTA affinity resin. Fractions containing CKIε fusion proteins, as judged by sodium dodecyl sulfate (SDS) gel electrophoresis, were pooled, dialyzed against 50 mM sodium phosphate buffer (pH=7) and 150 mM NaCl and concentrated prior to purification using gel filtration on Superdex 75 (GE Healthcare) in the same buffer. The purified protein was then transferred into the storage buffer (50 mM sodum phosphate buffer pH=8.0, 300 mM NaCl, 0.1% Tween, 10% glycerol) and stored at -80 ºC. For all mutants, any attempt to cleave off the tags resulted in the protein precipitation. Therefore, fusion, recombinant proteins (His6-CKIεΔC, MBP-P3ΔC, MBP-P4ΔC, SUMO- 6 P6ΔC) were purified and used in in vitro phosphorylation assays in the uncleaved form. The GST-bPDZ construct was prepared as described previously [2]. Construction of vectors for mammalian expression Open reading frames of human CKIε cDNA wt full-length (residues: 1-416) and three mutated versions (P3, P4, and P6) in pET vectors, were amplified by PCR using following specific primers: CKIεFW sense 5’-aagcttatggagctacgtgtggg-3’, and CKIεRV antisense 5’tctagattacttcccgagatggtcaa-3’. The PCR products were purified and inserted into pGEM-Teasy vector. All fragments were subsequently cloned into pcDNA3 vector by HindIII – XbaI sites. The truncated versions of CKIεΔC wt and P3ΔC, P4ΔC, and P6ΔC mutants (residues: 1-315) were amplified by PCR using primers CKIε-NFW sense 5’-aagcttgagctacgtgtgggg-3’, and CKIε-NRV antisense 5’-tctagattacctctcctcgcgttcg-3’, cloned into pGEM-T easy followed by subcloning at the HindIII/PstI site of pHAK-B3. siRNA-mediated knockdown siRNA-mediated knockdown has been performed as described previously ([3]). Following siRNAs have been used: control (sc37007), CK1δ (sc29910) and CK1ε (sc29914) (all from Santa Cruz Biotechnology) References 1. Berrow NS, Alderton D, Sainsbury S, Nettleship J, Assenberg R, Rahman N, Stuart DI, Owens RJ: A versatile ligation-independent cloning method suitable for highthroughput expression screening applications. Nucleic Acids Res 2007, 35(6):e45. 2. Jenny A, Reynolds-Kenneally J, Das G, Burnett M, Mlodzik M: Diego and Prickle regulate Frizzled planar cell polarity signalling by competing for Dishevelled binding. Nat Cell Biol 2005, 7(7):691-697. 3. Bryja V, Schulte G, Rawal N, Grahn A, Arenas E: Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J Cell Sci 2007, 120(Pt 4):586-595. Vítězslav Bryja, 2014    Attachments      #11      K. Tanneberger, A.S. Pfister, K. Brauburger, J. Schneikert, M.V. Hadjihannas, V. Kriz, G.  Schulte, V. Bryja and J. Behrens (2011): Amer1/WTX couples Wnt‐induced formation of  PtdIns(4,5)P2 to LRP6 phosphorylation. EMBO J. 30: 1433‐1443.       Impact factor (2011): 9.205  Times cited (without autocitations, WoS, Feb 21st 2014): 19  Significance: The first analysis of the mechanism behind the positive role of Amer1 in  Wnt/β‐catenin  signaling.  This  study  shows  that  Amer1  can  bridge  Dishevelled‐ controlled  formation  of  phosphatidyl  inositol  phosphates  with  downstream  phosphorylation of Lrp6, a key step in the activation of the pathway.  Contibution of the author/author´s team: Fluorescence recovery after photobleaching  (FRAP) assays describing Wnt‐3a‐induced Amer1 membrane dynamics.                Amer1/WTX couples Wnt-induced formation of PtdIns(4,5)P2 to LRP6 phosphorylation Kristina Tanneberger1 , Astrid S Pfister1 , Katharina Brauburger1 , Jean Schneikert1 , Michel V Hadjihannas1 , Vitezslav Kriz2,3 , Gunnar Schulte4 , Vitezslav Bryja2,3 and Ju¨ rgen Behrens1, * 1 Nikolaus-Fiebiger-Center, University Erlangen-Nu¨rnberg, Erlangen, Germany, 2 Faculty of Science, Institute of Experimental Biology, Masaryk University, Brno, Czech Republic, 3 Department of Cytokinetics, Institute of Biophysics, Academy of Sciences of the Czech Republic, Brno, Czech Republic and 4 Department of Physiology and Pharmacology, Karolinska Institutet, Stockholm, Sweden Phosphorylation of the Wnt receptor low-density lipoprotein receptor-related protein 6 (LRP6) by glycogen synthase kinase 3b (GSK3b) and casein kinase 1c (CK1c) is a key step in Wnt/b-catenin signalling, which requires Wnt-induced formation of phosphatidylinositol 4,5bisphosphate (PtdIns(4,5)P2). Here, we show that adenomatous polyposis coli membrane recruitment 1 (Amer1) (also called WTX), a membrane associated PtdIns(4,5)P2binding protein, is essential for the activation of Wnt signalling at the LRP6 receptor level. Knockdown of Amer1 reduces Wnt-induced LRP6 phosphorylation, Axin translocation to the plasma membrane and formation of LRP6 signalosomes. Overexpression of Amer1 promotes LRP6 phosphorylation, which requires interaction of Amer1 with PtdIns(4,5)P2. Amer1 translocates to the plasma membrane in a PtdIns(4,5)P2-dependent manner after Wnt treatment and is required for LRP6 phosphorylation stimulated by application of PtdIns(4,5)P2. Amer1 binds CK1c, recruits Axin and GSK3b to the plasma membrane and promotes complex formation between Axin and LRP6. Fusion of Amer1 to the cytoplasmic domain of LRP6 induces LRP6 phosphorylation and stimulates robust Wnt/b-catenin signalling. We propose a mechanism for Wnt receptor activation by which generation of PtdIns (4,5)P2 leads to recruitment of Amer1 to the plasma membrane, which acts as a scaffold protein to stimulate phosphorylation of LRP6. The EMBO Journal advance online publication, 8 February 2011; doi:10.1038/emboj.2011.28 Subject Categories: signal transduction Keywords: Amer1; LRP6; PtdIns(4,5)P2; Wnt; WTX Introduction The Wnt/b-catenin signalling pathway regulates cell proliferation, differentiation and apoptosis, and has an important role during embryonic development, adult tissue homoeostasis and various diseases including cancer (Lustig and Behrens, 2003; Clevers, 2006). In the absence of extracellular Wnt ligands the levels of cytoplasmic b-catenin are kept low by the action of a multiprotein destruction complex that targets b-catenin for proteasomal degradation. The core components of this complex are the scaffold proteins Axin and its homologue Conductin (Axin2), the tumour suppressor adenomatous polyposis coli (APC) and glycogen synthase kinase 3b (GSK3b), which phosphorylates b-catenin and thereby earmarks it for ubiquitin-mediated degradation in the proteasome (MacDonald et al, 2009). The binding of Wnt ligands to the transmembrane receptors Frizzled (Fz) and low-density lipoprotein receptor-related protein 6 (LRP6) initiates a signalling cascade that results in the inhibition of b-catenin phosphorylation and degradation leading to the activation of b-catenin-dependent transcription (Huang and He, 2008; Angers and Moon, 2009; MacDonald et al, 2009). A key step after Wnt stimulation is the phosphorylation of the intracellular domain (ICD) of LRP6 at five reiterated PPPSPxS motifs and adjacent Ser/Thr clusters (Tamai et al, 2004; Davidson et al, 2005; Zeng et al, 2005; MacDonald et al, 2008; Supplementary Figure S7A). Phosphorylation at the PPPSPxS motifs (e.g., at Ser1490) is mediated by GSK3b, whereas the Ser/Thr clusters (e.g., at Thr1479) are phosphorylated by casein kinase 1g (CK1g) (Davidson et al, 2005; Zeng et al, 2005). The phosphorylated PPPSPxS motifs provide docking sites for Axin (Mao et al, 2001; Tamai et al, 2004; Davidson et al, 2005) and can directly inhibit the activity of GSK3b (Cselenyi et al, 2008; Piao et al, 2008; Wu et al, 2009). Phosphorylation of LRP6 by GSK3b requires binding of Dishevelled (Dvl) to Fz, which in turn leads to recruitment of the Axin/GSK3b complex (Zeng et al, 2008). In contrast, CK1g is constitutively localized to the plasma membrane (Davidson et al, 2005). It was recently suggested that coclustering of Fz-LRP6 receptors together with Axin and Dvl in so-called LRP6 signalosomes is involved in LRP6 phosphorylation (Bilic et al, 2007). Signalosome formation depends on the ability of Dvl to dynamically polymerize, which might provide a high density of phosphorylation sites for GSK3b and CK1g (Bilic et al, 2007; Schwarz-Romond et al, 2007a, b). Recent evidence indicates that the Wnt-induced generation of phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) at the plasma membrane is required for LRP6 phosphorylation by GSK3b as well as CK1g and for signalosome formation (Pan et al, 2008). This process is mediated by Dvl, which binds and activates phosphatidylinositol 4-kinase type IIa (PI4KIIa) and phosphatidylinositol-4-phosphate 5-kinase type Ib (PIP5KIb) sequentially acting to produce PtdIns(4,5)P2 (Pan et al, 2008; Qin et al, 2009).Received: 9 August 2010; accepted: 18 January 2011 *Corresponding author. Nikolaus-Fiebiger-Center, University ErlangenNu¨rnberg, Glu¨ckstr. 6, 91054 Erlangen, Germany. Tel.: þ 49 9131 8529109; Fax: þ 49 9131 8529111; E-mail: jbehrens@molmed.uni-erlangen.de The EMBO Journal (2011), 1–11 | & 2011 European Molecular Biology Organization |All Rights Reserved 0261-4189/11 www.embojournal.org &2011 European Molecular Biology Organization The EMBO Journal EMBO THE EMBO JOURNAL THE EMBO JOURNAL 1 Amer1 (APC membrane recruitment 1) was initially described by our group as an APC-binding protein, which can associate with the plasma membrane via two N-terminal PtdIns(4,5)P2-binding domains (Grohmann et al, 2007). Amer1 is identical to the tumour suppressor WTX (Wilms Tumour gene on the X chromosome) mutated in a significant fraction of Wilms tumours (Rivera et al, 2007) and in the inherited disease OSCS (osteopathia striata congenita with cranial sclerosis) (Jenkins et al, 2009). Amer1 is found in complexes with b-catenin and components of the b-catenin destruction machinery such as APC, Axin and b-TrCP, and can block canonical Wnt signalling by inducing proteasomal degradation of b-catenin (Major et al, 2007). Studies in Zebrafish and Xenopus also suggest a negative role of Amer1 in Wnt signalling (Major et al, 2007). Several key players involved in LRP6 phosphorylation have been identified (Fz, Dvl, Axin, GSK3b, CK1g and PtdIns(4,5)P2), but a coherent picture of their interactions and the sequence of events are still missing. In particular, the mechanism by which Wnt-induced PtdIns(4,5)P2 formation results in LRP6 phosphorylation has remained elusive. In the present study, we identify an unexpected role of Amer1 as a positive regulator of Wnt signalling acting at the LRP6 receptor level by showing that Amer1 links Wnt-induced formation of PtdIns(4,5)P2 to LRP6 phosphorylation. Results Amer1 is required for Wnt-induced LRP6 phosphorylation To investigate whether Amer1 has a function in Wnt signalling at the receptor level, we knocked down its expression using two different siRNAs and analysed Wnt-induced phosphorylation of LRP6 in plasma membrane fractions by western blotting. We found that Amer1 knockdown prevented LRP6 phosphorylation at Ser1490 and Thr1479 in Wnt3Atreated HEK293T cells (Figure 1A and B; Supplementary Figure S1A). Amer1 knockdown also prevented LRP6 phosphorylation in SW480 colon carcinoma cells (Supplementary Figure S1B). Reciprocally, overexpression Figure 1 Amer1 is required for Wnt-induced LRP6 phosphorylation and signalosome formation. (A, B) Amer1 is required for LRP6 phosphorylation at Ser1490 (A) and Thr1479 (B). HEK293T cells stably expressing LRP6-EGFP were transfected with the indicated siRNAs and incubated with Wnt3A for 1 h. Membrane fractions were analysed by western blotting. (C) Overexpression of Amer1 promotes LRP6 phosphorylation at Ser1490. HEK293T cells stably expressing VSVG-LRP6 were transiently transfected with EGFP-Amer1 and treated with Wnt3A for 20 min. Note that Amer1 is expressed in two splice variants represented by the two bands on the anti-GFP western blot (cf. Supplementary Figure S3A). In panels (A) and (B), the anti-Amer1 antibody detects only the larger splice variant of endogenous Amer1. (D, E) Amer1 is required for Wnt-induced signalosome formation. (D) Signalosome formation in HeLa cells co-expressing LRP6-EYFP and Fz8-EYFP transfected with the indicated siRNAs. Images show EYFP fluorescence with and without Wnt3A treatment for 1 h. Right-hand panels represent higher magnifications of the boxed regions from the left. Arrowheads point to signalosomes. Scale bar is 10 mm. (E) Quantification of signalosome formation in cells from (D). Error bars indicate s.e.m. from three independent experiments. Amer1 regulates LRP6 phosphorylation K Tanneberger et al The EMBO Journal &2011 European Molecular Biology Organization2 of Amer1 stimulated LRP6 phosphorylation at Ser1490 (Figure 1C; Supplementary Figure S1C). Notably, the previously described recruitment of Axin to the plasma membrane after Wnt stimulation was also abolished after knockdown of Amer1 (Figure 1A; Mao et al, 2001). These data show that Amer1 is essential for Wnt-induced LRP6 phosphorylation and Axin translocation to the plasma membrane and point to a positive role of Amer1 in the activation of Wnt signalling at the receptor level. Amer1 is required for Wnt-induced signalosome formation Given the role of Amer1 in Wnt receptor activation we analysed whether it is involved in signalosome formation. Signalosomes were detected by monitoring the aggregation of YFP-tagged LRP6 and Fz8 after Wnt3A stimulation. Indeed knockdown of Amer1 efficiently reduced formation of LRP6 signalosomes (Figure 1D and E). These results show that Amer1 is required for Wnt-induced signalosome formation and corroborate a role for Amer1 in the activation of the Wnt pathway at the receptor level. Membrane localization of Amer1 through PtdIns(4,5)P2 binding is required for its effect on LRP6 phosphorylation We next analysed whether LRP6 activation depends on plasma membrane localization of Amer1, which is mediated by two short domains in its N-terminus. These domains bind to PtdIns(4,5)P2 and are characterized by a high proportion of highly conserved lysine residues known for their capacity to interact with PtdIns(4,5)P2 (Kagan and Medzhitov, 2006; Grohmann et al, 2007; Supplementary Figure S2A). Mutation of seven of these lysine residues to alanine (Amer1(7mLys)) abolished interaction with PtdIns(4,5)P2 as well as membrane association of Amer1 (Supplementary Figure S2B and C). Importantly, this mutant was defective in stimulating LRP6 phosphorylation in HEK293T cells (Figure 2A, left panels). Because the Amer1(7mLys) mutant is enriched in the nucleus, this experiment does not rule out a role of Amer1 in the cytoplasm. Fusion of the Amer1(7mLys) mutant to a nuclear export sequence (NES-Amer1(7mLys)) led to cytoplasmic localization of Amer1, but did not restore its ability to stimulate LRP6 phosphorylation (Figure 2A, right panels; Supplementary Figure S2C). Together, these results show that PtdIns(4,5)P2-mediated membrane association of Amer1 is required for its effect on LRP6 phosphorylation and that cytoplasmic localization alone is not sufficient. Amer1 is expressed in two splice isoforms termed Amer1-S1 or Amer1-S2, which differ by the presence or absence of amino acids 50–326, comprising a large part of the membrane association/PtdIns(4,5)P2-binding domain (Supplementary Figure S3A; Jenkins et al, 2009). While Amer1-S1, which binds to the plasma membrane, stimulated LRP6 phosphorylation, Amer1-S2 lacking the membrane association failed to do so (Supplementary Figure S3B and C). Importantly, specific knockdown of the Amer1-S1 isoform reduced Wnt-induced LRP6 phosphorylation, whereas knockdown of Amer1-S2 had no effect (Figure 2B; Supplementary Figure S3A and D). These data further support the importance of membrane localization of Amer1 for its effect on LRP6 phosphorylation. Wnt induces plasma membrane translocation of Amer1, which requires the formation of PtdIns(4,5)P2 Next, we studied whether activation of Wnt signalling alters the association of Amer1 with the plasma membrane and whether PtdIns(4,5)P2 is involved. Wnt3A treatment induced a rapid increase of endogenous Amer1 in the plasma membrane fraction of HEK293T cells, whereas total amounts of Amer1 in whole cell lysates were not altered (Figure 3A). To determine whether Wnt-induced association of Amer1 with the plasma membrane depends on the formation of PtdIns(4,5)P2 we knocked down PI4KIIa, which was shown to be essential for PtdIns(4,5)P2 synthesis after Wnt stimulation (Pan et al, 2008). Wnt-induced plasma membrane recruitment of Amer1 was strongly reduced when PI4KIIa was knocked down by two different siRNAs (Figure 3B; Supplementary Figure S4). Neomycin binds PtdIns(4,5)P2 and can block its interaction with proteins (Gabev et al, 1989; Pilot et al, 2006). Pre-treatment of cells with neomycin abolished Wnt-induced membrane association of Amer1 and at the same time strongly reduced LRP6 phosphorylation and Figure 2 Membrane localization of Amer1 through PtdIns(4,5)P2 binding is required for its effect on LRP6 phosphorylation. (A) N-terminal lysine mutants of Amer1 (EGFP-Amer1(7mLys), EGFP-NES-Amer1(7mLys)) lacking PtdIns(4,5)P2 binding and membrane association are deficient for LRP6 phosphorylation. HEK293T cells stably expressing VSVG-LRP6 were transiently transfected with EGFP-tagged Amer1 constructs as indicated and treated with Wnt3A for 20 min. For details on the constructs, see Supplementary Figure S2A–C. (B) Effect of specific knockdown of Amer1 splice variants Amer1-S1 and Amer1-S2 on LRP6 phosphorylation. HeLa cells transfected with the indicated Amer1 siRNAs were incubated with Wnt3A for 1 h and endogenous LRP6 was examined by western blotting (WB). Specific RT–PCR for expression of the splice variants is shown below. See also Supplementary Figure S3A–D. In (A) and (B) numbers below the lanes reflect relative levels of phosphorylated LRP6 (pS1490) normalized to LRP6 as determined by densitometry. Experiments were repeated at least three times. Amer1 regulates LRP6 phosphorylation K Tanneberger et al &2011 European Molecular Biology Organization The EMBO Journal 3 Wnt-induced recruitment of Axin to the plasma membrane (Figure 3C). Together, these data demonstrate that Amer1 is translocated to the plasma membrane in a PtdIns(4,5)P2dependent manner after Wnt stimulation. Wnt3A decreases membrane dynamics of Amer1 In order to test for effects of Wnt stimulation on the membrane dynamics of Amer1, we employed fluorescence recovery after photobleaching (FRAP) methodology. Selected membrane regions of EGFP-Amer1-expressing cells were bleached by a laser pulse and the recovery of EGFP fluorescence was analysed for 150 s (see Figure 3D for a typical experiment). Pre-incubation with Wnt3A decreased the mobile pool of Amer1 from B74% (95% confidence interval: 73.1–75.1%) to 55% (95% confidence interval: 54.4–56.7%) and at the same time increased the halftime required for recovery of Amer1 from 13 to 18 s (Figure 3E; Supplementary Table). This indicates that Wnt signalling generates an immobile pool of Amer1 at the plasma membrane that does not readily exchange with neighbouring or cytoplasmic Amer1. Interestingly, pre-treatment of cells with neomycin restored recovery in the presence of Wnt3A suggesting that Wnt-induced changes in the mobility of membrane Amer1 are dependent on PtdIns(4,5)P2 (Figure 3E). Amer1 is required for PtdIns(4,5)P2-induced LRP6 phosphorylation We previously found that treatment of cells with ionomycin which activates phospholipase C and thereby leads to breakdown of PtdIns(4,5)P2 resulted in release of Amer1 from the plasma membrane (Varnai and Balla, 1998; Grohmann et al, 2007). Interestingly, ionomycin treatment abolished LRP6 phosphorylation after Wnt treatment (Figure 4A). In line, inhibition of PtdIns(4,5)P2 formation by knockdown of PI4KIIa prevents LRP6 phosphorylation after Wnt stimulation (Pan et al, 2008; Figure 4B). To analyse whether this is due to loss of Amer1 from the plasma membrane we asked whether tethering of Amer1 to the membrane independently of PtdIns(4,5)P2 would be able to restore LRP6 phosphorylation. Indeed, reduced LRP6 phosphorylation after knockdown of PI4KIIa was prevented by Amer1 when fused to the transmembrane domain from the low-density lipoprotein (LDL) receptor (Figure 4B; Zeng et al, 2005). Recent data demonstrate that carrier-mediated transfer of exogenous PtdIns(4,5)P2 lipids into cells enhances Wnt-induced LRP6 phosphorylation (Pan et al, 2008). We found that knockdown of Amer1 abolished the stimulation of LRP6 phosphorylation by PtdIns(4,5)P2 (Figure 4C). These data demonstrate that Amer1 mediates the effect of PtdIns(4,5)P2 on LRP6 phos- phorylation. Figure 3 Wnt induces plasma membrane translocation of Amer1, which requires the formation of PtdIns(4,5)P2. (A) Wnt3A induces plasma membrane translocation of Amer1. HEK293T cells stably expressing LRP6-EGFP were incubated with Wnt3A for 1 h and then subjected to subcellular fractionation. Cytoplasmic fractions (C), membrane fractions (M) and whole cell lysates (WCL) were analysed by western blotting. a-Tubulin and LRP6 were used to mark cytoplasmic and membrane fractions, respectively. (B) Knockdown of PI4KIIa prevents Wnt-induced plasma membrane recruitment of Amer1. siRNA transfected HeLa cells were incubated with Wnt3A for 1 h and membrane fractions were analysed by western blotting. (C) Neomycin abolishes Wnt-induced membrane translocation of Amer1. HEK293Tcells stably expressing VSVGLRP6 were treated with 10 mM neomycin for 30 min before incubation with Wnt3A plus neomycin for 1 h and membrane fractions were analysed by western blotting. (D, E) Wnt decreases membrane dynamics of Amer1. HEK293 cells transiently transfected with EGFP-Amer1 were subjected to FRAP analysis. (D) Typical example of EGFP-Amer1 distribution and changes in EGFP fluorescence before and after bleaching. Scale bar is 10 mm. (E) Statistical analysis of FRAP experiments using cells pre-stimulated with PBS (control) or Wnt3A with or without 10 mM neomycin as indicated (N, number of cells analysed per condition). The graphs show mean values and s.e.m., and the best fitting curve model, which was used for calculation of the mobile pool of EGFP-Amer1 (% of fluorescence recovered) and of the recovery halftime (T1/2). Amer1 regulates LRP6 phosphorylation K Tanneberger et al The EMBO Journal &2011 European Molecular Biology Organization4 Amer1 recruits Axin and GSK3b to the plasma membrane and promotes complex formation between Axin and LRP6 Our data show that Amer1 is required for Wnt-induced Axin translocation to the plasma membrane (see Figure 1A). Therefore, Amer1 might stimulate LRP6 phosphorylation through recruitment of Axin (Zeng et al, 2005, 2008). It is possible, however, that membrane association of Axin is a consequence rather than a cause of increased LRP6 phosphorylation induced by Amer1 because the phosphorylated PPPSPxS motifs in the cytoplasmic domain of LRP6 can serve as docking sites for Axin (Mao et al, 2001; Tamai et al, 2004; Davidson et al, 2005). To rule out this possibility, we treated cells with LiCl in order to inhibit GSK3b-mediated LRP6 phosphorylation and monitored Wnt-induced Axin translocation to the plasma membrane. LiCl treatment efficiently inhibited LRP6 phosphorylation but had no effect on Axin recruitment (Figure 5A). This shows that phosphorylation of LRP6 is not required for the association of Axin with the plasma membrane after Wnt treatment, suggesting that Amer1 promotes Axin translocation independently of prior LRP6 phosphorylation. We therefore analysed whether Amer1 can directly recruit Axin and the associated GSK3b. Amer1 formed endogenous complexes with Axin as shown by immunoprecipitation (Supplementary Figure S5A). In agreement with previous reports, Axin was diffusely distributed in the cytoplasm in a dotty pattern when exogenously expressed in MCF-7 cells (e.g., Schwarz-Romond et al, 2005; Figure 5B). In contrast, Axin was localized to the plasma membrane when Amer1 was expressed (Figure 5B; Supplementary Figure S5B). Similarly, endogenous Conductin was redistributed by Amer1 to the plasma membrane in SW480 colon carcinoma cells (Supplementary Figure S5C). Axin and Conductin were not redirected to the plasma membrane by Amer1(7mLys) mutants, indicating that membrane association of Amer1 is required (Supplementary Figure S5B and C). Importantly, in the presence but not in the absence of Axin Amer1 was also able to recruit GSK3b to the plasma membrane (Figure 5B). In line, GSK3b co-immunoprecipitated with Amer1 in the presence of wild-type Axin but not in the presence of a mutant that lacks GSK3b binding (AxinL396Q; Zeng et al, 2008), indicating that Axin proteins link GSK3b to Amer1 (Figure 5C). Together, these data suggest that Amer1 stimulates LRP6 phosphorylation by recruiting the Axin/GSK3b complex. In support of this, a C-terminal deletion mutant of Amer1 that retains Axin/ Conductin binding (Amer1(2–601)) stimulated LRP6 phosphorylation whereas a mutant lacking the Axin/Conductinbinding region (Amer1(2–530)) failed to do so (Figure 5D; Supplementary Figure S6A–C). Next, we analysed whether Amer1 binds to LRP6 by coimmunoprecipitation experiments. We found that Amer1 and Figure 4 Amer1 is required for PtdIns(4,5)P2-induced LRP6 phosphorylation. (A) Ionomycin treatment prevents Wnt-induced LRP6 phosphorylation. HEK293T cells stably expressing VSVG-LRP6 were treated with Wnt3A in the presence or absence of 10 mM ionomycin for 30 min. (B) Amer1DN (amino acids 207–1135) linked to the transmembrane domain of the LDL receptor (RFP-TMD-Amer1DN) rescues phosphorylation of endogenous LRP6 after knockdown of PI4KIIa. HeLa cells stably expressing RFP or RFP-TMD-Amer1DN were transfected with the indicated siRNAs and incubated with Wnt3A for 1 h. The numbers below the lanes reflect relative levels of phosphorylated LRP6 (pS1490) normalized to LRP6 as determined by densitometry. Data are representative of four independent experiments. (C) Amer1 is required for LRP6 phosphorylation stimulated by PtdIns(4,5)P2. HEK293Tcells stably expressing VSVG-LRP6 and transfected with the indicated siRNAs were treated with PtdIns(4,5)P2 for 10 min before Wnt3A conditioned medium was added for another 20 min (left). Quantification of LRP6 phosphorylation from four independent experiments (right). Statistical analysis was done using an unpaired Student’s t-test. The P-value reflects statistically significant differences. The black bars represent the experiment shown on the left. Amer1 regulates LRP6 phosphorylation K Tanneberger et al &2011 European Molecular Biology Organization The EMBO Journal 5 LRP6 form complexes after overexpression and at the endogenous level (Figure 5E and F). Amer1 also co-immunoprecipitated with LRP6 lacking the extracellular domain (LRP6DE(1–4); Supplementary Figure S7A and B). Immunoprecipitation with serial LRP6 C-terminal deletion mutants (Davidson et al, 2005) showed that Amer1 interacts with a fragment retaining the membrane proximal PPPSPxS motif and flanking Ser/Thr clusters (LRP6DE(1–4)D87; Supplementary Figure S7A and B). Deletion of the PPPSPxS motif abolished the interaction (LRP6DE(1–4)D127). Amer1 did not interact with a fragment consisting only of the PPPSPxS motif and Ser/Thr clusters, indicating that these motifs are not sufficient for Amer1 binding (LDLRDN-miniC; Supplementary Figure S7A and B). Moreover, alanine substitutions Figure 5 Amer1 recruits Axin and GSK3b to the plasma membrane and promotes complex formation between Axin and LRP6. (A) Inhibition of LRP6 phosphorylation by LiCl does not prevent Wnt-induced Axin translocation to the plasma membrane. HEK293T cells stably expressing EGFP-LRP6 were incubated with 50 mM LiCl for 30 min before Wnt3A treatment for 1 h and membrane fractions were analysed by western blotting. (B, C) Amer1 associates with GSK3b and recruits it to the plasma membrane via the interaction with Axin. (B) MCF-7 cells were cotransfected as indicated above the panels. Expressed proteins were detected by CFP and YFP fluorescence and anti-Flag immunofluorescence. Scale bar is 20 mm. (C) GSK3b co-immunoprecipitates with EGFP-Amer1 in the presence of Flag-Axin but not Flag-AxinL396Q, which is defective in GSK3b binding. (D) The binding of Amer1 to Axin/Conductin is required for its effect on LRP6 phosphorylation. HEK293T cells stably expressing VSVG-LRP6 were transiently transfected with EGFP or EGFP-tagged Amer1 mutants as detailed in Supplementary Figure S6A–C. The numbers below the lanes reflect relative levels of phosphorylated LRP6 (pS1490) normalized to LRP6 as determined by densitometry. Data are representative of four independent experiments. (E, F) Amer1 interacts with LRP6. (E) Co-immunoprecipitation of LRP6 with EGFP-Amer1 upon transient transfection of HEK293T cells stably expressing VSVG-LRP6. (F) Co-immunoprecipitation of endogenous Amer1 and LRP6 from lysates of HEK293T cells. Immunoprecipitations were performed with mouse anti-Amer1 or control IgG antibodies. (G) Amer1 links Axin to LRP6. VSVG-LRP6 stably expressed in HEK293Tcells co-immunoprecipitated with Flag-Axin in the presence but not in the absence of EGFP-Amer1. (H) Amer1 interacts with CK1g. Co-immunoprecipitation of endogenous CK1g and Amer1 from lysates of SW480 cells. Immunoprecipitations were performed with mouse anti-Amer1 or control IgG antibodies. As CK1g levels in lysates were very low, immunoprecipitation with anti-CK1g antibodies is shown. Amer1 regulates LRP6 phosphorylation K Tanneberger et al The EMBO Journal &2011 European Molecular Biology Organization6 of serines and threonines in the PPPSPxS motifs (LRP6m10; Zeng et al, 2005) did not affect Amer1 binding to LRP6 (data not shown). Together, these data demonstrate that Amer1 binds close to the signalling motifs phosphorylated by GSK3b and CK1g, but that phosphorylation of these motifs is not required for Amer1 binding. Deletion analysis of Amer1 demonstrated that both central and C-terminal parts interact with LRP6, suggesting that there are multiple LRP6 interaction sites in Amer1 (data not shown). Axin is suggested to form a complex with LRP6 (Mao et al, 2001; Tamai et al, 2004; Davidson et al, 2005). Because Amer1 binds to both Axin and LRP6, it might promote complex formation between the two proteins. Indeed, while LRP6 was only poorly co-immunoprecipitated with Axin in the absence of Amer1, it was strongly immunoprecipitated when Amer1 was present (Figure 5G). Conversely, Axin might link LRP6 to Amer1. However, this is unlikely, because Axin did not increase the amounts of LRP6 co-immunoprecipitated with Amer1 (Supplementary Figure S7C). In co-immunoprecipitation experiments endogenous complexes of Amer1 and CK1g were found (Figure 5H). Amer1 did not interact with a CK1g mutant lacking the membrane association domain (Supplementary Figure S8A and B). These results are consistent with the finding that Amer1 is necessary for LRP6 phosphorylation at the CK1g site Thr1479 (cf. Figure 1B; Supplementary Figure S1A). An Amer1–LRP6 fusion protein activates downstream Wnt/b-catenin signalling independently of Wnt Our data indicate that Amer1 acts by recruiting the Axin/GSK3b complex to LRP6. To test whether close proximity between Amer1 and LRP6 would be sufficient to trigger LRP6 phosphorylation and, consequently, downstream b-catenin signalling, we fused Amer1 to the ICD of LRP6 (Supplementary Figure S9). After transfection in HEK293T cells phosphorylation of the LRP6-ICD was strongly stimulated by fusion to Amer1, but not by fusion to the membrane targeting domain of Amer1 (amino acids 2–209) or to Amer1DN lacking this domain (Figure 6A and B). Consistent with this, the LRP6-ICD–Amer1 fusion protein robustly induced stabilization of cytoplasmic b-catenin and activation of a TCF/b-catenin-dependent transcriptional reporter (Figure 6C). In contrast, the LRP6-ICD alone or its fusion to the Amer1 deletion mutants had no or only a minor stimulatory activity. Thus, Amer1 can directly induce LRP6 phosphorylation and activation of downstream Wnt/ b-catenin signalling when fused to the cytoplasmic domain of LRP6. Notably, Amer1 alone did not stimulate b-catenin stabilization or reporter gene expression in spite of its ability to induce LRP6 phosphorylation (see Figure 1C; Supplementary Figure S1C). This is presumably due to its concurrent function in promoting b-catenin degradation (Major et al, 2007), and indicates that its negative regulatory activity is abolished in the LRP6-ICD–Amer1 fusion construct (see Discussion). Figure 6 Fusion of Amer1 to the cytoplasmic domain of LRP6 induces LRP6 phosphorylation and stimulates downstream Wnt/b-catenin signalling. (A, B) Amer1 fused to the LRP6 intracellular domain (ICD) suffices to induce LRP6 phosphorylation. (A) Phosphorylation of the EGFP-LRP6-ICD–Amer1 fusion constructs (Supplementary Figure S9) as determined by anti-pS1490 western blotting. The band indicated by the asterisk most likely represents the Amer1-S2 splice variant generated from the LRP6-ICD–Amer1 fusion, which does not stimulate phosphorylation of the LRP6-ICD. (B) Quantification of the experiment from (A) as determined by densitometry. The intensity of the pS1490 bands was normalized to the intensity of the GFP bands with the LRP6-ICD set to 1. (C) The LRP6-ICD–Amer1 fusion protein activates a b-catenin-dependent luciferase reporter stably expressed in HEK293T cells (Major et al, 2007) and induces stabilization of cytoplasmic b-catenin. Fold changes of luciferase activity were determined by normalization to the EGFP control. Error bars indicate s.d. Amer1 regulates LRP6 phosphorylation K Tanneberger et al &2011 European Molecular Biology Organization The EMBO Journal 7 Discussion It is meanwhile well established that LRP6 phosphorylation is a key and early event in Wnt signalling. However, although several key players are known in this process it has remained elusive how LRP6 is linked to the activating kinases GSK3b and CK1g and how Wnt-induced formation of PtdIns(4,5)P2 is involved. We propose that Amer1 is an essential intermediate in the process of Wnt receptor activation and that it has a specific role in connecting PtdIns(4,5)P2 to LRP6 phosphorylation. Mechanism of LRP6 phosphorylation induced by Amer1 Our knockdown and overexpression experiments showed that Amer1 is necessary and sufficient for LRP6 phosphorylation. We suggest that this function is based on the complex formation of Amer1 with Axin/GSK3b and CK1g, and on the Wnt-regulated dynamic interaction of these complexes with the plasma membrane. It was previously shown that Axin and its relative Conductin are crucial for LRP6 phosphorylation through binding to GSK3b (Zeng et al, 2005, 2008). We found that Amer1 interacts with Axin and the associated GSK3b and can link Axin to LRP6. When directly fused to the cytoplasmic domain of LRP6 Amer1 promotes phosphorylation of LRP6 at the GSK3b phosphorylation site Ser1490, and induces downstream Wnt/b-catenin signalling. This suggests that Amer1 acts as a scaffold protein to recruit GSK3b to LRP6 via Axin, and thereby promotes LRP6 phosphorylation, very much like Axin acts as a scaffold for b-catenin phosphorylation (Behrens et al, 1998; Ikeda et al, 1998). Amer1 also interacts with CK1g, the second kinase responsible for LRP6 phosphorylation. Unlike GSK3b, CK1g is constitutively present at the plasma membrane (Davidson et al, 2005) and it is not known how it is activated by Wnt signalling. It is possible that Amer1 recruits CK1g from the lateral plasma membrane to the vicinity of LRP6 receptors and thereby positions CK1g for optimal phosphorylation of LRP6. In line with a scaffolding role, Amer1 binds close to the membrane proximal phosphorylation sites of GSK3b and CK1g in LRP6 (Supplementary Figure S7A and B). Of interest, Amer1 was initially identified as an APCbinding protein, which contains three independent APC interaction sites (Supplementary Figure S3A). We noticed that Wnt-induced LRP6 phosphorylation does not depend on APC (Supplementary Figure S10), suggesting that this interaction is not of relevance for the function of Amer1 in LRP6 phosphorylation. From our immunoprecipitation experiments and published proteomic studies (Major et al, 2007), Amer1/Axin/GSK3b and Amer1/CK1g complexes seem to form constitutively in cells. This raises the question why these complexes are inactive with respect to LRP6 phosphorylation in the absence of Wnts and how they become activated by Wnt signalling. We found that Amer1 is recruited to the plasma membrane after Wnt stimulation through formation of PtdIns(4,5)P2, and that membrane localization of Amer1 is essential for stimulation of LRP6 phosphorylation. Moreover, Amer1 is required for plasma membrane localization of Axin after Wnt stimulation and can directly recruit Axin/GSK3b to the membrane. Altogether, this points to a crucial role of Wntinduced membrane localization of Amer1 for activation of LRP6 phosphorylation. In the absence of Wnts, a fraction of Amer1 together with its associated proteins might be present at the plasma membrane but at too low concentrations to promote LRP6 phosphorylation, whereas after Wnt stimulation the mere increase of plasma membrane bound Amer1 leads to activation of LRP6 phosphorylation. Moreover, it could well be that PtdIns(4,5)P2 is preferentially formed in the vicinity of the Fz/LRP6 receptor complexes due to local engagement of Dvl by Fz, which stimulates PtdIns(4,5)P2 synthesis after Wnt stimulation (Pan et al, 2008). This could generate a high density of binding sites for Amer1 and its associated kinases at the receptors and thereby allow efficient phosphorylation of LRP6. In support of this, increased PtdIns(4,5)P2 levels were found to be associated with LRP6 aggregates, that is, signalosomes, as compared with nonaggregated LRP6 in sucrose density fractions (Pan et al, 2008). Altogether, we suggest a model (Figure 7) in which PtdIns(4,5)P2 molecules formed after Wnt stimulation through Dvl serve as docking sites for Amer1 at the plasma membrane and attract Amer1 to LRP6. Amer1 recruits Axin/ GSK3b and CK1g to these sites, binds to LRP6 and promotes LRP6 phosphorylation. Thus, Amer1 seems to act as a scaffold protein at the plasma membrane that connects LRP6 receptors to the activating kinases in a Wnt-inducible and PtdIns(4,5)P2-dependent manner. This model implies that the PtdIns(4,5)P2-dependent recruitment of Amer1 is of functional relevance for LRP6 phosphorylation and positions Amer1 within the Wnt-Dvl-PtdIns(4,5)P2 pathway described previously (Tamai et al, 2004; Davidson et al, 2005; Zeng et al, 2005, 2008; Pan et al, 2008). In fact, our studies show that Amer1 is required for LRP6 phosphorylation stimulated by PtdIns(4,5)P2 and that fusion of Amer1 to an unrelated membrane targeting domain rescues LRP6 phosphorylation after inhibition of PtdIns(4,5)P2 synthesis by knockdown of PI4KIIa. Integration of Amer1 into current models of LRP6 phosphorylation According to an initiation-amplification model, LRP6 phosphorylation by GSK3b occurs in two steps. Initial recruitment of the Axin/GSK3b complex to the plasma membrane allows phosphorylation at PPPSPxS motifs such as at Ser1490 and thus creates docking sites for more Axin/GSK3b complexes, which in turn induce further phosphorylation at other Figure 7 A model for the involvement of Amer1 in PtdIns(4,5)P2mediated LRP6 phosphorylation. Wnt binding to the Fz-LRP6 receptor complex leads to recruitment of Dvl, which induces the formation of PtdIns(4,5)P2 by binding and activating the phosphatidylinositol kinases PI4KII and PIP5KI. The generation of PtdIns (4,5)P2 in regions of receptor activity triggers the recruitment of Amer1 proteins, which in turn promote LRP6 phosphorylation by recruiting Axin/GSK3b and CK1g to LRP6. Amer1 regulates LRP6 phosphorylation K Tanneberger et al The EMBO Journal &2011 European Molecular Biology Organization8 PPPSPxS sites resulting in amplification of the signal (Zeng et al, 2008; MacDonald et al, 2009). It was proposed that initiation occurs through Fz/Dvl-mediated recruitment of Axin/GSK3b complexes to LRP6 (Cliffe et al, 2003; Zeng et al, 2008). Amer1 may have a similar role as Dvl in mediating the initial phosphorylation steps because it can recruit Axin/GSK3b independently of prior phosphorylation of LRP6. In fact, Dvl and Amer1 might cooperate in the recruitment of Axin/GSK3b: Dvl might recruit Axin by direct interaction, and/or indirectly by promoting PtdIns(4,5)P2 synthesis and, as a consequence, Amer1 translocation to the plasma membrane. Amer1 does not seem to be conserved in invertebrates such as Drosophila where phosphorylation of the LRP6 orthologue arrow is presumably also critical. This may indicate that initiation relies entirely on the Dvl-dependent Axin recruitment in these organisms and that Amer1 provides an additional level of regulation in vertebrates. According to the signalosome model, Fz-LRP6 receptor pairs aggregate due to polymerization of Dvl and co-cluster with Axin leading to LRP6 phosphorylation (Bilic et al, 2007; Schwarz-Romond et al, 2007a, b). We found that Amer1 is essential for signalosome formation. This role of Amer1 is probably independent of its function in LRP6 phosphorylation because signalosome formation did not require LRP6 phosphorylation by CK1g (Bilic et al, 2007). Instead, Amer1 might be directly involved in receptor aggregation, for example, through its ability to connect LRP6 to Axin, which in turn can interact with Dvl. Our FRAP analysis shows that Amer1 is recruited to an immobile pool at the plasma membrane upon Wnt stimulation, which might correspond to signalosomes. So far, we have not been able to detect Amer1 in signalosomes probably owing to the suboptimal stoichiometry of the many overexpressed proteins in this assay. Of note, PtdIns(4,5)P2 was shown to be required for signalosome formation (Pan et al, 2008), possibly through interaction with Amer1. Amer1 has a dual positive and negative role in Wnt signalling Our results show that Amer1 acts as an activator of the Wnt signalling pathway at the LRP6 receptor level, whereas loss of function and biochemical studies suggest a role of Amer1 as a negative regulator of Wnt signalling by inducing degradation of b-catenin (Major et al, 2007). Indeed, overexpression of Amer1 repressed the activity of a TCF/b-catenin-dependent transcriptional reporter (Supplementary Figure S11). Thus, Amer1 has a dual functional role as an activator and inhibitor of the Wnt pathway. Because of its downstream role in b-catenin degradation only the negative regulatory role becomes apparent in loss of function studies. The dual role of Amer1 is similar to that of Axin and GSK3b, which are required for activation of LRP receptors but are also essential for b-catenin degradation. In the case of GSK3b, these functions may be spatially separated between plasma membrane and cytoplasm (Zeng et al, 2005, 2008). In contrast, both activation and inhibition of Wnt signalling by Amer1 require its localization in the plasma membrane (Figure 2; KT and JB unpublished data) raising the question as to how these activities are regulated. It is possible that the differential localization of Amer1 between the general membrane compartment and clustered Fz/LRP6 receptors determines the balance between inhibiting and activating functions of Amer1 and provides a switch mechanism between these activities. In the simplest model, a constitutive pool of Amer1 at the plasma membrane might be involved in the steady state degradation of b-catenin. After Wnt stimulation, Amer1 is recruited close to LRP6 leading to its phosphorylation as discussed above. Thereafter, phosphorylated LRP6 might directly inhibit Amer1 function in b-catenin degradation by blocking GSK3b via its phosphorylated PPPSPxS motifs (Cselenyi et al, 2008; Piao et al, 2008; Wu et al, 2009) and/ or by interfering with binding of Amer1 to interaction partners such as b-catenin and APC. In line, when present in a fusion protein with the ICD of LRP6 Amer1 can lead to stabilization of b-catenin and activation of Wnt/b-catenindependent transcription, indicating that the function of Amer1 in b-catenin degradation is blocked in the complex with LRP6 (Figure 6C). Alternatively, Amer1 might initially act as an activator of the signal and turn into an inhibitor, as part of a negative feedback regulation mediated by the TCF/b-catenin-dependent upregulation of Conductin (Lustig et al, 2002). Materials and methods Cell culture, Wnt treatment and subcellular fractionation HEK293Tcells stably expressing VSVG-LRP6 (Zeng et al, 2005) were kindly provided by X He and HEK293T cells stably expressing LRP6-EGFP (Kategaya et al, 2009) or pBAR/Renilla (Major et al, 2007) by RT Moon. HeLa cells stably expressing RFP-TMDAmer1DN or RFP were selected after transfection in medium containing 1 mg/ml geneticin (G418, Invitrogen) and enriched using a MoFlo high-speed cell sorter (Dako Cytomation). Wnt3A conditioned medium was produced from mouse L cells stably expressing Wnt3A (American Type Culture Collection CRL-2647) and added 48 h after transfection for knockdown experiments, or 16 h after transfection for overexpression experiments. Ionomycin was obtained from Calbiochem. Subcellular fractionation of cells was carried out using the ProteoJET Membrane Protein Extraction Kit (Fermentas) according to the manufacturer’s instructions. Plasmids The pEGFP-Amer1(7mLys) mutant was obtained by PCR mutagenesis. pEGFP-NES-Amer1(7mLys) was created by the insertion of an oligonucleotide coding for a consensus MAPKK-NES (NLVDLQKK LEELELDEQQ) (Henderson and Eleftheriou, 2000) between the EGFP- and Amer1-coding sequences of pEGFP-Amer1(7mLys). To obtain the pEGFP-LRP6-ICD–Amer1 fusion constructs, the coding sequence of the LRP6-ICD (residues 1394–1613) was inserted between the coding sequences of EGFP and the respective Amer1 constructs. mRFP-TMD-Amer1DN was generated by replacing the LRP6-coding sequence of mRFP-daLRP6 (Krieghoff et al, 2006) with the transmembrane domain of the LDL receptor (residues 781–849; Zeng et al, 2005) fused to the Amer1(207–1135)-coding sequence. To generate pEGFP-Amer1-S1, three single nucleotide changes (147 A4C, 150 T4C, 153 G4A) were introduced leading to the ablation of the internal splice donor site without changing the amino-acid sequence. pEGFP-Amer1-S2 contains an in-frame deletion of residues 50–326. Details of the plasmids are available upon request. Lipid-binding assay and PtdIns(4,5)P2 delivery GST-Amer1 proteins were expressed in Escherichia coli BL21 and freshly purified before the experiment using Glutathione Sepharose 4B beads (GE Healthcare) as described previously (Grohmann et al, 2007). Membrane Lipid Strips (Echelon Biosciences Inc.) were incubated with the GST fusion proteins at a concentration of 1 mg/ml at 41C overnight and detected by an anti-GST antibody, according to the manufacturer’s instructions. For delivery of PtdIns(4,5)P2 into cells, lipids and Carrier 3 (Echelon Biosciences Inc.) were preincubated at a 1:1 molar ratio (100mM final concentration each) for 10min at room temperature and then added to the cells at a final concentration of 10mM. After incubation of the cells for 10min at 371C, Amer1 regulates LRP6 phosphorylation K Tanneberger et al &2011 European Molecular Biology Organization The EMBO Journal 9 the same volume of Wnt3A conditioned medium was added for another 20min. Immunofluorescence microscopy Immunofluorescence stainings were performed as described previously (Hadjihannas et al, 2006; Grohmann et al, 2007). For signalosome experiments, cells were transfected with 100 pmol siRNA together with 600 ng LRP6-EYFP, 200 ng MESD, 200 ng Fz8EYFP, 200 ng Flag-Axin, 200 ng Flag-GSK3b and 100 ng CFP-Dvl2 using TransIT-TKO (Mirus, Madison, WI, USA). Photographs were taken with a CCD camera (Visitron, Munich, Germany) on a Zeiss Axioplan 2 microscope (Zeiss, Oberkochen, Germany,  63 objective or  100 objective) and MetaMorph software (Molecular Devices). Amer1 antibodies The mouse monoclonal antibody against Amer1 has been described before (Grohmann et al, 2007). The Amer1-specific rabbit polyclonal antibody was produced by Pineda (Berlin, Germany) by immunizing rabbits with amino acids 2–285 of recombinant human Amer1 generated as a GST fusion in bacteria. The serum was affinity purified using CNBr-activated Sepharose 4B beads (GE Healthcare) coupled to GST-Amer1(2–285). Reporter assays b-Catenin reporter assays were carried out in HEK293T cells stably expressing a b-catenin responsive firefly luciferase reporter (pBAR) along with a Renilla luciferase, which serves as an internal control (Major et al, 2007). Cells were seeded in 12-well plates, transfected with 200 ng of the indicated constructs and harvested 24 h posttransfection. Firefly and Renilla luciferase activities were determined according to standard procedures, and firefly luciferase values were normalized to Renilla values. All experiments were performed in duplicates and reproduced at least twice. Fluorescence recovery after photobleaching HEK293 cells were plated on 35-mm glass bottom dish (MatTek Corp.) pre-coated with 0.1% collagen (Sigma) and transfected with the calcium phosphate method using 1.6 mg of EGFP-Amer1 plasmid/dish. FRAP analysis was carried out 24 h after transfection on a Zeiss LSM710 laser scanning microscope (488 nm laser line) with a C-Apochromat  40 water immersion objective, NA ¼ 1.2 (Zeiss, Jena, Germany). During live cell imaging, cells were kept at 371C in serum-free DMEM supplemented with 20 mM HEPES and 0.1% BSA. Stimulation with 100 ng/ml purified, recombinant Wnt3A (R&D Systems) and/or 10 mM neomycin (Sigma) was performed for 30 min prior image acquisition. One or two B5 mm2 square regions of interest (ROI) were selected per cell. Basal signal intensity was measured in 1 s intervals for 18 s. Applied laser power was 1% to minimize photobleaching. Then, the ROI was bleached for 4 s with two scans with maximum power using the 488-nm line argon laser. Fluorescence recovery was recorded in 1 s intervals for 154 s. Image processing and data analysis were done using the Zen 2009 software (Zeiss, Jena, Germany). The raw data were normalized with the average fluorescence of 20 data points before bleaching as 100% and the first value after bleaching as 0%. Sum of all datasets with stable recovery have been statistically analysed and the quantitative information (mobile pool, recovery halftime) was obtained by the non-linear one phase association using the least squares fitting method. The normalization and statistical analysis have been performed using GraphPad Prism software (GraphPad Software Inc.). RT–PCR RT–PCR was performed according to standard protocols (see Supplementary Experimental Procedures) using the following sense and antisense primers: Amer1-S1 (50 -GCGAATTCGGAGACCC AAAAGGATGAAGCTGCTCAG-30 , 50 -CCTTGCTCTTCCGGTGACGGC GGATACTGC-30 ), Amer1-S2 (50 -GCGAATTCGGAGACCCAAAAGG ATGAAGCTGCTCAG-30 , 50 -CATCATCATCTGGCAAGGCCATCTC-30 ) and GAPDH (50 -CTTCACCACCATGGAGAAGG-30 , 50 -CCTGCTTCACC ACCTTCTTG-30 ). Note that PCR for Amer1-S2 leads to the coamplification of products from Amer1-S1 and Amer1-S2 as the primers flank the alternatively spliced intron. GAPDH was used for normalization. Supplementary data Supplementary data are available at The EMBO Journal Online (http://www.embojournal.org). Acknowledgements We thank C Niehrs, X He, RT Moon, P De Camilli and A Kikuchi for cell lines and reagents; J Masˇek for technical assistance with FRAP analysis; I Wacker for helpful discussions; A Do¨bler for secretarial assistance. This work was supported by the DFG research grant BE 1550/6-1 and the IZKF research grant TPD12 to Ju¨rgen Behrens and the DFG research grant SCHN 1189/1-1 to Jean Schneikert. Vitezslav Bryja is supported by grants from the Czech Science Foundation (204/09/0498, 204/09/J030), EMBO Installation Grant, the Ministry of Education Youth and Sports of the Czech Republic (MSM 0021622430) and by Academy of Sciences of the Czech Republic (AVOZ50040507, AVOZ50040702). Gunnar Schulte is supported by Knut and Alice Wallenberg Foundation (KAW2008.0149), Swedish Research Council (K2008-68P-20810-01-4, K2008-68X-20805-01-4) and The Swedish Foundation for International Cooperation in Research and Higher Education (STINT). Author contributions: KT planned and performed most of the experimental work. KT, ASP, KB, MVH and JS performed co-immunoprecipitation and immunofluorescence experiments. ASP generated the Amer1-specific rabbit polyclonal antibody. VK, GS and VB performed FRAP experiments. JB coordinated the project and assisted with planning the experiments and data analysis. The manuscript was written by JB and KT. Conflict of interest The authors declare that they have no conflict of interest. References Angers S, Moon RT (2009) Proximal events in Wnt signal transduction. Nat Rev Mol Cell Biol 10: 468–477 Behrens J, Jerchow BA, Wurtele M, Grimm J, Asbrand C, Wirtz R, Kuhl M, Wedlich D, Birchmeier W (1998) Functional interaction of an axin homolog, conductin, with beta-catenin, APC, and GSK3beta. Science 280: 596–599 Bilic J, Huang YL, Davidson G, Zimmermann T, Cruciat CM, Bienz M, Niehrs C (2007) Wnt induces LRP6 signalosomes and promotes dishevelled-dependent LRP6 phosphorylation. Science 316: 1619–1622 Clevers H (2006) Wnt/beta-catenin signalling in development and disease. Cell 127: 469–480 Cliffe A, Hamada F, Bienz M (2003) A role of dishevelled in relocating axin to the plasma membrane during wingless signalling. Curr Biol 13: 960–966 Cselenyi CS, Jernigan KK, Tahinci E, Thorne CA, Lee LA, Lee E (2008) LRP6 transduces a canonical Wnt signal independently of axin degradation by inhibiting GSK3’s phosphorylation of betacatenin. Proc Natl Acad Sci USA 105: 8032–8037 Davidson G, Wu W, Shen J, Bilic J, Fenger U, Stannek P, Glinka A, Niehrs C (2005) Casein kinase 1 gamma couples Wnt receptor activation to cytoplasmic signal transduction. Nature 438: 867–872 Gabev E, Kasianowicz J, Abbott T, McLaughlin S (1989) Binding of neomycin to phosphatidylinositol 4,5-bisphosphate (PIP2). Biochim Biophys Acta 979: 105–112 Grohmann A, Tanneberger K, Alzner A, Schneikert J, Behrens J (2007) AMER1 regulates the distribution of the tumor suppressor APC between microtubules and the plasma membrane. J Cell Sci 120: 3738–3747 Hadjihannas MV, Bruckner M, Jerchow B, Birchmeier W, Dietmaier W, Behrens J (2006) Aberrant Wnt/beta-catenin signaling can induce chromosomal instability in colon cancer. Proc Natl Acad Sci USA 103: 10747–10752 Amer1 regulates LRP6 phosphorylation K Tanneberger et al The EMBO Journal &2011 European Molecular Biology Organization10 Henderson BR, Eleftheriou A (2000) A comparison of the activity, sequence specificity, and CRM1-dependence of different nuclear export signals. Exp Cell Res 256: 213–224 Huang H, He X (2008) Wnt/beta-catenin signalling: new (and old) players and new insights. Curr Opin Cell Biol 20: 119–125 Ikeda S, Kishida S, Yamamoto H, Murai H, Koyama S, Kikuchi A (1998) Axin, a negative regulator of the Wnt signalling pathway, forms a complex with GSK-3beta and beta-catenin and promotes GSK-3beta-dependent phosphorylation of beta-catenin. EMBO J 17: 1371–1384 Jenkins ZA, van Kogelenberg M, Morgan T, Jeffs A, Fukuzawa R, Pearl E, Thaller C, Hing AV, Porteous ME, Garcia-Minaur S, Bohring A, Lacombe D, Stewart F, Fiskerstrand T, Bindoff L, Berland S, Ades LC, Tchan M, David A, Wilson LC et al (2009) Germline mutations in WTX cause a sclerosing skeletal dysplasia but do not predispose to tumorigenesis. Nat Genet 41: 95–100 Kagan JC, Medzhitov R (2006) Phosphoinositide-mediated adaptor recruitment controls Toll-like receptor signalling. Cell 125: 943–955 Kategaya LS, Changkakoty B, Biechele T, Conrad WH, Kaykas A, Dasgupta R, Moon RT (2009) Bili inhibits Wnt/beta-catenin signalling by regulating the recruitment of axin to LRP6. PLoS One 4: e6129 Krieghoff E, Behrens J, Mayr B (2006) Nucleo-cytoplasmic distribution of {beta}-catenin is regulated by retention. J Cell Sci 119: 1453–1463 Lustig B, Behrens J (2003) The Wnt signalling pathway and its role in tumor development. J Cancer Res Clin Oncol 129: 199–221 Lustig B, Jerchow B, Sachs M, Weiler S, Pietsch T, Karsten U, van de Wetering M, Clevers H, Schlag PM, Birchmeier W, Behrens J (2002) Negative feedback loop of Wnt signaling through upregulation of conductin/axin2 in colorectal and liver tumors. Mol Cell Biol 22: 1184–1193 MacDonald BT, Tamai K, He X (2009) Wnt/beta-catenin signalling: components, mechanisms, and diseases. Dev Cell 17: 9–26 MacDonald BT, Yokota C, Tamai K, Zeng X, He X (2008) Wnt signal amplification via activity, cooperativity, and regulation of multiple intracellular PPPSP motifs in the Wnt co-receptor LRP6. J Biol Chem 283: 16115–16123 Major MB, Camp ND, Berndt JD, Yi X, Goldenberg SJ, Hubbert C, Biechele TL, Gingras AC, Zheng N, Maccoss MJ, Angers S, Moon RT (2007) Wilms tumor suppressor WTX negatively regulates WNT/beta-catenin signaling. Science 316: 1043–1046 Mao J, Wang J, Liu B, Pan W, Farr III GH, Flynn C, Yuan H, Takada S, Kimelman D, Li L, Wu D (2001) Low-density lipoprotein receptor-related protein-5 binds to axin and regulates the canonical Wnt signaling pathway. Mol Cell 7: 801–809 Pan W, Choi SC, Wang H, Qin Y, Volpicelli-Daley L, Swan L, Lucast L, Khoo C, Zhang X, Li L, Abrams CS, Sokol SY, Wu D (2008) Wnt3a-mediated formation of phosphatidylinositol 4,5-bisphosphate regulates LRP6 phosphorylation. Science 321: 1350–1353 Piao S, Lee SH, Kim H, Yum S, Stamos JL, Xu Y, Lee SJ, Lee J, Oh S, Han JK, Park BJ, Weis WI, Ha NC (2008) Direct inhibition of GSK3beta by the phosphorylated cytoplasmic domain of LRP6 in Wnt/beta-catenin signaling. PLoS One 3: e4046 Pilot F, Philippe JM, Lemmers C, Lecuit T (2006) Spatial control of actin organization at adherens junctions by a synaptotagmin-like protein Btsz. Nature 442: 580–584 Qin Y, Li L, Pan W, Wu D (2009) Regulation of phosphatidylinositol kinases and metabolism by Wnt3a and Dvl. J Biol Chem 284: 22544–22548 Rivera MN, Kim WJ, Wells J, Driscoll DR, Brannigan BW, Han M, Kim JC, Feinberg AP, Gerald WL, Vargas SO, Chin L, Iafrate AJ, Bell DW, Haber DA (2007) An X chromosome gene, WTX, is commonly inactivated in Wilms tumor. Science 315: 642–645 Schwarz-Romond T, Fiedler M, Shibata N, Butler PJ, Kikuchi A, Higuchi Y, Bienz M (2007a) The DIX domain of dishevelled confers Wnt signalling by dynamic polymerization. Nat Struct Mol Biol 14: 484–492 Schwarz-Romond T, Merrifield C, Nichols BJ, Bienz M (2005) The Wnt signalling effector dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J Cell Sci 118: 5269–5277 Schwarz-Romond T, Metcalfe C, Bienz M (2007b) Dynamic recruitment of axin by dishevelled protein assemblies. J Cell Sci 120: 2402–2412 Tamai K, Zeng X, Liu C, Zhang X, Harada Y, Chang Z, He X (2004) A mechanism for Wnt coreceptor activation. Mol Cell 13: 149–156 Varnai P, Balla T (1998) Visualization of phosphoinositides that bind pleckstrin homology domains: calcium- and agonist-induced dynamic changes and relationship to myo-[3H]inositol-labeled phosphoinositide pools. J Cell Biol 143: 501–510 Wu G, Huang H, Garcia Abreu J, He X (2009) Inhibition of GSK3 phosphorylation of beta-catenin via phosphorylated PPPSPXS motifs of Wnt coreceptor LRP6. PLoS One 4: e4926 Zeng X, Huang H, Tamai K, Zhang X, Harada Y, Yokota C, Almeida K, Wang J, Doble B, Woodgett J, Wynshaw-Boris A, Hsieh JC, He X (2008) Initiation of Wnt signaling: control of Wnt coreceptor Lrp6 phosphorylation/activation via frizzled, dishevelled and axin functions. Development 135: 367–375 Zeng X, Tamai K, Doble B, Li S, Huang H, Habas R, Okamura H, Woodgett J, He X (2005) A dual-kinase mechanism for Wnt co-receptor phosphorylation and activation. Nature 438: 873–877 Amer1 regulates LRP6 phosphorylation K Tanneberger et al &2011 European Molecular Biology Organization The EMBO Journal 11 SUPPLEMENTARY INFORMATION “Amer1/WTX couples Wnt-induced formation of PtdIns(4,5)P2 to LRP6 phosphorylation” Kristina Tanneberger, Astrid S. Pfister, Katharina Brauburger, Jean Schneikert, Michel V. Hadjihannas, Vitezslav Kriz, Gunnar Schulte, Vitezslav Bryja and Jürgen Behrens SUPPLEMENTARY FIGURES Supplementary Figure S1 (A, B) Amer1 is required for phosphorylation of endogenous LRP6 at Ser1490 and Thr1479. Cells were transfected with the indicated siRNAs and incubated with Wnt3A for 1 hour. Membrane fractions from HEK293T cells (A) and whole cell lysates from SW480 cells (B) were analysed by Western blotting. (C) Overexpression of Amer1 promotes phosphorylation of endogenous LRP6 at Ser1490. HEK293T cells were transfected with EGFP-Amer1 and treated with Wnt3A for 20 minutes. Membrane fractions were analysed by Western blotting. 2 Supplementary Figure S2 (A) Sequence comparison of the two N-terminal PtdIns(4,5)P2 binding sites of Amer1 from different species and of Amer2 (Grohmann et al, 2007). Highly conserved basic and aromatic residues are highlighted in red. Red boxes indicate lysine residues that have been mutated to alanine in this study. (B) Phospholipid-binding assays of Amer1 mutants. Membrane lipid strips were incubated with the indicated GST-Amer1 fusion proteins revealing a severely diminished ability of Amer1(7µLys) to bind to phosphatidylinositol lipids. The Coomassie staining below shows relative amounts of purified proteins used. (C) Fluorescence micrographs of MCF-7 cells transfected with EGFP-tagged Amer1 constructs as indicated above the panels. Scale bar is 20 µm. 3 Supplementary Figure S3 (A) Schematic representation of Amer1 splice variants Amer1-S1 and Amer1-S2 (Jenkins et al, 2009). In Amer1-S2 amino acids 50-326 are deleted in frame. The localization of target sequences of Amer1 siRNAs is depicted. siAmer1a targets both Amer1 splice variants while siAmer1-S1 and siAmer1-S2 specifically target Amer1-S1 and Amer1-S2, respectively. The two N-terminal membrane localization domains (M1, M2) are highlighted by light grey and the three APC interaction domains (A1-3) by dark grey shading (Grohmann et al, 2007). (B) Fluorescence micrographs of MCF-7 cells transfected with EGFP-tagged Amer1 splice variants as indicated above the panels. Scale bar is 20 µm. (C) Effect of Amer1 splice variants on LRP6 phosphorylation when overexpressed in HEK293T cells stably expressing VSVGLRP6. (D) Specific knockdown of Amer1 splice variants after transient transfection of Flag-tagged Amer1 together with the indicated siRNAs into HEK293T cells. 4 Supplementary Figure S4 Knockdown of PI4KIIα prevents Wnt-induced plasma membrane recruitment of Amer1. HEK293T cells were transfected with two different siRNAs, incubated with Wnt3A for 1 hour and membrane fractions were analysed by Western blotting. AntiPan-Cadherin Western blot is shown as loading control for membrane fractions. 5 Supplementary Figure S5 (A) Co-immunoprecipitation of endogenous Axin and Amer1 from lysates of SW480 cells. Immunoprecipitations were performed with mouse anti-Amer1 or control IgG antibodies. (B, C) Amer1 leads to plasma membrane recruitment of endogenous Axin (B) or Conductin (C). Amer1 mutants Amer1(7µLys) and NES-Amer1(7µLys), which are defective in membrane localization, are not able to recruit Axin/Conductin. (B) HEK293T cells were transfected with the indicated EGFP-Amer1 constructs and membrane fractions were analysed by Western blotting. VSVG-LRP6 stably expressed in these cells is shown as loading control for membrane fractions. (C) Double staining of EGFP-tagged Amer1 and the indicated Amer1 mutants (left panels, GFP fluorescence) and Conductin (middle panels, anti-Conductin immunofluorescence) in SW480 cells. Scale bar is 20 µm. 6 Supplementary Figure S6 (A, B) Co-immunoprecipitation of Flag-tagged Axin (A) or Conductin (B) with EGFPtagged Amer1 or Amer1 deletion mutants after immunoprecipitation with anti-GFP antibodies in HEK293T cells. (C) Schematic representation of Amer1 deletion mutants and their ability to bind to the indicated interaction partners. For the Amer1 scheme see Supplementary Figure S3A. 7 Supplementary Figure S7 8 (A) Structure of the LRP6 intracellular domain highlighting the five PPPSPxS motifs (labelled A-E) and adjacent Ser/Thr clusters, which are represented by the grey and red boxes, respectively. The sequence of PPPSPxS motif A and the two flanking CK1 clusters (cluster 1 and 2) is shown above. Red amino acids represent CK1 phosphorylation sites. Residues Thr1479 and Ser1490 detected by phospho-specific antibodies are indicated. Schematic representation of LRP6 deletion mutants and their ability to bind to Amer1 as shown in (B) is indicated. LRP6 mutants have been described in Davidson et al, 2005. Scheme adapted from Davidson et al, 2009. (B) Co-immunoprecipitation of Flag-tagged LRP6 mutants as detailed in (A) with EGFPtagged Amer1 after immunoprecipitation with anti-GFP antibodies in HEK293T cells. (C) Overexpression of Axin does not promote the interaction between Amer1 and LRP6. HEK293T cells stably expressing VSVG-LRP6 were transfected with FlagAmer1 with or without YFP-Axin and anti-Flag immunprecipitations were performed. Supplementary Figure S8 (A, B) Amer1 interacts with CK1γ. (A) Flag-Amer1 co-immunoprecipitated with EYFPCK1γ, but not with a CK1γ mutant lacking the C-terminal membrane association motif (EYFP-CK1γΔC). (B) Co-immunoprecipitation of Flag-CK1γ with EGFP-Amer1 after immunoprecipitation with anti-GFP antibodies in HEK293T cells. 9 Supplementary Figure S9 Fluorescence micrographs of MCF-7 cells expressing the indicated EGFP-tagged fusion constructs. For the Amer1 scheme see Supplementary Figure S3A. Scale bar is 20 µm. Supplementary Figure S10 Knockdown of APC does not affect LRP6 phosphorylation at Ser1490. HEK293 cells stably expressing either an APC shRNA (293iAPC) or control shRNA (293control) (Schneikert and Behrens, 2006) were incubated with Wnt3A for 1 hour. Membrane fractions (M) and whole cell lysates (WCL) were analysed by Western blotting. 10 Supplementary Figure S11 Overexpression of Amer1 represses a TCF/β-catenin dependent transcriptional reporter. HEK293T cells stably expressing a β-catenin responsive firefly luciferase (Major et al, 2007) were transfected with RFP-Wnt3A, RFP-daLRP6 and EGFPAmer1 as indicated. Fold changes of luciferase activity are presented relative to control transfected cell. Error bars indicate standard deviations. 11 SUPPLEMENTARY MATERIALS AND METHODS Cell culture and transfection All cell lines were cultured in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal calf serum (FCS) and 1% penicillin/streptomycin (PAA Laboratories, Pasching, Austria) at 37°C in a humidified atmosphere of 10% CO2. Plasmid transfections were performed using either polyethylenimin (Sigma) for HEK293T and SW480 cells or TransIT-TKO (Mirus, Madison, WI, USA) for MCF-7 cells. siRNAs were transfected using Oligofectamine (Invitrogen) according to the manufacturer’s instructions. Plasmids and siRNAs The following plasmids have been described previously: pEGFP-Amer1, pcDNAFlag-Amer1 (Grohmann et al, 2007); mRFP-daLRP6, mRFP-Wnt3A, mYFP-Axin, mYFP-Conductin, mCFP-Axin, mCFP-Conductin, mCFP-Dvl2 (Krieghoff et al, 2006); pcDNA3.1-Flag, pcDNA-Flag-Conductin, GSK3β (Behrens et al, 1998). LRP6-EYFP, Fz8-EYFP, MESD, EYFP-CK1γ, EYFP-CK1γΔC, Flag-LRP6ΔE(1-4), FlagLRP6ΔE(1-4)Δ87, Flag-LRP6ΔE(1-4)Δ127, Flag-LRP6ΔE(1-4)Δ162 and FlagLDLRΔN-miniC were kindly provided by C. Niehrs, VSVG-LRP6, VSVG-LRP6m10 and pcDNA-Flag-Axin L396Q by X. He and pcDNA-Flag-Axin by A. Kikuchi. Deletion mutants of Amer1 were generated by restriction digests or PCR amplification. The sequences of the siRNA oligonucleotides are: siGFP, 5’-GCTACCTGTTCCATGGCCA-3’; siLuc, 5’-CTTACGCTGAGTACTTCGA-3’; siAmer1a, 5’-GGGAGTACCCGTGAACAAA-3’; siAmer1b, 5’-CCTCTATGCCCAAGCCAAA-3’; siAmer1-S1, 5’- 12 CCACCAGCTACTGAGAAAA-3’; siAmer1-S2, 5’-GGCCCAGGTTGTGGTGACA-3’; siPI4KIIα, 5’-GGATCATTGCTGTCTTCAA-3’ (Pan et al, 2008); siPI4KIIα-2, 5’GGAAGAGGACCTATATGAA-3’ (Pan et al, 2008). All siRNA oligonucleotides were purchased from Eurogentec. Preparation of protein lysates, immunoprecipitation and Western blotting For cytoplasmic or whole cell lysates cells were lysed for 10 minutes at 4°C in hypotonic buffer (25 mM Tris-HCl, pH 8, 1 mM EDTA, 10 mM NaF, 1 mM DTT and 1 mM PMSF) or Triton-X-100 buffer (20 mM Tris-HCl, pH 7.4, 150 mM NaCl, 5 mM EDTA, 1% Triton X-100, 10 mM NaF, 1 mM DTT and 1 mM PMSF), respectively. Lysates were cleared at 16,000 g for 10 minutes. For immunoprecipitation, lysates were incubated for 4 hours at 4°C with the appropriate antibody and Protein A/G PLUS agarose beads (Santa Cruz Biotechnology Inc.) or with Anti-FLAG M2 affinity gel beads (Sigma). Immunoprecipitates were collected, washed four times with low salt NET buffer (50 mM Tris-HCl, pH 8, 150 mM NaCl, 5 mM EDTA, 1% Triton-X-100, 10 mM NaF) and eluted with SDS sample buffer. For Western blotting (Lustig et al, 2002) proteins were visualized with a luminoimager (LAS-3000, Fuji) using Enhanced Chemiluminescence reagent (Perkin Elmer) and quantified using the AIDA image analyser software v. 3.52 (Raytest, Straubenhardt, Germany). Antibodies The mouse monoclonal antibody against Conductin has been described before (Lustig et al, 2002). Rabbit anti-pT1479 and rabbit anti-PI4KIIα were kindly provided by C. Niehrs and P. De Camilli, respectively. Commercial antibodies were obtained from Abcam (mouse anti-LRP6 (1C10)), Cell Signaling (rabbit anti-Axin1 (C76H11), rabbit anti-GSK3β (27C10), rabbit anti-LRP6 (C5C7), rabbit anti-pLRP6 (Ser1490)), 13 R&D Systems (sheep anti-CK1γ), Roche (mouse anti-GFP, mixture of clones 7.1 and 13.1), Santa Cruz Biotechnology Inc. (rabbit anti-β-catenin (H102)), Serotec (rat antiα-tubulin, clone YL1/2) and Sigma (rabbit anti-Flag, mouse anti-Flag, mouse anti-GST, rabbit anti-Pan-Cadherin). Secondary antibodies coupled to horseradish peroxidase or Cy3 were purchased from Jackson ImmunoResearch. RT-PCR Total cellular RNA was isolated with the RNeasy mini kit (Qiagen) and possible genomic contaminations were removed by treatment with DNase I. Single stranded cDNA was synthesized from 1 µg total cellular RNA using the AffinityScript QPCR cDNA Synthesis Kit (Stratagene) according to the manufacturer’s instructions. 14 SUPPLEMENTARY REFERENCES Behrens J, Jerchow BA, Wurtele M, Grimm J, Asbrand C, Wirtz R, Kuhl M, Wedlich D Birchmeier W (1998) Functional interaction of an axin homolog, conductin, with beta-catenin, APC, and GSK3beta. Science 280: 596-599 Davidson G, Shen J, Huang YL, Su Y, Karaulanov E, Bartscherer K, Hassler C, Stannek P, Boutros M, Niehrs C (2009) Cell cycle control of wnt receptor activation. Dev Cell 17: 788-799 Davidson G, Wu W, Shen J, Bilic J, Fenger U, Stannek P, Glinka A, Niehrs C (2005) Casein kinase 1 gamma couples Wnt receptor activation to cytoplasmic signal transduction. Nature 438: 867-872 Grohmann A, Tanneberger K, Alzner A, Schneikert J, Behrens J (2007) AMER1 regulates the distribution of the tumor suppressor APC between microtubules and the plasma membrane. J Cell Sci 120: 3738-3747 Jenkins ZA, van Kogelenberg M, Morgan T, Jeffs A, Fukuzawa R, Pearl E, Thaller C, Hing AV, Porteous ME, Garcia-Minaur S, Bohring A, Lacombe D, Stewart F, Fiskerstrand T, Bindoff L, Berland S, Ades LC, Tchan M, David A, Wilson LC et al (2009) Germline mutations in WTX cause a sclerosing skeletal dysplasia but do not predispose to tumorigenesis. Nat Genet 41: 95-100 Krieghoff E, Behrens J, Mayr B (2006) Nucleo-cytoplasmic distribution of {beta}catenin is regulated by retention. J Cell Sci 119: 1453-1463 Lustig B, Jerchow B, Sachs M, Weiler S, Pietsch T, Karsten U, van de Wetering M, Clevers H, Schlag PM, Birchmeier W, Behrens J (2002) Negative feedback loop of Wnt signaling through upregulation of conductin/axin2 in colorectal and liver tumors. Mol Cell Biol 22: 1184-1193 Major MB, Camp ND, Berndt JD, Yi X, Goldenberg SJ, Hubbert C, Biechele TL, Gingras AC, Zheng N, Maccoss MJ, Angers S, Moon RT (2007) Wilms tumor suppressor WTX negatively regulates WNT/beta-catenin signaling. Science 316: 1043-1046 Pan W, Choi SC, Wang H, Qin Y, Volpicelli-Daley L, Swan L, Lucast L, Khoo C, Zhang X, Li L, Abrams CS, Sokol SY, Wu D (2008) Wnt3a-mediated formation of phosphatidylinositol 4,5-bisphosphate regulates LRP6 phosphorylation. Science 321: 1350-1353 Schneikert J, Behrens J (2006) Truncated APC is required for cell proliferation and DNA replication. Int J Cancer 119: 74-79 15 Vítězslav Bryja, 2014    Attachments      #12      Bernatik  O,  Sri  Ganji  R,  Cervenka  I,  Polonio  T,  Schulte  G,  Bryja  V  (2011):  Sequential  activation  and  inactivation  of  Dishevelled  in  the  Wnt/‐catenin  pathway  by  casein  kinases. J. Biol. Chemistry 286: 10396‐10410.      Impact factor (2011): 4.773  Times cited (without autocitations, WoS, Feb 21st 2014): 12  Significance:  This  detailed  analysis  for  the  first  time  rigorously  described  the  role  of  individual Dvl domains in several independent phenotypes associated with CK1‐ controlled phosphorylation. Surprisingly, this analysis lead to discovery of distinct  processes  controlled  by  CK1  and  identified  intrinsic  negative  feedback  loop  controlled by CK1.  Contibution of the author/author´s team: Completely performed in my lab with the  exception of FRET assays.                Sequential Activation and Inactivation of Dishevelled in the Wnt/␤-Catenin Pathway by Casein Kinases□S Received for publication,July 29, 2010, and in revised form, January 7, 2011 Published, JBC Papers in Press,February 1, 2011, DOI 10.1074/jbc.M110.169870 Ondrej Bernatik‡ , Ranjani Sri Ganji‡ , Jacomijn P. Dijksterhuis§ , Peter Konik¶ , Igor Cervenka‡ , Tilman Polonio§1 , Pavel Krejci‡ʈ **2 , Gunnar Schulte§3 , and Vitezslav Bryja‡ʈ4 From the ‡ Institute of Experimental Biology, Masaryk University, 61137 Brno, Czech Republic, the § Section of Receptor Biology and Signaling, Department of Physiology and Pharmacology, Karolinska Institutet, SE-171 77 Stockholm, Sweden, ¶ Faculty of Science, University of South Bohemia, 37005 Ceske Budejovice, Czech Republic, the ʈ Department of Cytokinetics, Institute of Biophysics ASCR, 61265 Brno, Czech Republic, and **Medical Genetics Institute, Cedars-Sinai Medical Center, Los Angeles, California 90048 Dishevelled (Dvl) is a key component in the Wnt/␤-catenin signaling pathway. Dvl can multimerize to form dynamic protein aggregates, which are required for the activation of downstream signaling. Upon pathway activation by Wnts, Dvl becomes phosphorylated to yield phosphorylated and shifted (PS) Dvl. Both activation of Dvl in Wnt/␤-catenin signaling and Wnt-induced PS-Dvl formation are dependent on casein kinase 1 (CK1) ␦/⑀ activity. However, the overexpression of CK1 was shown to dissolve Dvl aggregates, and endogenous PS-Dvl forms irrespectiveofwhetherornottheactivatingWnttriggerstheWnt/ ␤-catenin pathway. Using a combination of gain-of-function, lossof-function, and domain mapping approaches, we attempted to solve this discrepancy regarding the role of CK1⑀ in Dvl biology. We analyzed mutual interaction of CK1␦/⑀ and two other Dvl kinases, CK2 and PAR1, in the Wnt/␤-catenin pathway. We show that CK2 acts as a constitutive kinase whose activity is required for the further action of CK1⑀. Furthermore, we demonstrate that the two consequences of CK1⑀ phosphorylation are separated both spatially and functionally; first, CK1⑀-mediated induction of TCF/ LEF-driven transcription (associated with dynamic recruitment of Axin1) is mediated via a PDZ-proline-rich region of Dvl. Second, CK1⑀-mediated formation of PS-Dvl is mediated by the Dvl3 C terminus.Furthermore,wedemonstratewithseveralmethodsthat PS-Dvl has decreased ability to polymerize with other Dvls and could, thus, act as the inactive signaling intermediate. We propose a multistep and multikinase model for Dvl activation in the Wnt/ ␤-catenin pathway that uncovers a built-in de-activation mechanism that is triggered by activating phosphorylation of Dvl by CK1␦/⑀. The Wnt signaling pathway is a well conserved signaling pathway necessary during embryonic development whose dysregulation in adult tissues is linked to cancer (1, 2). Wnts are secreted glycoproteins that can activate several downstream cascades including the Wnt/␤-catenin (canonical) pathway and other (noncanonical) pathways. These noncanonical pathways include planar cell polarity, the Wnt/Ca2ϩ pathway, and other less-defined signaling cascades. The Wnt/␤-catenin pathway initiates with the binding of Wnt to its receptor Frizzled (3) and the co-receptor LRP5/6. Upon LRP5/6 phosphorylation, the complex of Frizzled and LRP5/6 recruits Axin to the cell membrane. The ␤-catenin destruction complex, composed of Axin, adenomatosis polyposis coli, and glycogen synthase kinase 3␤, is subsequently disrupted. ␤-Catenin accumulates and is then translocated to the nucleus where it binds to the TCF55 /LEF family of transcription factors. The binding of ␤-catenin to TCF/LEF removes transcriptional repressors such as Groucho and initiates the transcription of TCF/LEF target genes (1, 4). The cytoplasmic protein Dishevelled (Dvl; Dvl1, Dvl2, Dvl3 in mammalian cells) is critically required for Wnt/␤-catenin signal transduction. Dvl acts as a scaffolding protein, which interacts with many proteins and serves as a key signaling intermediate between the WNT receptor Frizzled and downstream components in the Wnt/␤-catenin and non-canonical Wnt pathways (5–8). The current model of Wnt/␤-catenin signal transduction proposes Dvl as a core protein of dynamic protein assemblies called signalosomes (9). The signalosome hypothesis proposes that upon stimulation of the Wnt pathway, Dvl protein multimerizes via its DIX domains and forms a platform that recruits other proteins including Axin and that is required for the phosphorylation of LRP6 (9–11). When Dvl is overexpressed in mammalian cells, it is present in dynamic protein aggregates and is visible as Dvl punctae, which most likely represents Dvl multimers (12, 13). Although Dvl multimerization has yet to be convincingly demonstrated at the endogenous level, it is clear that endogenous Dvl becomes phosphorylated after the addition of Wnt. Wnt-induced activation of Dvl is visible as a Wnt-induced shift in the electrophoretic mobility of all three Dvl isoforms, forming the so called phosphorylated □S The on-line version of this article (available at http://www.jbc.org) contains supplemental Figs. 1–6. 1 Present address: Cellular Senescence Group, German Cancer Research Center (DKFZ), Im Neuenheimer Feld 242, 69120 Heidelberg, Germany. 2 Supported by Czech Science Foundation Grant 301/09/0587. 3 Supported by Knut and Alice Wallenberg Foundation Grant KAW2008.0149, Swedish Research Council Grants K2008-68P-20810-01-4 and K2008-68X- 20805-01-4, Swedish Cancer Society Grant CAN 2008/539, and The Swedish Foundation for International Cooperation in Research and Higher Education (STINT). 4 Supported by Ministry of Education, Youth, and Sports of the Czech Republic Grant MSM0021622430, Academy of Sciences of the Czech Republic Grants KJB501630801, AVOZ50040507, and AVOZ50040702, Czech Science Foundation Grants 204/09/H058, 204/09/0498, and 204/09/J030, and an EMBO Installation grant. To whom correspondence should be addressed: Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlarska 2, 61137 Brno, Czech Republic. Tel.: 420-549493291; Fax: 420-541211214; E-mail: bryja@sci.muni.cz. 5 The abbreviations used are: TCF, T-cell-specific transcription factor; PS, phosphorylated and shifted; CK1, casein kinase 1; MEF, mouse embryonal fibroblast; rm, recombinant mouse; x, Xenopus; Dvl, Dishevelled; LEF, lymphoid enhancer binding factor; TBCA, (E)-3-(2,3,4,5-tetrabromophenyl)acrylic acid; TBBt, tetrabromobenzotriazole; aa, amino acids. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 286, NO. 12, pp. 10396–10410, March 25, 2011 © 2011 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A. 10396 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom http://www.jbc.org/content/suppl/2011/02/01/M110.169870.DC1.html Supplemental Material can be found at: and shifted (PS) Dvl (14–16). It has been clearly demonstrated that both the activation of Dvl in the Wnt/␤-catenin pathway (17–21) and Wnt-induced PS-Dvl formation are dependent on casein kinase 1 (CK1) ␦/⑀ activity (14, 16). It is, thus, conceivable that CK1␦/⑀ phosphorylation of Dvl promotes the formation of signalosomes/Dvl aggregates, which is unique for the Wnt/␤catenin pathway. However, these predictions have not been supported by experimental data. First, overexpression of CK1 dissolved (and not promoted) Dvl aggregates in all tested cell types (14, 20, 22). Second, endogenous PS-Dvl formed irrespective of whether or not the activating Wnt triggered the Wnt/␤catenin pathway (14, 15, 23). In the present study we attempted to solve this obvious conundrum in our current understanding of the role of CK1⑀ in Dvl biology. We analyzed the mutual interaction of CK1␦/⑀, and two other Dvl kinases, CK2 and PAR1, in the Wnt/␤catenin pathway. Using a combination of gain-of-function, loss-of-function, and domain mapping approaches, we proposed a multistep and multikinase model for Dvl activation in the Wnt/␤-catenin pathway. This approach uncovered a built-in negative feedback loop triggered after activating phosphorylation of Dvl by CK1␦/⑀. MATERIALS AND METHODS Cell Culture, Transfection, and Treatments—WT mouse embryonal fibroblasts (MEFs) and HEK-293T cells were propagated in DMEM, 10% FCS, 2 mM L-glutamine, 50 units/ml penicillin, and 50 units/ml streptomycin. Cells (40,000–60,000 per well) were seeded in 24-well plates either directly (for biochemical analysis) or on sterile coverslips (for microscopy). The next day cells were transfected using polyethyleneimine in a stoichiometry of 2.5 ␮l per 1 ␮g of DNA. Cells were harvested for immunoblotting or immunocytochemistry 24 h after transfection. The following plasmids have been described previously: Dvl2-Myc (42), ca-␤-catenin (43), FLAG-Dvl3 deletion constructs (32), Xenopus (x) Dsh and xPAR1 constructs (25), xCK2 constructs (44), xCK1⑀ constructs (45), GST-DSH (36), and FLAG-Dvl1 (46). Treatment with the D4476, IC261, TBBt, or TBCA (Calbiochem) dissolved in DMSO was done in 24-well plates in the presence of 1 ␮l per well FuGENE 6 reagent to increase cell penetration. For analysis of cellular signaling, the cells were stimulated with mouse Wnt-3a or -5a (R&D Systems, Minneapolis, MN) for 2 h if not otherwise stated. Control stimulations were done with 0.1% BSA in PBS. RNA Interference—MEF or HEK-293T cells were transfected with siRNA using neofection according to the manufacturer’s instructions (Ambion). In brief, siRNAs (0.55 ␮l of 20 ␮M siRNA) were mixed with Lipofectamine 2000 (1.45 ␮l; Invitrogen) and Opti-MEM (48 ␮l; Invitrogen) and incubated for 20 min at room temperature. The transfection mixture (50 ␮l) was added to the 24-well plate and mixed with a suspension of freshly trypsinized cells (25,000 cells/well in 350 ␮l of complete media) resulting in a final concentration of 30 nM siRNA. When a combination of two different siRNAs was used, each siRNA was used at 30 nM, and the control siRNA was at 60 nM. The transfection was terminated after 5 h by changing the culture media. Cells were then treated with Wnt-3a or -5a inhibitors or transfected according to the scheme used. The used siRNAs were from Ambion (mCK1⑀ catalog no. 188528, mCK1␦ catalog no. 88388) and from Santa Cruz Biotechnology (hCK1⑀ sc 29914, hCK1␦ sc 29910, hCK2␣ sc-29918, mCK2␣ sc-29919, control siRNA sc-37007). Dual-Luciferase Assay, Western Blotting, Solubility Assay, Dephosphorylation Assay, and Immunoprecipitation—For the luciferase reporter assay, cells were transfected with 0.1 ␮g of Super8X TopFlash construct and 0.1 ␮g of Renilla luciferase construct per well in a 24-well plate 24 h after seeding. For the TopFlash assay, a Promega Dual-Luciferase assay kit was used according to the manufacturer’s instructions. Relative luciferase units were measured on a MLX luminometer (Dynex Technologies) and normalized to the Renilla luciferase expression 24 h post-transfection. Results were shown as the means with S.D. of at least three independent experiments. Immunoblotting and sample preparations were performed as previously described (14). For solubility assays, cells were seeded out on 12-well plates and transfected according to the scheme. Twenty-four hours post-transfection, cells were scraped into 200 ␮l of buffer containing 50 mM Tris, pH 8.5, 150 mM NaCl, 1 mM MgCl2, and protease inhibitors (Roche Applied Science, 11836145001) at 4 °C, and lysis was carried out for 15 min. Cells were then subjected to three freeze/thaw cycles. The lysates were centrifuged at 16,100 ϫ g, and the supernatant was collected. The pellet was then resuspended in 200 ␮l of lysis buffer. Both supernatant and pellet were prepared for Western blotting by adding 40 ␮l of 5ϫ Laemmli buffer, sonicated, and boiled at 95 °C for 5 min. The antibodies used include phosphoLRP6 (Ser-1490, #2568), Dvl3 (#3218), Dvl2 (#3224), Axin1 (#2087) from Cell Signaling Technologies, actin (C-11, sc-1615), Dvl3 (sc-8027), Dvl2 (sc-8026), CK1⑀ (sc-6471), and Myc (sc-40) from Santa Cruz Biotechnology, CK2␣ (#611610) from BD Transduction Laboratories, and FLAG M2 (F1804) from Sigma. For dephosphorylation assay cells were washed with buffer containing 100 mM Tris, pH 8.5, 150 mM NaCl, 0.2 mM EDTA. Then they were scraped into dephosphorylation buffer containing 50 mM Tris, pH 8.5, 150 mM NaCl, 1 mM MgCl2, and protease inhibitors (Roche Applied Science, 11836145001). Cells were lysed by a 22-gauge needle and then spun down at 16,100 ϫ g. Supernatant was divided into two parts. The pellet was resuspended in dephosphorylation buffer and also divided into two parts. One part of the supernatant and pellet was treated for 1 h by 80 units of alkaline phosphatase (Sigma P8361). 5ϫ Laemmli buffer was added to each sample, and samples were sonicated and boiled for 5 min. Immunoprecipitation was performed as previously described (22). The antibodies used for immunoprecipitation were anti-FLAG (F1804; Sigma), anti-CK1⑀ (sc-6471), antiDvl3 (sc-8027), anti-Dvl2 (sc-8026), and mouse IgG (sc-2025) (all Santa Cruz Biotechnology). GST Pulldown Assay—Production of recombinant GST-Dsh (36) was induced by adding 0.2 mM isopropyl 1-thio-␤-D-galactopyranoside and grown for 4 h at 37 °C. The bacteria were spun down at 4000 ϫ g at 4 °C for 10 min, and the pellet was resuspended in 10 ml of GST lysis buffer (50 mM Tris-Cl, pH 7.5, 150 mM NaCl, 5 mM MgCl2, protease inhibitors (Roche Applied Science)) and stored at Ϫ80 °C. Then the solution was thawed, Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10397 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom sonicated 3 ϫ 30 s, and spun down 15,000 ϫ g at 4 °C for 15 min, and the supernatant was then used for incubation with GST beads for 2 h on the rotator (100 ␮l of beads per 100 ml of original bacterial culture). After incubation beads were washed 3 times with 1 ml of GST lysis buffer and frozen at Ϫ80 °C as 25% slurry in GST lysis buffer ϩ 20% glycerol. HEK-293T cells were transfected according to the scheme, grown for 24 or 48 h, and lysed in 0.5% Nonidet P-40 lysis buffer (0.5% Nonidet P-40, 50 mM Tris, pH 7.4, 1 mM EDTA, 150 mM NaCl) with added protease inhibitors (Roche Applied Science), phosphatase inhibitors (Calbiochem), and 10 mM N-ethylmaleimide (Sigma). The lysate was spun down at 16,100 ϫ g at 4 °C for 10 min, and the supernatant was used for overnight incubation with 15 ␮l of solid GST beads containing GST recombinant proteins on the rotator. After the incubation samples were washed with 800 ␮l of 0.5% Nonidet P-40 lysis buffer, the GST beads were collected at 0.1 ϫ g at 4 °C for 1 min, the supernatant was aspirated; this washing was repeated 6 times. Proteins were eluted with 45 ␮l of 2ϫ Laemmli buffer. Immunocytochemistry—HEK-293 cells were seeded at ϳ2 ϫ 105 cells/well on collagen-coated coverslips in 24-well plates. Cells were then transfected the next day, and 24 h later they were fixed in fresh 4% paraformaldehyde, permeabilized with 0.05% Triton X-100, blocked with 3% BSA, 0,25% Triton, 0. 01% NaN3 for 1 h, and incubated overnight with primary antibodies. The next day coverslips were washed in PBS and incubated with secondary antibodies Cy2 and Cy3 (Jackson ImmunoResearch), washed with PBS, stained with DAPI (1:5000), and mounted on coverslips. Cells were visualized using a Zeiss LSM 710 confocal microscope. Two hundred positive cells per coverslip were analyzed and scored as described previously (14). Graphs show percentages of individual localization patterns. Fo¨rster Resonance Electron Transfer (FRET)—HEK-293 cells were seeded in a 24-well plate on sterile glass coverslips and grown overnight (62,500 cells per well). Cells were transfected with 0.1 ␮g of Dvl2-GFP, 0.1 ␮g of FLAG-Dvl3, and either 0.25 ␮g of pcDNA 3.1 or CK1⑀ using the calcium phosphate method. The next day, cells were fixed in 4% paraformaldehyde in PBS for 15 min. Indirect immunostaining was done after blocking using anti-FLAG M2 (Sigma, catalogue no. F1804) and donkey anti-mouse Cy3-conjugated secondary antibody (Jackson ImmunoResearch). To control for co-transfection efficiency of Dvl isoforms and CK1⑀, separate samples were also stained for CK1⑀ (Santa Cruz, catalogue no. sc-6471) and donkey anti-goat Cy-5-conjugated secondary antibody. Glass coverslips were mounted on microscopy slides using glycerol gelatin (Sigma). FRET between Dvl2-GFP and FLAG-Dvl3-Cy3 was performed on a Zeiss LSM710 confocal microscope and determined by photo acceptor bleaching employing 100% laser power of 561 nm diode laser for 20 s. Signal intensity was routinely reduced by 80–90%. Images were acquired before and after photo bleaching with excitation/emission range 488/493–545 nm (GFP) and 561/562–681 nm (Cy3). Quantification of GFP emission pre- and post-bleaching was determined with region of interest analysis in at least 15 individual cells per experiment and condition, employing the ZEN 2009 software from ZEISS. Data were normalized to the GFP intensity before bleaching. For statistical analysis, Student’s t test was performed. Mass Spectrometry (MS/MS) Analysis—Protein pellets from MEF cells (5 ϫ 15-cm-diameter confluent plates/sample) were stored at Ϫ80 °C until analysis. After thawing they were immunoprecipitated using 5 ␮g of anti-Dvl2 (sc-8026, Santa Cruz Biotechnology) and anti-Dvl3 (sc-8027, Santa Cruz Biotechnology) antibodies. Proteomic grade trypsin (Sigma) was added at a concentration of 100 ng/␮l per sample. For label-free quantitative analysis, yeast alcohol dehydrogenase (Sigma) was added as a standard protein, with final concentrations of 50 fmol/␮l per sample. After adding trypsin, samples were incubated at 37 °C for 12 h. Digested peptides were desalted using ZipTip C18 pipette tips (Millipore Corp.) according to the manufacturer’s protocol. MS analysis was performed on a NanoAcquity UPLC (Waters) on-line coupled to an electrospray ionization Q-Tof Premier (Waters) mass spectrometer. 1 ␮l of sample was diluted in water and loaded onto a 180-␮m ϫ 20-mm nanoAcquity UPLC Symmetry trap column (Waters) packed with 5-␮m BEH C-18 beads. After 1 min of trapping, peptides were eluted through a 75-␮m ϫ 150-mm nanoAcquity (Waters) analytical column packed with 1.7-␮m BEH C-18 beads at a flow rate of 400 nl/min using a gradient of 3–40% acetonitrile with 0.1% formic acid for 35 min at a temperature of 35 °C. Effluent was directly fed into the electrospray ionization source of the mass spectrometer. Raw data were acquired in data independent MSE Identity (Waters) mode. Precursor ion spectra were acquired with collision energy of 5 V, and fragment ion spectra were acquired with collision energy of 20–35-V ramp in alternating 1-s scans. Peptide spectra and fragment spectra were acquired with 2 and 5 ppm tolerance, respectively. Raw data were then subjected to a data base search using species specific Uniprot and NCBI mouse protein data base by the PLGS2.3 software (Waters). Acetyl N-terminal, deamidation N and Q, carbamidomethyl C, and oxidation M were set as variable modifications. Identification of three consecutive y- or b-ions was required for a positive peptide match. Protein quantification is based on peak intensities of at least three positively identified peptides belonging to one protein compared with peak intensities of the standard protein (Dvl2). Quantitative analysis was performed by the PLGS2.3 software (Waters). RESULTS CK2␣, PAR1, and CK1⑀ Promote Dvl-dependent Wnt/ ␤-Catenin Signaling via Distinct Mechanisms—Three Dvl kinases, CK1⑀, CK2␣, and PAR1, have been previously shown to play a positive role in Wnt/␤-catenin signaling (17, 24, 25). We addressed the question of the interplay between these three kinases and Dvl by coexpressing them alone (as wild type or kinase inactive variants) or in combination with FLAG-Dvl3 in HEK-293 cells. For CK2 overexpression, we co-expressed both kinase subunits (CK2␣ and CK2␤) if not mentioned otherwise. We used TCF/LEF reporter (TopFlash) (26) activity and electrophoretic migration of FLAG-Dvl3 as a readout (Fig. 1). FLAG-Dvl3 can be detected on Western blot as two bands that we call Dvl and P (phosphorylated)-Dvl. Only CK1⑀, but not CK2 or PAR-1, was able to promote the formation of novel, Casein Kinases in Dishevelled Biology 10398 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom - w t 8S/TA - w t 8S/TA 0 1 2 3 *** CK2/PAR1 XDsh TopFLASHfoldinduction FLAG CK1ε CK2α Myc - + - - + + + + + + + + + + + - - wt KI wt KI wt KI wt wt wtwt KI KI KI KI - - - - - - - wt KI - - wt KI - - - - - - - - - - wt KI - - wt KI FLAG CK1ε CK2α PAR1 myc 0 1 2 3 4 * TopFLASHfoldinduction - w t 8S/TA - w t 8S/TA 0 2 4 6 8 10 ns CK1εε XDsh TopFLASHfoldinduction FLAG CK1ε CK2α Myc - + - - + + + + + + + + + + + - - wt KI wt KI wt KI wt wtwt wt KI KI KI KI - - - - - - - wt KI - - wt KI - - - - - - - - - - wt KI - - wt KI FLAG CK2α CK1ε PAR1 myc 0.0 0.5 1.0 1.5 2.0 2.5 * TopFLASHfoldinduction FLAG CK1ε CK2α Myc - + - - + + + + + + + + + + + - - wt KI wt KI wt KI wt wtwt wt KI KI KI KI - - - - - - - wt KI - - wt KI - - - - - - - - - - wt KI - - wt KI FLAG PAR1 myc CK1ε CK2α 0 1 2 3 4 * TopFLASHfoldinduction A. B. C. D. E. FIGURE 1. A–C, cotransfection of CK1⑀, CK2␣/␤, and PAR1 kinases with FLAG-Dvl3. FLAG-Dvl3 (0.1 ␮g) plasmid was cotransfected with 0.4 ␮g of each kinase or a corresponding amount of pcDNA3 as control. Samples were analyzed for the activation of TCF/LEF-dependent transcription using the TopFlash reporter system (graphs summarize three independent experiments) and for the analysis of electrophoretic migration of FLAG-Dvl3. Western blot analysis of CK1⑀, CK2␣, and PAR1-Myc kinases provides a control for efficiency of transfection. All the measurements have been performed in three independent replicates. The arrow indicates PS-Dvl3, the filled arrowhead indicates P-Dvl3, and the open arrowhead indicates non-modified Dvl3. D and E, HEK-293 cells were transfected with wild type Xenopus Dsh (XDsh) or XDsh with eight serines in the b-region mutated to alanines (8S/TA) and corresponding kinases. XDsh-8S/A does not lack the TopFlash-inducing potential for CK1⑀ but fails to induce TopFlash to the same extent in CK2/PAR1-transfected cells. Data represent the means Ϯ S.D. *, p Ͻ 0.05; **, p Ͻ 0.01; ***, p Ͻ 0.001, one-way analysis of variance, Tukey post test. ns, not significant. Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10399 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom even more slowly migrating forms of Dvl3 labeled as PS-Dvl (see the Western blot panels in Fig. 1, A–C). Such changes in the electrophoretic mobility of FLAG-Dvl3 induced by CK1⑀ coexpression correlated with robust activation of the TopFlash reporter, which reflected activation of the Wnt/␤-catenin pathway (see the graphs in Fig. 1, A–C). CK2 was able to partially activate the TopFlash reporter, whereas the activation of the reporter by PAR1 was negligible. However, co-expression of PAR1 was sufficient to significantly increase the reporter activation induced solely by either CK1⑀ or CK2 (see Fig. 1, A and B). The most prominent TopFlash induction was in the sample in which CK1⑀ and PAR1 were coexpressed together (highest bar in Fig. 1A). PAR1 in combination with CK2 was able to shift Dvl3 electrophoretic migration from Dvl to P-Dvl (Fig. 1C, lane 9) but never to PS-Dvl (Fig. 1, A–C, all samples with CK1⑀). This difference suggested that the effects of CK2/PAR1 were distinct from the effects of CK1⑀. As CK2 and PAR1 kinases were shown to mainly phosphorylate Dvl in the basic region/PDZ region (24, 25), this observation would suggest that they cooperate or were redundant in the phosphorylation of Dvl in the Wnt/␤catenin pathway. In line with these observations, when the Ser/ Thr cluster in the basic region preceding PDZ domain of Dishevelled was mutated to alanines (25), it had a reduced capacity to activate the TopFlash reporter after phosphorylation by CK2, PAR1, or the combination of both (Fig. 1D, supplemental Fig. 1). It was, however, fully active after phosphorylation by CK1⑀ (Fig. 1E). These results demonstrate that CK1⑀ and CK2/PAR1 were both capable of activating the Wnt/␤catenin pathway in a Dvl-dependent manner, but the mechanisms were distinct. Efficient activation of downstream signaling accompanied by the formation of PS-Dvl can only be achieved by the phosphorylation by CK1⑀. Because we have not observed any functional differences between CK2 and PAR1, we used the combination of the two kinases in the further work. Down-regulation of CK1␦/⑀ but Not of CK2-diminished Formation of PS-Dvl in MEF Cells after Wnt Stimulation—Gainof-function experiments suggested that CK1 and CK2/PAR1 phosphorylated Dvl differently. To test the requirement of CK1 and CK2 for endogenous PS-Dvl formation, we used a loss-offunction approach in MEFs. MEFs represent a very suitable system for the analysis of Dvl biology because they respond efficiently to Wnt stimulation by the formation of PS-Dvl. We have used several antibodies against endogenous Dvl2 and Dvl3 that show slightly different sensitivity against individual phospho-variants of Dvl. We down-regulated CK1⑀ and -␦, which are two isoforms previously shown to be redundant and required for PS-Dvl formation (14), or CK2␣ using siRNA-mediated knockdown in MEF cells. Stimulation with recombinant mouse (rm) Wnt3a (50 ng/ml) or rmWnt5a (100 ng/ml) for 2 h efficiently promoted PS-Dvl formation in MEF cells (Fig. 2A). Consistent with the earlier findings, the knockdown of CK1␦/⑀ blocked the formation of PS-Dvl after Wnt3a or Wnt5a treatment (Fig. 2A). However, we failed to see any effect of CK2␣ knockdown on the Wnt-induced PS-Dvl formation. This experiment suggested that CK1␦/⑀ was indeed necessary for the dynamic Wnt-induced formation of PS-Dvl and that CK2 was not. To confirm the knockdown data, we pretreated MEF cells with CK1 (100 ␮M D4476 and 50 ␮M IC261)- and CK2 (50 ␮M TBCA, 100 ␮M TBBt)-specific inhibitors for 30 min and subsequently with rmWnt3a (50 ng/ml) or rmWnt5a (100 ng/ml) for 2 h. As we show in Fig. 2B (D4476 and TBCA) and supplemental Fig. 2 (IC261 and TBBt), CK1-specific inhibitors were able to block the Wnt-induced formation of PSDvl. In contrast, the CK2 inhibitors failed to block or diminish the formation of the Wnt-induced PS-Dvl band. On the other hand, they were able to generally decrease Dvl phosphorylation, which also resulted in the appearance of a thirdfastest migrating band (Dvl3, antibody CST#3218). The CK2␣ activity toward Dvl has been shown to be constitutive (24, 27). Thus, we propose that the CK2 inhibitor-induced third band was a non-modified Dvl3, as the molecular mass corresponded to the in silico-determined mass of Dvl3. Interestingly, treatment of cell lysates with alkaline phosphatase completely removed PS-Dvl but only partially promoted appearance of Dvl (supplemental Fig. 2B). This suggests that CK2-induced modification of Dvl is not solely phosphorylation but may be more complex and involve other modifications such as mono-ubiquitination. It is also worth noting that CK1 inhibition decreased Wnt3a-induced phosphorylation of LRP6 at Ser-1490, which is known to be a Dvl-dependent process (9, 28). These loss-of-function experiments suggest that endogenous Dvl is constitutively modified (but not only phosphorylated) in a CK2-dependent manner (P-Dvl) and after Wnt stimulation is further phosphorylated by CK1␦/⑀ to form PS-Dvl. To demonstrate gradual modification of Dvl3 by CK2 (to P-Dvl) and CK1␦/⑀ (to PS-Dvl), we first eliminated endogenous CK1⑀ by siRNA and overexpressed FLAG-Dvl3 with Xenopus CK2 and CK1⑀, which were not targeted by siRNA. As we show in Fig. 2C, under these conditions (right side of the panel) exogenous CK2 clearly shifts Dvl to P-Dvl (filled arrowhead), and CK1⑀ to PS-Dvl (full arrow). The combination of results shown in Figs. 1 and 2, thus, suggests that CK2/PAR1 and CK1⑀ may phosphorylate Dvl sequentially with CK2 activity being constitutive and CK1⑀ activity being induced by Wnts. Both CK1 and CK2 Kinase Activities Are Irreplaceably Required for Wnt/␤-Catenin Activation via Dvl3—We hypothesized that phosphorylations by CK2 and CK1 represent distinct, although sequential steps in the chain of events required for Dvl activation in the Wnt/␤-catenin pathway. To test this hypothesis, we induced TCF/LEF transcription by co-expressing FLAG-Dvl3 and CK1⑀ or FLAG-Dvl3 and CK2/PAR1, and we tested the effects of CK1 and CK2 inhibition, respectively. As we show in Fig. 3, A and B, the CK1⑀ inhibitors D4476 and IC261 were able to diminish TopFlash induction by both FLAG-Dvl3ϩCK1⑀ and FLAG-Dvl3ϩCK2/PAR1. Strikingly, the two CK2 inhibitors TBCA and TBBt showed very similar effects and blocked a Dvl3-dependent increase in TCF/LEFdriven transcription induced by both CK1⑀ and CK2/PAR1. Importantly, inhibition of casein kinases did not show a consistently negative effect on TCF/LEF-mediated transcription induced by constitutively active (S33A) ␤-catenin (Fig. 3C). This result was also not observed in the absence of FLAG-Dvl3, which demonstrates that (i) the inhibitors did not interfere with Casein Kinases in Dishevelled Biology 10400 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom siRNA Ctrl siRNA CK1ε/δ siRNA CK2α 0.1% BSA in PBS Wnt-3a (50ng) Wnt-5a (100ng) + - - + - - + - - - - - - -+ + + + - - + - - +- pLrp6 (S1490) Dvl2(CST#3224) Dvl3(sc-8027) CK2α CK1ε actin pLrp6 (S1490) Dvl2(CST#3224) Dvl3(sc-8027) actin 0.1% BSA in PBS Wnt-3a (50ng) Wnt-5a (100ng) Ctrl D4476 TBCA + - - + - - + - - - - - - -+ + + + - - + - - +- - Dvl3(CST#3218) Flag-Dvl3 CK1ε CK2 siCtrl siCK1ε - + + + + + - - - --- - + + + + + - - - --- Flag CK2α CK1ε A. B. C. FIGURE2.A,MEFcellsweretransfectedwithsiRNAforCK1⑀and-␦orCK2␣.24hpost-transfectioncellsweretreatedwithrecombinantmouseWnt3a(50ng/ml) or Wnt5a (100 ng/ml). Cell lysates were harvested 2 h after treatment. After Wnt 3a/5a treatment, CK1⑀/␦ siRNA diminishes the formation of PS-Dvl, whereas CK2␣siRNAdoesnot.WesternblotanalysisforCK1⑀andCK2␣demonstratesknockdownefficiency;phosphorylationofLrp6wasusedtodeterminetheWnt3a activity. B, MEF cells were pretreated with the CK1 (D4476 100 ␮M) and CK2 (TBCA 100 ␮M) inhibitors. Cells were treated with rmWnt3a (50 ng/ml) or Wnt5a (100 ng/ml) after 30 min of pretreatment. Cell lysates were harvested 2 h after treatment. The CK1 inhibitor D4476 was able to diminish PS-Dvl formation after Wnt treatment, whereas CK2 inhibitors were not able to do so. Unlike CK1 inhibitors, CK2 inhibitors were able to cause the formation of lowest molecular weight form of Dvl3. C, HEK-293 cells were transfected with siRNA against CK1⑀ and after 24 h transfected with corresponding constructs. In the absence of endogenous CK1⑀, CK2 clearly promotes P-Dvl, whereas PS-Dvl formation is only induced by CK1⑀. The arrow indicates PS-Dvl3, the filled arrowhead indicates P-Dvl3, and the open arrowhead indicates non-modified Dvl3. Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10401 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom the signaling downstream of ␤-catenin, and (ii) the observed effects were Dvl3-dependent. To demonstrate how endogenous CK1⑀ and CK2␣ contributed to the activation of the TCF/LEF transcription induced by overexpressed kinases, we down-regulated endogenous CK1⑀ and CK2␣ in HEK-293 cells using siRNAs. Used siRNAs do not target exogenous xCK1⑀ or xCK2/xPAR1 kinases, which were together with FLAG-Dvl3 overexpressed in cells depleted of CK1⑀ or CK2␣ (Fig. 3, E and F). Knockdown of CK1⑀ and CK2␣ decreased TopFlash induction in FLAG-Dvl3ϩxCK1⑀, which suggested that endogenous CK2␣ activity was required for the action of overexpressed CK1⑀. Furthermore, down-regulation of endogenous CK1⑀ was able to diminish the xCK2/xPAR1-induced TopFlash. CK2␣ siRNA did not show any effect in this assay, which suggests that xCK2␣ was able to fully rescue the lack of endogenous CK2␣ kinase. These results suggested that the phosphorylation of Dvl3 by CK1 and CK2 represented individual steps in the sequence of events. Accordingly, the activation of the Wnt pathway and inter- ctrl D vl3 D vl3+C K 2 ctrl D vl3 D vl3+C K 2 ctrl D vl3 D vl3+C K 2 0.0 0.5 1.0 1.5 siCtrl siCK1ε/δ siCK2α ** ns TopFLASHratio DMSO D4476 IC261 TBBt TBCA DMSO D4476 IC261 TBBt TBCA 0.0 0.5 1.0 1.5 β-cateninCtrl TopFLASHratio DMSO D4476 IC261 TBBt TBCA DMSO D4476 IC261 TBBt TBCA 0.0 0.2 0.4 0.6 Ctrl Dvl3 TopFLASHratio ctrl D vl3 ε D vl3+C K 1 ctrl D vl3 ε D vl3+C K 1 ctrl D vl3 ε D vl3+C K 1 0 1 2 3 siCtrl siCK1ε/δ siCK2α *** ** TopFLASHratio DMSO D4476 IC261 TBBt TBCA DMSO D4476 IC261 TBBt TBCA 0 1 2 3 4 5 Ctrl Dvl3+CK1ε ** TopFLASHratio DMSO D4476 IC261 TBBt TBCA DMSO D4476 IC261 TBBt TBCA 0 1 2 3 4 5 Ctrl Dvl3+CK2 *** TopFLASHratio A. B. C. D. F.E. FIGURE 3. A–D, HEK-293 cells were transfected with the indicated constructs and TopFlash reporter. Twenty-four hours post-transfection, cells were treated with CK1 (D4476 100 ␮M, IC261 50 ␮M) and CK2 (TBCA 50 ␮M, TBBt 100 ␮M) inhibitors. Both CK1 and CK2 inhibitors were able to significantly block TCF/LEFdriven transcription in both FLAG-Dvl3 ϩ CK1⑀ (A)- or FLAG-Dvl3ϩCK2/PAR1 (B)-transfected cells. The inhibitors did not show any effect in cells transfected with constitutively active ␤-catenin (C) or in the absence of overexpressed Dvl3 (D). E and F, HEK-293 cells were transfected with siRNA against CK1⑀ or CK2␣. After 24 h they were transfected with the corresponding constructs and TopFlash ϩ Renilla constructs. siRNA against CK1⑀ and CK2␣ were able to diminish DvlϩCK1⑀-induced TopFlash, whereas DvlϩCK2/PAR1-induced TopFlash could not be diminished by the siRNA. Data represent the means Ϯ S.D. **, p Ͻ 0.01; ***, p Ͻ 0.001; one-way analysis of variance, Tukey post test. ns, not significant. Casein Kinases in Dishevelled Biology 10402 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom ruption of any of these steps completely interfered with the ability of Dvl3 to promote Wnt/␤-catenin signaling. Both CK1⑀ and CK2 Promote Even Intracellular Distribution of Dvl—Dvl was usually found to be present in dynamic multiprotein aggregates (12, 13) called Dvl dots or puncta. Based on the cellular context, overexpressed Dvl2 or Dvl3 was usually present either in dots or was evenly distributed (Dvl2 exampled in Fig. 4A). CK1⑀ has been shown earlier to strongly promote even localization of Dvl (14, 20). Based on these findings, in the next step we tested the role of CK2 on Dvl distribution. Dvl2Myc was overexpressed with the indicated kinases in HEK-293 cells, and the cells were scored for their pattern of subcellular localization of Dvl2. As we show in Fig. 4B, CK1⑀ promoted even localization of Dvl, as previously reported. Interestingly, overexpression of CK2 showed very similar, albeit somewhat weaker, effects than CK1⑀. These effects are not due to reduced levels of Dvl2-Myc after kinase coexpression (inset in Fig. 4A). Importantly, CK1 inhibition by D4476 and CK2 inhibition by TBBt were both able to reduce the effects of either CK1⑀ or CK2 (Fig. 4C). These results suggested that the even distribution of Dvl was similar to the TCF/LEF-dependent transcription promoted by a mechanism that was non-redundantly dependent on both CK2 and CK1⑀. Analysis of Dvl Domains Required for the Effects of CK1⑀ and CK2—Dvl proteins have three very well conserved domains (see the scheme in Fig. 5): the N-terminal DIX domain, the PDZ domain, and the DEP domain. Apart from these main domains, there are two well conserved stretches of amino acids: a basic region (b) preceding the PDZ domain and a proline-rich cluster (Pro) between the PDZ and DEP domains. The DIX domain has been shown to function mainly in the canonical Wnt/␤-catenin pathway (29), whereas the function of the DEP domain has been mostly associated with noncanonical pathways (30, 31). The PDZ domain serves as a scaffold for many cooperating proteins and has been shown to be necessary in both pathways. To test the roles of and mutual relationship between CK1⑀ and CK2 in Dvl biology, we analyzed a set of diverse human Dvl3 mutants. We took advantage of the previously described deletion ctrl D 4476 TB B t ctrl D 4476 TB B t ctrl D 4476 TB B t 0.0 0.2 0.4 0.6 0.8 even small puncta large puncta ctrl CK1εε CK2 DistributionpatternofDvl2-Myc * ns * ctrl ε C K 1 C K 2 0.0 0.2 0.4 0.6 0.8 1.0 even small puncta large puncta DistributionpatternofDvl2-Myc * * + + +Dvl2-myc - + -CK1ε - - +CK2αβ A. B. C. FIGURE 4. HEK-293 cells were transfected with the indicated constructs and were then subjected to immunocytochemistry. A, shown is a typical localization pattern of Dvl2-Myc, which was scored in the experimental samples. The inset shows levels and electrophoretic migration of Dvl2-Myc in the analyzed cells. B, both kinases were able to change Dvl intracellular distribution from dotted to even. This effect can be reduced by treatment with CK1 (D4476) and CK2 (TBBt) inhibitors (C). Quantification is based on three independent replicates. *. p Ͻ 0.05, one-way analysis of variance (cells with even distribution), Tukey post test. ns, not significant. Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10403 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom mutants of Dvl3 (32) and analyzed the function of individual domains for Dvl3 behavior, which was modified by CK1 or CK2/PAR1. The parameters examined after CK1⑀ and in some cases after CK2 overexpression included (i) the ability to bind CK1⑀ (endogenous and overexpressed), endogenous CK2␣, and endogenous Axin1, (ii) the ability to promote TCF/LEF-dependent transcription (see Fig. 3), (iii) intracellular localization of Dvl3 (see Fig. 4), and (iv) electrophoretic mobility shift of Dvl3 (see Fig. 1). As we show in the protein-protein interaction analysis (Fig. 5), full-length Dvl3 was able to bind both endogenous and overexpressed CK1⑀, endogenous CK2␣, and upon CK1⑀ overexpression, full-length Dvl3 also bound endogenous Axin1. Moreover, full-length Dvl3 acted as a scaffold, which was required for the co-immunoprecipitation of CK1⑀ with endogenous CK2␣ and Axin1. Analysis of mutants showed that CK1⑀ could be co-immunoprecipitated with at least two distinct domains of Dvl3. For high affinity interactions, which were indicated by the presence of endogenous CK1⑀ in the Dvl3 pulldown assay, only a DIX domain was dispensable (see aa 82–716 mutant). However, overexpressed CK1⑀ could be also found in association with all other mutants tested, with the exception of the extreme C terminus (mutant aa 496–716). Our results confirm previous findings that map the binding domains of CK1 to the DEP domain or its close proximity (compare aa 332–716 and 496–716 mutants and Kishida et al. (19)) and to the basic region preceding the PDZ domain (see mutant aa 1–246 and Klein et al. (33) and Klimowski et al. (34)). Endogenous CK2␣ could be co-immunoprecipitated with all the mutants of Dvl3, which contained a basic region between the DIX and PDZ domains,. This finding was in agreement with our data in Fig. 1D and with a previously published report (24). Interestingly, the binding of CK2 to Dvl3 was promoted by the co-expression of CK1⑀ (see e.g. aa 1–422 and 1–495 mutants). In line with this observation, CK1⑀ could be precipitated with endogenous CK2 in a Dvl3-dependent manner only when the basic region/PDZ (CK2 binding site) and the proline-rich-DEP domain (CK1 binding site) were present (WT, aa 80–716 and 1–495 mutants). We also, for the first time, showed that CK1⑀ overexpression could induce the recruitment of endogenous Axin1 to Dvl3 (Fig. 5, panel WB: Axin1), which is very relevant for the inactivation of the destruction complex and for further downstream ␤-catenindependent signaling. Axin1 bound only Dvl3 mutants containing both a DIX domain, which was previously reported as necessary for binding to Axin1 (29), and a PDZ-Pro-rich FIGURE 5. HEK-293 cells were cotransfected with the indicated combinations of FLAG-Dvl3 mutants and CK1⑀. The amino acids present in the mutants and the domain structure is schematized (b, basic region; Pro, proline-rich region). Two days later cells were lysed and subjected to immunoprecipitation (IP) with the corresponding antibodies (FLAG M2, CK1⑀). The presence of Dvl3, CK1⑀, CK2␣, and Axin1 in the pull-down assay was analyzed by Western blotting (WB). FLAG-Dvl3 constructs used in this experiment are schematized on the left hand side of the figure. Casein Kinases in Dishevelled Biology 10404 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom region, which probably contains crucial CK1⑀ phosphorylation sites required for Axin1 recruitment. In the next step we tested the functional properties of Dvl3 mutants with respect to the effects of CK1 and CK2. Dvl3 and its mutants were transfected into HEK-293T cells and tested for their ability to induce TCF/LEF-dependent transcription. As we show in Fig. 6A, the ability of Dvl3 truncations to induce transcription was very weak, with only 2-fold induction of TopFlash by WT and aa 1–422 and 1–495 mutants. There was no induction by mutants lacking the DIX domain, but there was approximately an 8-fold induction by mutant aa 1–246, which contained the DIX domain and a basic region. Expression of CK1⑀ alone increased the TCF/LEF reporter ϳ8-fold in this experimental setup (Fig. 6B), which we believe represented the effects of CK1⑀ on endogenous Dvl or on other endogenous components of the Wnt pathway. CK1⑀ together with DIXdeficient mutants (and also mutant aa 1–246) did not activate TopFlash above this level. In contrast, the activation capacity of both mutants, which contain a DIX, PDZ, and proline-rich region (aa 1–422 and 1–495) but lacked the C terminus, was even higher than the full-length Dvl after phosphorylation by CK1⑀ (Fig. 6B). This effect does not seem to be unique to Dvl3 because the Dvl1 mutant lacking the C terminus (Dvl1 aa 1–502) behaves very similarly (supplemental Fig. 3A). CK2/ PAR1 increased TopFlash activation to a lesser extent than CK1⑀, and the activity of none of the mutants, with the exception of mutant aa 1–495, which lacked only the C terminus, was potentiated more than 3-fold (Fig. 6C). Individual deletion mutants intrinsically differ in other basic parameters than in its ability to promote the Wnt pathway. The presence of the DIX domain, which was shown to be required for polymerization of Dvl (11), strongly determined the solubility and localization of Dvl. Mutants containing the DIX domain were insoluble when mildly lysed by freezing/thawing (supplemental Fig. 3B) and formed large aggregates when they were visualized by immunocytochemistry (Fig. 7). On the other hand, mutants lacking the DIX domain were more soluble (supplemental Fig. 3B) and were more evenly localized (Fig. 7). Our data were also in good agreement with the previous report, which mapped the nuclear export signal C-terminal from the DEP domain (35) because the normally cytoplasmic Dvl3 was also found in the nucleus after deletion of the C terminus (supplemental Fig. 4A). As we demonstrated above, CK1⑀ and CK2/ PAR1 changed the distribution of full-length Dvl3 from punctate to uniform (Fig. 4). The positive effects of CK1⑀ and CK2/ PAR1 on even localization of Dvl3 were conserved in the aa 82–716 and 332–716 mutants (Fig. 7, supplemental Fig. 4B). Under basal conditions, these two mutants were localized more uniformly than WT, and because they lacked their own DIX domain, we speculate that they aggregated with endogenous Dvl isoforms. Importantly, the distribution pattern of DIX-containing mutants, which also lacked the N terminus, was not modulated by CK1⑀ (Fig. 7) or CK2/PAR1 (not shown). The ability of CK1⑀ to homogenize the localization of Dvl3 correlated well with its ability to promote the efficient formation of hyperphosphorylated Dvl3 (probably corresponding to PSDvl3). This form in the overexpressed FLAG-Dvl3 was visible as the third, most slowly migrating band, induced only in the conditions where CK1 was co-expressed (Fig. 5, or in more detail, supplemental Fig. 5). CK1⑀ and CK2 also caused the formation of a shifted form of Dvl3 in other mutants with the exception of aa 496–716 (C terminus). However, in these cases it promoted predominantly, although not exclusively, the ratio of P-Dvl/Dvl bands, which were also present after Dvl3 overexpression without the kinase (supplemental Fig. 5). ctrl D vl3 82-716332-716496-716 1-246 1-422 1-495 0.00 0.02 0.04 0.06 0.08 TopFLASHratio 0 20 40 60 80 ctrl Dvl3 82-716 332-716 496-716 1-246 1-422 1-495 Ck1ε - + - + - + - + - + - + - + - + TopFLASHratio 0 5 10 15 20 TopFLASHratio ctrl Dvl3 82-716 332-716 496-716 1-246 1-422 1-495 CK2/PAR1 - + - + - + - + - + - + - + - + A. B. C. FIGURE 6. HEK-293 cells were transfected with the indicated constructs and a TopFlash reporter. A, the ability of FLAG-Dvl3 constructs to promote TCF/LEF-dependent transcription is shown. B and C, FLAG-Dvl3 constructs were co-transfected with CK1⑀ (B) or CK2/PAR1 (C). The graphs show -fold change increase in the TopFlash reporter for each particular construct induced by the co-expressed kinase. Graphs summarize data from three independent experiments (means Ϯ S.D.). Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10405 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom All the results regarding the deletion mutants of Dvl3 are summarized for clarity in Table 1. Based on the analysis of deletion mutants, we concluded the following previously unpublished information; (i) the roles of CK1⑀ and CK2 were distinct and involved different regions of Dishevelled, (ii) the C terminus showed an inhibitory function and was required for CK1⑀-induced change in the subcellular localization from punctate to even, (iii) the PDZ and Pro-rich rich region FIGURE 7. HEK-293 cells were transfected with corresponding constructs and then subjected to immunocytochemistry. Intracellular distribution of Dvl was observed and quantified 24 h after transfection. Typical subcellular distribution of Dvl3 mutants and quantitative analysis of distribution pattern (200 cells in each of three independent experiments) is shown. wt, wild type; ki, kinase inactive. Casein Kinases in Dishevelled Biology 10406 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom were required for CK1⑀-induced activation of Dvl in the Wnt/␤-catenin pathway. CK1⑀-phosphorylated PS-Dvl Has Decreased Ability to Polymerize with Other Dvls—The findings above suggested that CK1⑀ has apart from its clear positive role in Dvl/␤-catenin signaling also a subsequent negative role, which is at least in part mediated by formation of PS-Dvl and the Dvl-C terminus. Because CK1⑀ has primarily a strong positive role, it is not possible to analyze the negative consequences of CK1⑀ phosphorylation by looking globally at downstream functional readouts. It is more likely that if indeed CK1⑀ affects Dvl signaling negatively, it will interfere with the properties of Dvl required for efficient downstream signaling in the ␤-catenin pathway. A suitable measure of this function is the ability of Dvls to polymerize via their DIX domains, which was shown to be crucial for Dvl function. In addition, polymerization-deficient mutants of Dvl were shown to be inactive (9, 11). To test the ability of PS-Dvl phosphorylated by CK1⑀ to bind other Dvls, we have performed GST pulldown assays with GSTDvl (36) as bait. As we show in Fig. 8A, GST-Dvl is able to efficiently pull down FLAG-tagged Dvl3. Importantly, P-Dvl binds to GST-Dvl most efficiently, whereas non-modified Dvl or PS-Dvl triggered by CK1⑀ show much lower affinity to bacterially produced GST-Dvl (see Fig. 8B for quantification). Results of GST pulldown might be affected by protein modifications appearing in the cell extract after lysis. To avoid possible lysis artifacts, we performed FRET experiments in HEK- 293 cells expressing Dvl2-GFP and FLAG-Dvl3 with and without overexpression of CK1⑀. After fixation and indirect immunostaining of FLAG-Dvl3 with a Cy3-coupled secondary antibody, the electron acceptor (Cy3) was photobleached. Due to the destruction of the electron acceptor, FRET does not occur, and emission by the electron donor (GFP) increases compared with the prebleach situation when the two FRET partners are in close proximity to each other. In fact, FRET was measurable after photo acceptor bleaching with 115.8 Ϯ 1.2% (mean Ϯ S.E., n ϭ 30) in Dvl2-GFP- and FLAG-Dvl3-expressing cells, which showed a predominant punctate distribution of the two Dvl isoforms. With CK1⑀ overexpression, Dvl2-GFP and FLAG-Dvl3 showed a more even distribution as reported previously, and the GFP emission after photobleaching remained unchanged compared with the GFP emission before bleaching (107.0 Ϯ 1.2%; mean Ϯ S.E., n ϭ 30) (Fig. 8C). Typical appearance of cells used for FRET is shown in supplemental Fig. 6. We concluded that reduced FRET indicates the dissociation of the two Dvl isoforms upon CK1⑀ overexpression. The results above suggested that persistent phosphorylation of Dvl by CK1⑀, which leads to the formation of PS-Dvl, causes decreased ability of PS-Dvl to multimerize. However, PS-Dvl is also formed after Wnt3a stimulation (see Fig. 2). To exclude pos- sibleartifactsofoverexpression,wedecidedtotestwhetherendogenous PS-Dvl has also decreased ability to polymerize. Therefore, westimulatedMEFcellswithWnt3a,immunoprecipitatedendogenous Dvl2, and analyzed the amount of Dvl3 in the pulldown assay by Western blotting. The antibody used for precipitation does not cross-react with endogenous Dvl1 and Dvl3 (Ref. 37 and data not shown). As we show in Fig. 8D, the amount of Dvl3 in complexwithDvl2decreasedafter2hofWnt3astimulation,when PS-Dvl represented a dominant Dvl form. To further confirm this finding and to avoid possible cross-reactivity of used antibodies, we have repeated the co-immunoprecipitation experiment with endogenous Dvl2 and Dvl3 as bait and TABLE 1 Summary of deletion mutants of Dvl3 Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10407 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom analyzed the composition of the complexes by MS/MS. First, we have noticed that the number of samples where Dvl2 co-immunoprecipitated with Dvl3 and Dvl3 co-immunoprecipitated with Dvl2 dramatically decreased after 3 h of Wnt3a stimulation (Fig. 8E).Pleasenotethatonlysampleswherethebaitwasdetectedwith a significant score were considered for the analysis. We conclude from this non-quantitative analysis that the amount of the respective Dvl that co-immunoprecipitated with the bait after Wnt3a treatment was generally lower and more frequently below the detection limit of MS/MS technology. We could also perform MS/MS-based relative quantification of individual Dvls in a few Wnt3a-treated samples where Dvl2 was used as bait and where Dvl3 was detected in sufficient peptide quality and in sufficient number of technical replicates (see “Materials and Methods” for details). The amount of Dvl2 (bait) was set to 10 arbitrary units, and Dvl3 in the samples was quantified relatively to Dvl2. The results of this analysis showed that the amount of endogenous Dvl3 in complex with Dvl2 is indeed lower after Wnt3a stimulation (Fig. 8F). These results are in agreement with observations obtained by Western blot detection and by the non-quantitative MS/MS analysis. In summary, our findings shown in Fig. 8 demonstrate that PS-Dvl has decreased ability to bind other Dvls and suggest that PS-Dvl represents a negative signaling intermediate with decreased ability to polymerize. DISCUSSION In this study we attempted to answer two main questions related to the Wnt/␤-catenin signaling pathway; that is, how control Wnt3a 0 1 2 3 4 RelativeamountofDvl3 incomplexwithDvl2 0 1 2 3 pulldown/inputratio PS-Dvl3P-Dvl3Dvl3 Detected Dvl2 Detected Dvl3 0.0 0.2 0.4 0.6 0.8 control Wnt3a N=10 N=8 N=8 N=6 IP Dvl3 sc8027 IP Dvl2 sc8026 proportionofsamples whereDvlwasdetected IP TCL IgG Dvl2 sc8026 - +- - +Wnt-3a Dvl2 sc8026 Dvl3 CST#3218 Dvl3 sc8027 pcDNA3 CK1ε 0 20 40 60 80 100 105 110 115 *** FRET(Dvl2-Dvl3) + + +FLAG- Dvl3 + - + -CK1ε + GST GST-Dvl + + - + input Dvl3 P-Dvl3 PS-Dvl3 A. B. C. D. E. F. FIGURE 8. A and B, FLAG-Dvl3 was overexpressed either alone or with CK1⑀in HEK-293 cells and subjected to GST pulldown assays with GST-Dvl (or GST tag only as the negative control). FLAG-Dvl3 in the pulldown and input (TCL) was visualized by Western blotting using FLAG antibody. The ratio of individual Dvl forms (pulldown/input) in the shown experiment was quantified by densitometry (B). C, Dvl2-GFP and FLAG-Dvl3 were overexpressed in HEK-293 cells, fixed, and stained for FLAG (Cy3-conjugated secondary antibody). As a measure of proximity between Dvl2-GFP and FLAG-Dvl3, FRET between GFP and Cy3 was determined by photoacceptor bleaching in cells co-transfected with pcDNA3.1 or CK1⑀. The results from 30 individual cells are summarized in a bar graph (C). GFPfluorescencebeforephotobleachingwassetto100%fornormalization.StatisticalsignificancewasanalysesbyStudent’sttest.***,pϽ0.0001.D,MEFcells were either treated with control or Wnt3a-conditioned medium for 3 h, lysed, and subjected to immunoprecipitation (IP) with Dvl2-specific antibody. Unspecific antibody (IgG) was used as a control. The amount of Dvl2 and Dvl3 in the immunoprecipitates and in the input (TCL) was analyzed by Western blotting using several antibodies specific for individual Dvl isoforms. E and F, MEF cells were treated with control and Wnt3a-conditioned media, and endogenous Dvl2 or Dvl3 (baits) was immunoprecipitated. Composition of protein mixtures was analyzed by MS/MS. Graphs (E) indicate the proportion of samples where Dvl2 was found in complex with Dvl3 (bait), and Dvl3 was found in complex with Dvl2 (bait). Bait was identified in 100% of samples in this analysis. N indicates the number of experiments. F, relative amounts of endogenous Dvl3 in samples immunoprecipitated with anti-Dvl2 antibody are shown. Quantification is based on peak intensities with relative bait abundance set to 10. Casein Kinases in Dishevelled Biology 10408 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom Dvl integrates phosphorylation by individual kinases and how CK1⑀ at the same time strongly induces Wnt/␤-catenin signaling and dissolves Dvl puncta, which are required for signal activation. Based on our results, we propose a model that is consistent with the experimental evidence brought by us and others. This model sheds light on some of the so far controversial issues of Dvl biology and is schematized in Fig. 9. Our functional data show that CK2/PAR1 and CK1⑀ cannot replace each other. The effects of either kinase on Dvldependent induction of TCF/LEF-driven transcription can be blocked by inhibition/knockdown of the second kinase. This suggests that CK1⑀ and CK2 (or mammalian homologues of PAR1 from the MARK family of kinases, which may have similar function) act sequentially during the process of Dvl activation. Activity of CK1⑀ is required for the effects of CK2 to take place, and vice versa, the activity of CK2 is required for the effects of CK1⑀ to take place. In line with some earlier reports (24, 27), our data suggest that CK2 activity is constitutive or at least high enough to keep endogenous Dvl predominantly modified in the CK2-dependent manner in untreated cells. It remains to be tested as to the significance of the Wnt-induced increase in CK2 activity reported earlier by Gao and Wang (38). As a consequence, the levels of endogenous dephosphorylated Dvl were very low and only increased upon inhibition of CK2. In contrast, CK1⑀ seems to be a true Wnt-activated kinase required for the formation of PS-Dvl, which is the only hallmark of the activation of endogenous Dvl known to date. Our data suggest that CK1⑀ acts on Dvl modified by a CK2-dependent process. Furthermore, inhibition or low levels of CK2 can block or reduce the ability of CK1 to further phosphorylate Dvl and activate TCF/LEF-mediated transcription. The gradual effects of CK2 and CK1 toward Dvl can be visualized as a phosphorylation-dependent shift both in endogenous and exogenous Dvl (partially phosphorylated by CK2; Fig. 2C). It should be noted, however, that P-Dvl cannot be fully transformed to Dvl by phosphatase treatment, and thus, posttranslational modifications other than phosphorylation triggered by CK2 might be involved. The effects of CK1⑀-mediated phosphorylation of Dvl have been well described earlier in many cellular models. The most prominent CK1⑀-induced changes involve (i) the ability of CK1⑀ to activate Dvl-dependent TCF/LEF-driven transcription (17–20) and (ii) the ability to dissolve Dvl polymers visible as Dvl puncta (14, 20). We show in this study that these two consequences of CK1⑀ phosphorylation are separated both spatially and functionally. First, CK1⑀-mediated induction of TopFlash associated with dynamic recruitment of Axin1 was mediated via the PDZ-Pro-rich region of Dvl (given that DIX domain as the required prerequisite was present). Meanwhile, the ability of CK1⑀ to promote even localization of Dvl3 was mediated by the Dvl3 C terminus. The C-terminal part of Dvl was required for even localization of Dvl, which likely corresponds to the inability to form signalosomes. Dvl3 mutants lacking C terminus were hyperactive, showed punctate localization, and got less hyperphosphorylated after CK1⑀ co-expression, at least as assessed by the mobility shift. However, our data do not allow a direct link of electrophoretic migration of Dvl and Dvl localization. Although there is a strong correlation between these two features of Dvl (both strongly promoted by C terminus), there are few notable exceptions from this rule; for example, CK2 is unable to promote Dvl3 hyperphosphorylation but still efficiently shifts Dvl localization from punctate to even. However, our findings clearly demonstrated that the ability of CK1 to activate downstream signaling and promote even localization of Dvl are separate events with distinct functions. This observation sheds light on some of the unclear issues of Dishevelled biology. It is well established that CK1⑀ is required for signaling downstream of Dvl, which involves Dvl polymerization and subsequent formation of signalosomes, Lrp6 phosphorylation, and stabilization of ␤-catenin (9, 11). However, activation of Dvl-mediated (and CK1-dependent) downstream signaling measured either as the formation of signalosomes (9), phosphorylation of LRP6 (39), internalization of LRP6 (40), or dephosphorylation of ␤-catenin (16) appears in less than 10–15 min, whereas the earliest formation of PS-Dvl is detected first after 30 min (14–16, 41). Functional separation of CK1-triggered downstream signaling and Dvl depolymerization allowed us to formulate the hypothesis that hyperphosphorylated Dvl (PS-Dvl), which is predominantly evenly distributed, is in an inactive form. Furthermore, our data suggest that PS-Dvl is part of a Dvl-inactivating mechanism maintained by CK1⑀. CK1⑀ can, thus, control both the initiation and the termination of signaling (see Fig. 9). According to our results the C-terminal part of Dvl carries the ability to switch off the activity of Dvl in the Wnt/␤-catenin signaling upon CK1⑀ phosphorylation. In line with our findings, the C terminus of Dvl was shown to be an important region for the interaction with known negative regulators of the Wnt/␤catenin pathway. For example, KLHL12-cullin 3 system, which triggers ubiquitination and degradation of Dvl, binds the Dvl3 C terminus (32). The other inhibitor of the Wnt/␤-catenin pathway, atypical receptor-tyrosine kinase Ror2, binds to the C terDvl P-Dvl P-Dvl PS-Dvl puncta puncta puncta even bPDZ C-terminus CK2/PAR1 CK1 CK1? PDZ, Pro DIX, PDZ P P P P P P P Wnt TCF/LEF transcription ON OFF FIGURE 9. Model for the function of casein kinases in Dvl biology. Casein kinase 2 (and probably also PAR1 and its mammalian homologues) behave like constitutively active kinases. Under normal conditions CK2 phosphorylates Dvl in the basic domain/PDZ (bPDZ) to trigger formation of P-Dvl. P- Dvl was then subject to phosphorylation by CK1␦/⑀, which becomes activated upon Wnt stimulation. CK1␦/⑀ phosphorylation of P-Dvl in the PDZ and/or Pro-rich region of Dvl allows recruitment of axin and phosphorylation of Lrp6 (usually detected after 10–15 min) and an executive phase of pathway activation (TCF/LEF-driven transcription), which is dependent on the DIX domains. Dvl in these functional states was capable of polymerization with an electrophoretic migration that resembles P-Dvl. Further phosphorylation of Dvl by CK1 led to the formation of PS-Dvl (usually detected 30–45 min after Wnt stimulation) and ultimately resulted in the inactivation of Dvl in the ␤-catenin pathway. This functional state of Dvl, which requires the C terminus of Dvl, leads to Dvl depolymerization and uniform subcellular distribution. Using this mechanism, CK1␦/⑀ can control both the activation and the termination of the pathway. Casein Kinases in Dishevelled Biology MARCH 25, 2011•VOLUME 286•NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 10409 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom minus of Dvl3 and is the only known Dvl interaction partner that binds specifically to PS-Dvl after CK1 phosphorylation (41). Moreover, we have shown recently that negative effects mediated by the Dvl3 C terminus are dependent on Ror2 (41). Our work, thus, suggests that PS-Dvl, which is the only universal marker of Dvl activation in both canonical and non-canonical Wnt pathways, may be a general inactive signaling intermediate. This view was supported by the fact that PS-Dvl formed irrespective of whether or not activated Wnt triggered the Wnt/␤-catenin pathway (14–16). We have also shown before that PS-Dvl, triggered by Wnt3a and Wnt5a, was indistinguishable in terms of its dynamics, the electrophoretic migration, and the kinases involved in its formation (14). Although it is becoming clear that the C terminus of Dvl is an important functional domain of PS-Dvl, most likely by providing an interaction interface for negative regulators of Wnt/␤-catenin pathway, future work is required to uncover the molecular details of this phenomenon. Acknowledgments—We thank Dr. S. Yanagawa (Kyoto University, Japan), Dr. Randy Moon (University of Washington, Seattle), Dr. S. Byers (Georgetown University, Washington D. C.), Drs. Olga Ossipova, Sergei Sokol, and Marek Mlodzik (Mount Sinai School of Medicine), Dr. Isabel Dominguez (Boston University School of Medicine), Dr. Madelone Maurice (UMC Utrecht), and Dr. J. M. Graff (University of Texas, Dallas) for providing plasmids. LWnt3A cells (ATCC#CRL-2647) were provided by Dr. Vladimir Korinek (Institute of Molecular Genetics, Prague). REFERENCES 1. Clevers, H. (2006) Cell 127, 469–480 2. Malanchi, I., and Huelsken, J. (2009) Curr. Opin. Oncol. 21, 41–46 3. Schulte, G., and Bryja, V. (2007) Trends Pharmacol. Sci. 28, 518–525 4. Angers, S., and Moon, R. T. (2009) Nat. Rev. Mol. Cell Biol. 10, 468–477 5. Malbon, C. C., and Wang, H. Y. (2006) Curr. Top. Dev. Biol. 72, 153–166 6. Wallingford, J. B., and Habas, R. (2005) Development 132, 4421–4436 7. Gao, C., and Chen, Y. G. (2010) Cell. Signal. 22, 717–727 8. Wharton, K. A., Jr. (2003) Dev. Biol. 253, 1–17 9. Bilic, J., Huang, Y. L., Davidson, G., Zimmermann, T., Cruciat, C. M., Bienz, M., and Niehrs, C. (2007) Science 316, 1619–1622 10. Schwarz-Romond, T., Metcalfe, C., and Bienz, M. (2007) J. Cell Sci. 120, 2402–2412 11. Schwarz-Romond, T., Fiedler, M., Shibata, N., Butler, P. J., Kikuchi, A., Higuchi, Y., and Bienz, M. (2007) Nat. Struct. Mol. Biol. 14, 484–492 12. Schwarz-Romond, T., Merrifield, C., Nichols, B. J., and Bienz, M. (2005) J. Cell Sci. 118, 5269–5277 13. Smalley, M. J., Signoret, N., Robertson, D., Tilley, A., Hann, A., Ewan, K., Ding, Y., Paterson, H., and Dale, T. C. (2005) J. Cell Sci. 118, 5279–5289 14. Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007) J. Cell Sci. 120, 586–595 15. Gonza´lez-Sancho, J. M., Brennan, K. R., Castelo-Soccio, L. A., and Brown, A. M. (2004) Mol. Cell. Biol. 24, 4757–4768 16. Bryja, V., Schulte, G., and Arenas, E. (2007) Cell. Signal. 19, 610–616 17. Peters, J. M., McKay, R. M., McKay, J. P., and Graff, J. M. (1999) Nature 401, 345–350 18. Sakanaka, C., Leong, P., Xu, L., Harrison, S. D., and Williams, L. T. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 12548–12552 19. Kishida, M., Hino, Si., Michiue, T., Yamamoto, H., Kishida, S., Fukui, A., Asashima, M., and Kikuchi, A. (2001) J. Biol. Chem. 276, 33147–33155 20. Cong, F., Schweizer, L., and Varmus, H. (2004) Mol. Cell. Biol. 24, 2000–2011 21. Foldynova´-Trantírkova´, S., Sekyrova´, P., Tmejova´, K., Brumovska´, E., Bernatík, O., Blankenfeldt, W., Krejcí, P., Kozubík, A., Dolezal, T., Trantírek, L., and Bryja, V. (2010) Breast Cancer Res. 12, R30 22. Bryja, V., Gradl, D., Schambony, A., Arenas, E., and Schulte, G. (2007) Proc. Natl. Acad. Sci. U.S.A. 104, 6690–6695 23. Bryja, V., Schambony, A., Caja´nek, L., Dominguez, I., Arenas, E., and Schulte, G. (2008) EMBO Rep. 9, 1244–1250 24. Willert, K., Brink, M., Wodarz, A., Varmus, H., and Nusse, R. (1997) EMBO J. 16, 3089–3096 25. Ossipova, O., Dhawan, S., Sokol, S., and Green, J. B. (2005) Dev. Cell 8, 829–841 26. Korinek, V., Barker, N., Morin, P. J., van Wichen, D., de Weger, R., Kinzler, K. W., Vogelstein, B., and Clevers, H. (1997) Science 275, 1784–1787 27. Song, D. H., Sussman, D. J., and Seldin, D. C. (2000) J. Biol. Chem. 275, 23790–23797 28. Zeng, X., Huang, H., Tamai, K., Zhang, X., Harada, Y., Yokota, C., Almeida, K., Wang, J., Doble, B., Woodgett, J., Wynshaw-Boris, A., Hsieh, J. C., and He, X. (2008) Development 135, 367–375 29. Kishida, S., Yamamoto, H., Hino, S., Ikeda, S., Kishida, M., and Kikuchi, A. (1999) Mol. Cell. Biol. 19, 4414–4422 30. Li, L., Yuan, H., Xie, W., Mao, J., Caruso, A. M., McMahon, A., Sussman, D. J., and Wu, D. (1999) J. Biol. Chem. 274, 129–134 31. Boutros, M., Paricio, N., Strutt, D. I., and Mlodzik, M. (1998) Cell 94, 109–118 32. Angers, S., Thorpe, C. J., Biechele, T. L., Goldenberg, S. J., Zheng, N., MacCoss, M. J., and Moon, R. T. (2006) Nat. Cell Biol. 8, 348–357 33. Klein, T. J., Jenny, A., Djiane, A., and Mlodzik, M. (2006) Curr. Biol. 16, 1337–1343 34. Klimowski, L. K., Garcia, B. A., Shabanowitz, J., Hunt, D. F., and Virshup, D. M. (2006) FEBS J. 273, 4594–4602 35. Itoh, K., Brott, B. K., Bae, G. U., Ratcliffe, M. J., and Sokol, S. Y. (2005) J. Biol. 4, 3 36. Jenny, A., Reynolds-Kenneally, J., Das, G., Burnett, M., and Mlodzik, M. (2005) Nat. Cell Biol. 7, 691–697 37. Lee, Y. N., Gao, Y., and Wang, H. Y. (2008) Cell. Signal. 20, 443–452 38. Gao, Y., and Wang, H. Y. (2006) J. Biol. Chem. 281, 18394–18400 39. Khan, Z., Vijayakumar, S., de la Torre, T. V., Rotolo, S., and Bafico, A. (2007) Mol. Cell. Biol. 27, 7291–7301 40. Yamamoto, H., Komekado, H., and Kikuchi, A. (2006) Dev. Cell 11, 213–223 41. Witte, F., Bernatik, O., Kirchner, K., Masek, J., Mahl, A., Krejci, P., Mundlos, S., Schambony, A., Bryja, V., and Stricker, S. (2010) FASEB J. 24, 2417–2426 42. Lee, J. S., Ishimoto, A., and Yanagawa, S. (1999) J. Biol. Chem. 274, 21464–21470 43. Orford, K., Crockett, C., Jensen, J. P., Weissman, A. M., and Byers, S. W. (1997) J. Biol. Chem. 272, 24735–24738 44. Dominguez, I., Mizuno, J., Wu, H., Imbrie, G. A., Symes, K., and Seldin, D. C. (2005) Mol. Cell. Biochem. 274, 125–131 45. McKay, R. M., Peters, J. M., and Graff, J. M. (2001) Dev. Biol. 235, 388–396 46. Tauriello, D. V., Haegebarth, A., Kuper, I., Edelmann, M. J., Henraat, M., Canninga-van Dijk, M. R., Kessler, B. M., Clevers, H., and Maurice, M. M. (2010) Mol. Cell 37, 607–619 Casein Kinases in Dishevelled Biology 10410 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286•NUMBER 12•MARCH 25, 2011 atKarolinskainstitutetlibrary,onMarch25,2011www.jbc.orgDownloadedfrom       Vítězslav Bryja, 2014    Attachments      #13      Chaki M, Airik R, Ghosh AK, Giles RH, Chen R, Slaats GG, Wang H, Hurd TW, Zhou W,  Cluckey A, Gee HY, Ramaswami G, Hong CJ, Hamilton BA, Cervenka I, Ganji RS, Bryja V,  Arts HH, van Reeuwijk J, Oud MM, Letteboer SJ, Roepman R, Husson H, Ibraghimov‐ Beskrovnaya O, Yasunaga T, Walz G, Eley L, Sayer JA, Schermer B, Liebau MC, Benzing T,  Le  Corre  S,  Drummond  I,  Janssen  S,  Allen  SJ,  Natarajan  S,  O'Toole  JF,  Attanasio  M,  Saunier S, Antignac C, Koenekoop RK, Ren H, Lopez I, Nayir A, Stoetzel C, Dollfus H,  Massoudi R, Gleeson JG, Andreoli SP, Doherty DG, Lindstrad A, Golzio C, Katsanis N,  Pape  L,  Abboud  EB,  Al‐Rajhi  AA,  Lewis  RA,  Omran  H,  Lee  EY,  Wang  S,  Sekiguchi  JM,  Saunders R, Johnson CA, Garner E, Vanselow K, Andersen JS, Shlomai J, Nurnberg G,  Nurnberg P, Levy S, Smogorzewska A, Otto EA, Hildebrandt F. (2012): Exome capture  reveals  ZNF423  and CEP164 mutations,  linking  renal  ciliopathies  to  DNA  Damage  response signaling. Cell. 150(3): 533‐48.      Impact factor (2012): 31.957  Times cited (without autocitations, WoS, Feb 21st 2014): 33  Significance: Identification of CEP164 as a gene mutated in ciliopathies. Because it is  known  that  CEP164  controls  cilia  formation  and  DNA  damage,  this  protein  provided  a  link  between  these  two  processes  in  ciliopathies.  As  part  of  this  project, interaction between CEP164 and Dvl3 was revealed. This aspect pointed  towards the role of Dishevelled in the centrosomal biology.  Contibution of the author/author´s team: Identification of CEP164‐Dvl interaction by  mass spectrometry, identification of binding domains between the two proteins  and  analysis  of  disease‐specific  CEP164  mutants  with  respect  to  its  interaction  with Dvl3.                Exome Capture Reveals ZNF423 and CEP164 Mutations, Linking Renal Ciliopathies to DNA Damage Response Signaling Moumita Chaki,1,39 Rannar Airik,1,39 Amiya K. Ghosh,1 Rachel H. Giles,3 Rui Chen,4 Gisela G. Slaats,3 Hui Wang,4 Toby W. Hurd,1 Weibin Zhou,1 Andrew Cluckey,1 Heon Yung Gee,1 Gokul Ramaswami,1 Chen-Jei Hong,6 Bruce A. Hamilton,6 Igor Cervenka,7 Ranjani Sri Ganji,7 Vitezslav Bryja,7,8 Heleen H. Arts,9 Jeroen van Reeuwijk,9 Machteld M. Oud,9 Stef J.F. Letteboer,9 Ronald Roepman,9 Herve´ Husson,10 Oxana Ibraghimov-Beskrovnaya,10 Takayuki Yasunaga,11 Gerd Walz,11 Lorraine Eley,12 John A. Sayer,12 Bernhard Schermer,13,14,16 Max C. Liebau,13,15 Thomas Benzing,13,14,16 Stephanie Le Corre,19 Iain Drummond,19 Sabine Janssen,1 Susan J. Allen,1 Sivakumar Natarajan,1 John F. O’Toole,20 Massimo Attanasio,21 Sophie Saunier,22 Corinne Antignac,22 Robert K. Koenekoop,23 Huanan Ren,23 Irma Lopez,23 Ahmet Nayir,24 Corinne Stoetzel,25 Helene Dollfus,25 Rustin Massoudi,26 Joseph G. Gleeson,26 Sharon P. Andreoli,27 Dan G. Doherty,28 Anna Lindstrad,29 Christelle Golzio,29 Nicholas Katsanis,29 Lars Pape,30 Emad B. Abboud,31 Ali A. Al-Rajhi,31 Richard A. Lewis,5 Heymut Omran,32 Eva Y.-H.P. Lee,33 Shaohui Wang,33 JoAnn M. Sekiguchi,2 Rudel Saunders,2 Colin A. Johnson,34 Elizabeth Garner,35 Katja Vanselow,36 Jens S. Andersen,36 Joseph Shlomai,18 Gudrun Nurnberg,14,17 Peter Nurnberg,14,17 Shawn Levy,37 Agata Smogorzewska,35 Edgar A. Otto,1 and Friedhelm Hildebrandt1,2,38,* 1Department of Pediatrics and Communicable Diseases 2Department of Human Genetics University of Michigan, Ann Arbor, MI 48109, USA 3Department of Nephrology and Hypertension, University Medical Center, Utrecht, The Netherlands 4HGSC Department of Molecular and Human Genetics 5Department of Ophthalmology Baylor College of Medicine, Houston, TX, USA 6Department of Medicine, Division of Medical Genetics, Department of Cellular and Molecular Medicine, and Institute for Genomic Medicine, George Palade Laboratories, Room 256, UCSD School of Medicine, 9500 Gilman Drive, San Diego, CA 92093, USA 7Institute of Experimental Biology, Faculty of Science, Masaryk University, 61137 Brno, Czech Republic 8Department of Cytokinetics, Institute of Biophysics, Academy of Sciences of the Czech Republic 61265 Brno, Czech Republic 9Department of Human Genetics, Nijmegen Centre for Molecular Life Sciences and Institute for Genetic and Metabolic Disease, Radboud University Nijmegen Medical Centre, 6525 GA, Nijmegen, The Netherlands 10Genzyme Corporation, Cell Biology, Framingham, MA 01701, USA 11Renal Division, University Freiburg Medical Center, 7900 Freiburg, Germany 12Institute of Genetic Medicine, Newcastle University, Newcastle upon Tyne, NE1 3BZ, UK 13Department II of Internal Medicine and Center for Molecular Medicine 14Cologne Excellence Cluster on Cellular Stress Responses in Aging-Associated Diseases 15Department of Pediatrics and Adolescent Medicine 16Systems Biology of Aging 17Cologne Center for Genomics and for Molecular Medicine University of Cologne, 50937 Cologne, Germany 18Department of Microbiology and Molecular Genetics, The Institute for Medical Research Israel-Canada (IMRIC), Faculty of Medicine, The Hebrew University-Hadassah Medical School, Jerusalem, 91120 Israel 19Nephrology Division, Massachusetts General Hospital and Department of Genetics, Harvard Medical School, Charlestown, MA 02129, USA 20Division of Nephrology, Department of Internal Medicine, MetroHealth Medical Center, and Case Western Reserve University School of Medicine, Cleveland, OH 44109-1998, USA 21Department of Internal Medicine and Eugene McDermott Center for Growth and Development, University of Texas Southwestern Medical Center, Dallas TX 75390, USA 22Inserm U983, Paris Descartes University, Hoˆ pital Necker-Enfants Malades, Assistance Publique-Hoˆ pitaux de Paris, Paris, France 23McGill Ocular Genetics Laboratory, Montreal Children’s Hospital, McGill University Health Centre, Montreal, H3H 1P3, Canada 24Department of Pediatric Nephrology, Faculty of Medicine, University of Istanbul, Istanbul, Turkey 25Laboratoire de Ge´ ne´ tique Me´ dicale EA3949, Equipe AVENIR-Inserm, Faculte´ de Me´ decine, Universite´ de Strasbourg, 11 rue Humann, 67000 Strasbourg, France 26Howard Hughes Medical Institute, Department of Pediatrics, University of California, San Diego, La Jolla, CA 92093, USA 27Department of Pediatrics, James Whitcomb Riley Hospital for Children, Indiana University Medical Center, Indianapolis, IN 46202, USA 28Division of Genetic Medicine, Department of Pediatrics, University of Washington, Center for Integrative Brain Research, Seattle Children’s Hospital, Seattle, WA, 98101 USA 29Center for Human Disease Modeling, Duke University Medical Center, Durham, NC 27710, USA 30Department of Pediatric Nephrology, Hannover Medical School, Hannover 30625, Germany 31King Khaled Eye Specialist Hospital, Riyadh 11462, Kingdom of Saudi Arabia Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 533 32Klinik und Poliklinik fu¨ r Kinder- und Jugendmedizin, Allgemeine Pa¨ diatrie, Universita¨ tsklinikum Mu¨ nster, Mu¨ nster 48149, Germany 33University of California Irvine, Department of Biological Chemistry, Irvine, CA 92697, USA 34Leeds Institute of Molecular Medicine, St James’s University Hospital, Leeds, LS9 7TF, UK 35Laboratory of Genome Maintenance, The Rockefeller University, New York, NY 10065, USA 36Department of Biochemistry and Molecular Biology, University of Southern Denmark, DK-5230, Odense, Denmark 37HudsonAlpha Institute for Biotechnology, 601 Genome Way, Huntsville, AL 35806, USA 38Howard Hughes Medical Institute, Chevy Chase, MD 20815, USA 39These authors contributed equally to this work *Correspondence: fhilde@umich.edu http://dx.doi.org/10.1016/j.cell.2012.06.028 SUMMARY Nephronophthisis-related ciliopathies (NPHP-RC) are degenerative recessive diseases that affect kidney, retina, and brain. Genetic defects in NPHP gene products that localize to cilia and centrosomes defined them as "ciliopathies.’’ However, disease mechanisms remain poorly understood. Here, we identify by whole-exome resequencing, mutations of MRE11, ZNF423, and CEP164 as causing NPHPRC. All three genes function within the DNA damage response (DDR) pathway. We demonstrate that, upon induced DNA damage, the NPHP-RC proteins ZNF423, CEP164, and NPHP10 colocalize to nuclear foci positive for TIP60, known to activate ATM at sites of DNA damage. We show that knockdown of CEP164 or ZNF423 causes sensitivity to DNA damaging agents and that cep164 knockdown in zebrafish results in dysregulated DDR and an NPHP-RC phenotype. Our findings link degenerative diseases of the kidney and retina, disorders of increasing prevalence, to mechanisms of DDR. INTRODUCTION Nephronophthisis (NPHP) is a recessive cystic kidney disease that represents the most frequent genetic cause of end-stage kidney disease in the first three decades of life. NPHP-related ciliopathies (NPHP-RC) are single-gene recessive disorders that affect kidney, retina, brain, and liver by prenatal-onset dysplasia or by organ degeneration and fibrosis in early adulthood. Identification of recessive mutations in more than ten different genes (NPHP1-NPHP13) revealed that their gene products share localization at the primary cilia-centrosomes complex and mitotic spindle poles in a cell-cycle-dependent manner, characterizing them as retinal-renal ‘‘ciliopathies’’ (Ansley et al., 2003; Hildebrandt et al., 2011). Multiple signaling pathways downstream of cilia have been implicated in the disease mechanisms of NPHP-RC, including Wnt signaling (Germino, 2005; Simons et al., 2005) and Shh signaling (Huangfu and Anderson, 2005; Huangfu et al., 2003). However, despite convergence of ciliopathy pathogenesis at cilia and centrosomes it remains largely unknown what signaling pathways downstream of cilia and centrosome function operate in the disease mechanisms that generate the NPHP-RC phenotypes. Centrosomal proteins have been recently implicated in DNA damage response (DDR). Both pericentrin (PCNT), a core centrosomal protein (Doxsey et al., 1994), as well as CEP152, encoding a centrosomal protein required for centriolar duplication (Blachon et al., 2008), are defective in Seckel syndrome, an autosomal-recessive disorder characterized by dwarfism, microcephaly, and mental retardation (Griffith et al., 2008; Kalay et al., 2011; Rauch et al., 2008). PCNT- and CEP152 mutant cells are also defective in ATR-dependent DDR signaling, consistent with the fact that the first mutation identified in Seckel syndrome was in ataxia-telangiectasia mutated and RAD3-related (ATR), a key phosphoinositide 3-kinase-related protein kinase involved in DDR signaling (O’Driscoll et al., 2003), but the mechanism of the signaling defect is not fully understood. The known NPHP genes explain less than 50% of all cases with NPHP-RC, and many of the single-gene causes of NPHPRC are still unknown (Otto et al., 2011). The finding that some of the recently identified genetic causes of NPHP-RC are exceedingly rare (Attanasio et al., 2007) necessitates the ability to identify novel single-gene causes of NPHP-RC in single affected families. To achieve this goal, we developed a strategy that combines homozygosity mapping (HM) with whole-exome resequencing (WER) (Otto et al., 2010). Because this approach allows identification of multiple different causes of NPHP-RC within a short time frame, it has the potential of delineating pathogenic pathways. Using this approach, we identify here mutations in three NPHP-RC genes, MRE11, ZNF423, and CEP164, which together suggest involvement of a DDR signaling pathway in NPHP-RC pathogenesis. RESULTS Whole-Exome Resequencing Accelerates Discovery of NPHP-RC Genes Identification of monogenic causes of ciliopathies is limited by their rarity (Attanasio et al., 2007), necessitating methods to identify ciliopathy-causing genes in single families by using WER. However, WER typically yields hundreds of variants from normal reference sequence (Ng et al., 2009), whereas only a single-gene mutation will represent the disease cause. To overcome this limitation, we here combined WER with HM (Hildebrandt et al., 2009) in sib pairs affected with NPHP-RC and performed functional analysis of the identified genes (Otto et al., 2010). 534 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. HM yielded positional candidate regions of homozygosity by descent (Hildebrandt et al., 2009) in families A3471 (two regions), F874 (nine regions), and KKESH001-7 (14 regions) (Figure 1), who had one or more features of NPHP-RC, including NPHP, retinal degeneration, liver fibrosis, or cerebellar degeneration/ hypoplasia (Table 1). We then performed WER in one affected individual of each of the three NPHP-RC families (Ng et al., 2009; Otto et al., 2010). Remarkably, each of three NPHP-RC A B C D E F Figure 1. Identification of Recessive Mutations in MRE11, ZNF423, and CEP164 in NPHP-RC Using HM and WER Data regarding HM and mutations are shown for family A3471 with MRE11 mutation (A and B), family F874 with ZNF423 mutation (C and D), and family KKESH001 with CEP164 mutation (E and F). (A, C, and E) Nonparametric lod scores (NPL) are plotted across the human genome in three families (A3471, F874, and KKESH001) with NPHP-RC (see also Table 1). The x axis shows SNP positions on human chromosomes concatenated from p-ter (left) to q-ter (right). Genetic distance is given in cM. Maximum NPL peaks (Hildebrandt et al., 2009) (red circles) indicate candidate regions of homozygosity by descent. The genes MRE11, ZNF423, and CEP164 are positioned (arrow heads) within one of the maximum NPL peaks. (B, D, and F) Homozygous mutations detected in families with NPHPRC. Family number (underlined), mutation (arrowheads), and predicted translational changes (in parenthesis) are indicated (see also Table 1). Sequence traces are shown for mutations above normal controls. (For additional mutations in other families see also Table 1 and Figure S2). Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 535 Table1.MutationsofMRE11,ZNF423andCEP164infamilieswithNPHP-RC FamilyIndividuals Ethnic Origin Nucleotide Alterationa,b (Hg19Position) Deduced Protein Change Exon (State) ContinuousAmino AcidSequence Conservation Parental Consanguinity Kidney (AgeatESKF)Eye(AgeatRD)Other(atAge) MRE11 A3471-21and-22Pakistanic.1897C>T (Chr11:94,170,372) p.R633X16(hom)N/AYesNorenalfailureNormal-21:-21:CVA(MRI),ataxia, dysarthria,myoclonus; -22:CVA(MRI),ataxia ZNF423 F874-21and-22Turkeyc.2738C>T(Chr16: 49,670,325) p.P913L5(hom)(D.rerio)YesNPHPNDCVHInfantileNPHP Situsinversus A106-21and-22Icelandc.1518delC(Chr16: 49,671,545) p.P506fsX435(het)(X.tropicalis)NoPKDLCACVH(Joubert) A111-21?c.3829C>T(Chr16: 49,525,212) p.H1277Y9(het)(D.rerio)?PKDRDCVH,NPHP,perinatal breathingabnormality, tonguetumor CEP164 F319-21and-22Turkeyc.32A>C(Chr11: 117,209,334) p.Q11P3(hom)(Ch.Reinhardtii)YesNPHP,noBx; -21:(8years); -22:(8years) -21:RD(11yr, notyetblind);-22: noRPat8yrs -21:obesity?no LF;-22:obesity?LF? F59-21,-22,-23USA (Europe) c.277C>T,(Chr11: 117,222,588) c.1573C>T(Chr11: 117,252,580) p.R93W, p.Q525X 5(het), 13(het) (Ch.Reinhardtii), N/A NoNPHP,noBx; -21:(9years); -22:(8years); -23:normal -21:RD(6years); -22:LCA(legally blindat5months); -23:(2years) -22:NY(birth), mildAI;-23:seizuresc , substantialDD,mildID NPH505NDc.1726C>T(Chr11: 117,257,920) p.R576X15(hom)N/AYesNPHP,Bx(8yr)RDandflatERG (notblind) CVH,FD,bilateralPD, bronchiectasis(1mo), abnormalLFT,obesity KKESH001-7Saudic.4383A>G(Chr11: 117,282,884) p.X1460W extX57 33(hom)N/AYesnormal(RD)LCA,flat ERG(blind<2yr) N/A AI,aorticinsufficiency;Bx,Kidneybiopsy;CVH,cerebellervermishypoplasia;DD,developmentaldelay;ERG,electroretinogram;ESKF,end-stagekidneyfailure;FD,facialdysmorphism;het, heterozygous;hom,homozygous;ID,intellectualdisability;LCA,Lebercongenitalamaurosis;LF,liverfibrosis;LFT,liverfunctiontests;MRI,magneticresonanceimaging;N/A,notapplicable; ND,nodata;NPHP,nephronophthisis;NPHP-RC,nephronophthisis-relatedciliopathies;NY,nystagmus;PD,polydactyly;RD,retinaldegeneration;SS,shortstature. a Allmutationswereabsentfrom>270healthycontrolindividualsandfromtheESPExomeVariantServerdatabase,excepttheCEP164variantp.R576X(allelefrequencyinEuropeanAmericans 1/7,019). b cDNAmutationnumberingisbasedonhumanreferencesequencesNM_014956.4forMRE11,NM_015069.2forZNF423,andNP_055771forCEP164,where+1correspondstotheAofATG starttranslationcodon. c Seizureswereintractable,generalizedand/orpartialcomplex. 536 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. genes consecutively identified by this approach, MRE11, ZNF423, and CEP164, suggested a functional connection to the DDR pathway (Figure 1; Table 1). A Mutation of MRE11 in Progressive Cerebellar Degeneration Suggests Link to DDR In family A3471, two siblings had cerebellar vermis hypoplasia (CVH), a central feature of NPHP-RC (Table 1). Homozygosity mapping yielded two candidate loci (Figure 1A). WER detected a homozygous truncation mutation (p.R633X) of MRE11 (Figure 1B; Table 1) previously described for CVH in another Pakistani family (Stewart et al., 1999), suggesting a founder effect for this allele. MRE11 is an essential component of the ATM-Chk2 pathway of DDR (Figure S1 available online), where it recruits ATM (ataxia telangiectasia-mutated) to sites of DNA doublestrand breaks (Figure S1A). Rediscovery of this MRE11 mutation in family A3471 thus generated an unexpected link between NPHP-RC phenotype and the ATM pathway of DDR signaling (Figure S1A). Patients with the NPHP-RC Joubert Syndrome Have Defects in ZNF423 Another link of NPHP-RC to the ATM pathway of DDR signaling emerged from HM and WER in two siblings (F874) with infantile onset NPHP, CVH, and situs inversus (Table 1). SNP mapping yielded nine candidate regions of homozygosity by descent (Figure 1C). We identified in both affected individuals a homozygous missense mutation (p.P913L; conserved in vertebrates) of ZNF423 (Figure 1D). In addition, when examining 96 additional Joubert syndrome (JS) subjects, we detected two heterozygous-only mutations of ZNF423: p.P506fsX43 in family A106 and p.H1277Y in individual A111-21 (Table 1). Mutations of the mouse ortholog Zfp423 cause reduced proliferation and abnormal development of midline neural progenitors resulting in a loss of the cerebellar vermis (Alcaraz et al., 2006; Cheng et al., 2007) similar to that seen in JS patients with CVH. ZNF423 encodes a protein with 30 zinc fingers (Figure 2A). Mouse models display phenotypic variability that is subject to modifier genes, environment, and stochastic effects (Alcaraz et al., 2011; Alcaraz et al., 2006), consistent with the variable presentations of NPHP-RC patients. The homozygous mutation p.P913L, located between zinc fingers 21 and 22 (Figure 2A), most likely exerts recessive loss-of-function, analogous to the Zfp423 mouse models. We next examined whether the heterozygous-only mutations (Table 1) lead to loss of function via a dominant mechanism, using a proliferation assay in P19 cells (Figures 2B–2D). Mutations were engineered into a FLAG-tagged ZNF423 cDNA and assayed by a S-phase index, defined as the proportion of transfected cells that incorporate BrdU in 1 hr, 48 hr after transfection. Simple loss-of-function alleles should not interfere with endogenous Zfp423 activity in this assay. Indeed, overexpression of either wild-type or the homozygous p.P913L allele had no effect (Figure 2D). However, transfection with either the p.P506fsX43 frame-shifting allele, which removes the zinc fingers required for SMAD (similar to mothers against decapentaplegic) and EBF (early B cell factor) interactions, and the H1277Y substitution allele, which destroys the terminal zinc finger required for EBF interaction, reduced the mitotic index to little more than half that of cells transfected with green fluorescent protein (GFP) control vector or other alleles of ZNF423 (Figures 2B– 2D). A dominant mechanism is plausible for the two heterozygous mutations, as each is predicted to interfere selectively with a subset of interaction domains (Figure 2A). Neither subject had siblings, and DNA from parents was not available to determine whether the mutations occurred de novo. We detected five additional putative mutations in highly conserved (including histidine knuckle) residues of ZNF423 among JS families (Table S1). Although these mutations have not been confirmed functionally, the high incidence of predicted deleterious mutations found in patients but absent from 270 healthy control individuals, dbSNP, and 1000 Genomes Project data further support identification of ZNF423 as a causal gene in NPHP-RC and JS. ZNF423/OAZ was recently shown to interact with the DNA ds-damage sensor PARP1 (poly-ADP ribosyl polymerase 1) (Ku et al., 2003), which recruits MRE11 and ATM to sites of DNA damage (Figure S1A). This indirectly linked ZNF423 to the ATM pathway of DNA damage signaling (Figure S1A). We therefore tested whether ZNF423 mutations affect interaction between ZNF423 and PARP1. Coimmunoprecipitation verified the association of ZNF423 and PARP1 in reciprocal assays (Figure 2E). More importantly, the truncating mutation P506fsX43, which we detected in a JS patient (Table 1), abrogates this interaction (Figure 2E), whereas H1277Y inhibits multimerization of ZNF423 (Figure 2E). In addition, depletion of ZNF423 mRNA caused sensitivity to DNA damaging agents (see below). Furthermore, we identified ZNF423 as a direct interaction partner of CEP290/NPHP6, which is mutated in NPHP-RC (Sayer et al., 2006; Valente et al., 2006). In a yeast two-hybrid screen of human fetal brain library with a CEP290 (JAS2; amino acids 1917–2479) ‘‘bait’’ we found three in-frame ‘‘prey’’ sequences corresponding to ZNF423 (amino acids 178-406). This interaction was confirmed (Figures 2F and 2G). CEP290/NPHP6 is known to interact with the NPHP-RC protein NPHP5 (Scha¨ fer et al., 2008) and localizes to the ciliary transition zone (Sang et al., 2011). Mutations of CEP164 Cause NPHP-RC We obtained 14 candidate regions by HM in a Saudi family (KKESH001) of first-cousin parents with a child who had LCA (which can be allelic with NPHP-RC) with nystagmus, hyperopic discs, vascular attenuation, diffuse retinal pigment epithelium atrophy, and nonrecordable ERG (Table 1) (Figure 1E). Using WER we detected a homozygous point mutation in CEP164 (centrosomal protein 164 kDa) that abolished the termination codon, adding 57 amino acid residues to the open reading frame (p.X1460WextX57) (Figure 1F, Table 1). The mutation was absent from 96 Saudi healthy controls and from 224 North American LCA patients who lack mutations in other known LCA genes. We performed exon-PCR and Sanger sequencing of all 31 coding exons for one affected individual in each of 856 different NPHP-RC families (see Extended Experimental Procedures). We detected both mutated CEP164 alleles in each of three additional families with NPHP-RC (Table 1; Figure S2). We thereby identified recessive mutations of CEP164 as an additional cause of NPHP-RC. Because of the significant overlap of phenotypic Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 537 features with other forms of NPHP-RC we introduce the alias ‘‘NPHP14’’ for ZNF423 and ‘‘NPHP15’’ for the CEP164 protein. Although the number of families with CEP164 mutation is small, our findings support a gradient of genotype-phenotype correlations characteristic of NPHP-RC (Table 1), in which null mutations cause the severe dysplastic phenotypes of Meckel syndrome and JS, whereas hypomorphic alleles cause the milder degenerative phenotypes of NPHP and SLSN (Hildebrandt et al., 2011). CEP164 is transcribed into three common isoforms (Figures S2A–S2C) and is part of the photoreceptor sensory cilium proteome (Liu et al., 2007). To study subcellular localization of the CEP164 protein, we utilized antibodies against human CEP164 for immunoblotting and immunofluorescence (Figure S3). Mutation of CEP164 Interferes with Ciliogenesis By confocal microscopy of GFP-labeled CEP164 protein with other labels, we show that CEP164 colocalizes in hTERTRPE cells with the mother centriole, with the mitotic spindle poles, and with the abscission structure in a cell-cycle-dependent way (Figure S4), a feature characteristic of proteins involved in single-gene ciliopathies (Otto et al., 2010; Graser A B C D E F G Figure 2. Two ZNF423 Mutations Have Dominant Negative Characteristics, ZNF423 Mutation Abrogates Interaction with PARP1, and ZNF423 Directly Interacts with the NPHP-RC Protein CEP290/NPHP6 (A) Amino acid residues altered by NPHP-RC mutations in ZNF423 are drawn in relation to functional annotation of its 30 Zn-fingers. (B–D) S-phase index assay (fraction of transfected cells incorporating BrdU) for P19 cells transfected with either wild-type or mutant ZNF423. (B) Representative field of cells transfected with wildtype ZNF423 shows high frequency of BrdU+ FLAG+ double-positive cells. (C) ZNF423–H1277Y transfected cells exhibits fewer FLAG–positive cells and a lower proportion that are double positive. (D) S-phase index measured in duplicate transfections for each of three DNA preparations per construct. A GFP construct was used as a nonspecific control. Constructs with P506fsX43 and H1277Y mutations detected in NPHP-RC show significantly reduced S-phase index (p < 10À5 , ANOVA with post-hoc Tukey honestly significant difference [HSD]). (E) ZNF423 interacts with PARP1. P19 cells were cotransfected with expression constructs for N terminally FLAG-tagged human full-length ZNF423 and V5-tagged human full length PARP1. Comparable amounts of both proteins were present in all lysates (lower panels). Proteins were precipitated, using anti-V5 (upper panels) and anti-FLAG antibodies (middle panels), respectively. Reciprocal coIP demonstrates interaction between ZNF423 and PARP1. Note that the ZFN423 mutation P506fsX43 abrogates this interaction (arrowhead) and that mutation H1277Y diminishes ZNF423 multimerization (arrow). (F–G) ZNF423 directly interacts with CEP290/ NPHP6. (F) A human fetal brain yeast two-hybrid library screened with human CEP290/NPHP6 (JAS2; aa 1917–2479) fused to the DNA-binding domain of GAL4 (pDEST32) identified ZNF423 as a direct interaction partner of CEP290/NPHP6. The interaction was confirmed using direct yeast two-hybrid assay where 1 and 2 represent colony growth of CEP290 bait with ZNF423 prey. a–e are controls for colony growth on medium deficient in histidine, leucine and tryptophan. (G) HEK293T were cotransfected with human V5-tagged partial human V5-CEP290 clone and GFP-tagged fulllength human ZNF423 clone. Immunoprecipitation with anti-V5 (lane 2), but not control IgG (lane 3) precipitated both the V5-tagged CEP290 (arrowhead) as well as GFP-tagged ZNF423 (arrow). 538 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. et al., 2007) and that this colocalization is abrogated by mutations (Figure 3, Figures S3C–S3F). We thus demonstrated lack of centrosomal localization for the truncating mutation p.Q525X and for an equivalent of the p.X1460WextX57 mutation. Loss of function of several genes that cause nephronophthisis in NPHP-RC cause disruption of 3D architecture of renal epithelial cell culture (Otto et al., 2010; Sang et al., 2011). To evaluate CEP164 by this criterion, we transfected murine kidney IMCD3 cells with siRNA oligonucleotides against murine Cep164, or random sequences (Ctrl) in 3D spheroid growth assays. Cells transfected with siCep164 developed spheroids with overall normal architecture and size, but with markedly reduced frequency of cilia (Figures 3E–3H). We conclude that Cep164 affects ciliogenesis or maintenance but that the overall architecture of renal 3D growths is not as grossly affected as we have previously seen for knockdown of other NPHP-RC genes (Sang et al., 2011). A B C D E F G H Figure 3. Expression of Mutant CEP164 in Renal Epithelial Cells Abrogates Localization to Centrosomes (A) Immunofluorescence using a-SDCCAG8/ NPHP10-CG antibody in MDCK cells, labels both centrioles, whereas a-CEP164-ENR antibody demonstrates CEP164 staining at the mother centriole only. (B) Inducible overexpression of N terminally GFP-tagged human full-length CEP164 isoform 1 (NGFP-CEP164-WT) in IMCD3 cells demonstrates, in addition to a cytoplasmic expression pattern, localization at one of the two centrioles (inset, arrow heads) consistent with selective localization to the mother centriole (Graser et al., 2007). Both centrioles are stained with an anti-gtubulin antibody. (C) In contrast, the centrosomal signal is abrogated upon overexpression of an N terminally GFP-tagged truncated CEP164 construct representing the mutation p.Q525X. (D) The number of centrosomes positive for CEP164 is reduced upon overexpression of C terminally GFP-tagged human full-length CEP164 isoform 1 (CGFP-hCEP164-WT), which mimics the mutation p.X1460WextX57 that causes a readthrough of the stop-codon X1460, adding 57 aberrant amino acid residues to the C terminus of CEP164. Similar data were obtained upon CEP164 expression in hTERT-RPE cells (see also Figures S3B–S3D). IMCD3 cells were stably transfected with the respective CEP164 constructs in a retroviral vector for doxycyclin-inducible expression (pRetroX-Tight-Pur). Scale bars, 10 mm. (E–H) Knockdown of Cep164 disrupts ciliary frequency. (E) Depletion of Cep164 by siRNA (F) causes a ciliary defect in 3D spheroid growth assays. IMCD3 cells transfected with either siCtrl or siCep164 were grown to spheroids in 72 hr and immunostained for acetylated tubulin (red). DAPI stains nuclei (blue). Doxycycline induced stably transfected NGFP-hCEP164-WT (green).Space bar represents 5 mm. (G) Nuclei and cilia were scored within a single spheroid to generate ciliary frequencies. siCep164 transfected cells manifest lower cilia frequencies (33%) compared to control transfected IMCD3 cells (49%). Induction of NGFP-hCEP164WT in siCep164 transfected cells rescues this ciliary defect (57%). 50 spheroids per condition were analyzed in three independent experiments. Error bars represent SEM, n = 3, *p value < 0.0002. (H) Ciliary frequency is not rescued by mutant CEP164. Ciliary frequencies are reduced in siCep164 transfected IMCD3 cells (39%) compared to control siCtrl transfected IMCD3 cells (54%). Induction of NGFP-hCEP164Q525X does not rescue this ciliary defect (34%). 50 spheroids per condition were analyzed. Error bars represent SEM, ***p value < 0.0002. See also Figure S3. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 539 NPHP-RC Proteins Colocalize with the DDR Protein TIP60 to Nuclear Foci A noncentrosomal localization for CEP164 was described by demonstrating its translocation to nuclear foci in response to DNA damage (Pan and Lee, 2009; Sivasubramaniam et al., 2008). CEP164 plays a role in DDR signaling where it interacts with the DDR protein ATRIP (Figure S2C), is activated by the DDR proteins ATM and ATR, and is necessary for checkpoint- 1 (Chk1) activation. Abrogation of CEP164 function leads to loss of G2/M cell-cycle checkpoint and aberrant nuclear divisions (Sivasubramaniam et al., 2008). Localization of SDCCAG8 (alias NPHP10) (Otto et al., 2010), shows nuclear foci in hTERT-RPE cells in addition to its centrosomal localization (Figures 4B–4C). Transient shRNA knockdown confirmed specificity of the signal (Figures S4B–S4D). SDCCAG8/NPHP10 did not colocalize with markers for PLM bodies (Janderova´ -Rossmeislova´ et al., 2007) or CENP-C (marking chromosomal centromeres) (Figures S5A and S5B). In contrast, SDCCAG8/NPHP10 fully colocalized with SC35 in hTERT-RPE cells (Figures 4A–4C). SC35, also known as serine/arginine-rich splicing factor 2 (SRSF2), plays a role in DDR by controlling cell fate decisions in response to DNA damaging agents (Edmond et al., 2011; Reinhardt et al., 2011). SC35 marks hubs of enhanced gene expression (Szczerbal and Bridger, 2010), is phosphorylated by topoisomerase I (Elias et al., 2003), and is required for genomic stability during mammalian organogenesis (Xiao et al., 2007). Moreover, ZNF423 also fully colocalizes (Figure 4D), and CEP164 partially colocalizes (Figure 4E) with SC35 in nuclear foci. Consequently, ZNF423 and CEP164 also colocalize with SDCCAG8/NPHP10 in SC35positive nuclear foci (Figures 4F and 4G). SC35 functions within a TIP60 complex, in which TIP60 acetylates SC35 on lysine 52 (Figure S1B), modifying the role of SC35 in the promotion of apoptosis and inhibition of G2/M arrest (Edmond et al., 2011), which is regulated by the checkpoint proteins Chk1 and Chk2 (Figure S1D). Interestingly, the TIP60 protein, together with the heterotrimeric MRN complex (of which MRE11 is a component) constitutes the major activator of ATM within the ATM pathway of DDR signaling (Ciccia and Elledge, 2010) (Figure S1A). In hTERT-RPE cells the ATM activator TIP60 colocalizes to nuclear foci with SC35/SRSF2 (Figure 4H) and partially with the identified NPHP-RC protein CEP164 (Figure 4I). We thus identify a group of NPHP-RC proteins and demonstrate that they colocalize to nuclear foci with the DDR proteins TIP60 and SC35. These gene products include the identified NPHPRC proteins ZNF423 and CEP164 as well as SDCCAG8/ NPHP10. Interestingly, the protein OFD1, which is mutated in the ciliopathy oral-facial-digital syndrome, is part of the TIP60 complex. We recently identified OFD1 as a direct interaction partner of SDCCAG8/NPHP10 (Figure S1B) (Otto et al., 2010). Cep164 Associates with DDR Proteins and Its Loss Causes DDR Defects Because one of the central mechanisms controlled by DDR signaling is cell-cycle regulation through phosphorylation of checkpoint-1 (Chk1) and checkpoint-2 (Chk2) proteins (Figure S1D), we tested whether Chk proteins are recruited to SC35/SRSF2-positive nuclear foci. SC35 and p317-Chk1 colocalize to nuclear foci in hTERT-RPE cells (Figure 4J). We then tested whether localization of CEP164 to nuclear foci was inducible by DNA damage. Following irradiation with 20–50 J/m2 of UV light, CEP164-positive nuclear foci condensed to larger size and colocalized with TIP60 and Chk1 to foci of similar size (Figures 4K–4O). TIP60 and p317-Chk1 colocalized within these foci (Figure 4P). We thus demonstrate that CEP164 translocates in response to DNA damage to nuclear foci that contain the DDR proteins TIP60 and Chk1. Lagging chromosomes on anaphase spindles (‘‘anaphase lag’’) are a hallmark of many mutations that affect mitotic checkpoint integrity. We show that siCep164 knockdown in IMCD3 cells increased anaphase lag from 1% in siCtrl controls to 21% in siCep164-treated cells (Figures 5A and 5B, p = 0.04). This phenomenon was specific, since doxycycline-inducible expression of WT-CEP164 during Cep164 siRNA knockdown reduced the incidence of anaphase lag to just 4% (Figure 5B). These data indicate a requirement for Cep164 at the G2/M checkpoint. The DDR pathway can be activated by the CDK inhibitor roscovitine, which also reduces Chk1 expression (Maude and Enders, 2005). Roscovitine reduces the development of kidney cysts in the Nphp9 mouse model, Jck (Bukanov et al., 2006). We therefore tested the influence of roscovitine (targeting CDK2, 5, 7 and 9) on DDR activation in IMCD3 cells. Immunofluorescence shows increased uniform distribution of gH2AX (activated H2AX phosphorylated at Ser139) in the nucleus of IMCD3 cells upon roscovitine treatment in irradiated cells, indicating partial DDR activation (Figure 5C). Second, in cells treated with roscovitine, UV irradiation caused enhanced gH2AX staining with a prominent nuclear foci pattern, characteristic of strong DDR activation (Figure 5D). Immunoblotting showed that roscovitine decreased the amount of CEP164 present in both control and UV-irradiated cells (Figures 5E and 5F). This was most likely due to translocation of CEP164 into the nucleus upon roscovitine treatment, as shown by subcellular fractionation (Figure 5F). As expected, UV radiation increased phosphorylation of Chk1 at Ser317 (p-Chk1) (Figure 5E), and roscovitine decreased Chk1 protein expression and abrogated UV-induced p-Chk1 in both cytoplasm and nucleus (Figures 5G and 5H). These data indicate that CDK inhibition by roscovitine causes nuclear translocation of CEP164 and inhibits Chk1 activation. gH2AX activation by roscovitine may restore cell-cycle control by Chk2 activation instead (Maude and Enders, 2005). Human Wild-Type CEP164 but Not Its NPHP-RC Truncation Mutant Rescues IMCD3 Cell Proliferation In clonally selected IMCD3 cells expressing wild-type human CEP164 cDNA construct N-GFP-CEP164-WT under doxycycline (Dox) control, depletion of endogenous mouse Cep164 retarded proliferation in comparison to either undepleted control cells or undepleted cells that were Dox-induced to overexpress N-GFP-CEP164-WT alone (Figure 5G). Cep164-depleted growth was rescued by Dox-induced expression of human N-GFPCEP164-WT (Figure 5G). Cells expressing truncated cDNA construct N-GFP-CEP164-Q525X, modeling the NPHP-RC mutation in family F59, exhibited retarded growth, even when the endogenous Cep164 was present (Figure 5H), consistent with a dominant negative effect. Further depletion of the endogenous 540 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Figure 4. Colocalization upon Immunofluorescence of the NPHP-RC Proteins SDCCAG8/NPHP10, ZNF423 and CEP164 to Nuclear Foci that Are Positive for the DDR Signaling Proteins SC35, TIP60 and Chk1 in hTERT-RPE Cells (A–G) Colocalization of NPHP-RC proteins with SC35 in nuclear foci. SDCCAG8/NPHP10 (A-C) and ZNF423 (D) fully colocalize to nuclear foci with SC35, and (E) CEP164 partially colocalizes with SC35. SDCCAG8/ NPHP10 also colocalizes with the identified NPHP-RC proteins ZNF423 (F) and CEP164 (G). (H–J) Colocalization of NPHP-RC proteins with the DDR protein TIP60 and Chk1 to nuclear foci. (H) TIP60 fully colocalizes with SC35. (I) TIP60 partially colocalizes with CEP164. (J) Chk1 fully colocalizes with SC35/ SRSF2. DNA is stained in blue with DAPI. Scale bars, 5 mm. (K–P) Colocalization of DDR and NPHP proteins upon induction of DDR by UV radiation in HeLa cells. (K) Following irradiation of HeLa cells with UV light at 20 J/m2 a strong immunofluorescence signal of an anti-gH2AX antibody indicates activation of DDR. (L–M) Upon irradiation with UV light, CEP164-positive nuclear foci condense and colocalize with TIP60 foci of similar size. (N–O) In untreated cells (N) a pattern of broad CEP164 speckles, which are Chk1-negative and locate to DAPI-negative domains, changes to a pattern of multiple smaller foci (O) that are double positive for both CEP164-N11 and Chk1. (P) p317-Chk1 fully colocalizes with TIP60 to nuclear foci and to the centrosome (arrowhead). See also Figures S4, S5A, and S5B. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 541 Figure 5. Knockdown of Cep164 Causes Anaphase Lag and Retarded Cell Growth (A and B) Knockdown of CEP164 causes anaphase lag. siCep164 knockdown in IMCD3 cells increased anaphase lag incidence from 1% after siCtrl to 21% after siCep164-treated cells (n > 250 anaphases, five independent experiments). CREST antiserum (red) and DAPI (blue) confirmed the presence of incomplete mitotic congression and unattached kinetochores during late anaphase (white arrows). Doxycycline-inducible expression of WT-CEP164 during Cep164 siRNA knockdown reduced the incidence of anaphase lag to 4%, whereas untransfected IMCD3 cells had no detectable anaphase lag (0%) (B). Bars represent SEM, p values (Student’s t test) are indicated above the bar graph. (C–F) The effect of roscovitine on UV-induced DDR. Cells were UV irradiated with 30 J/m2 and analyzed 1 hr after UV irradiation. Where indicated, cells were preincubated for 24 hr with the CDK inhibitor roscovitine (80 mM). (C and D) Immunofluorescence analysis showed that roscovitine triggered uniform nuclear 542 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Cep164 in N-GFP-CEP164-Q525X-expressing cells showed an additive effect on growth retardation, confirming the dominant negative effect of N-GFP-CEP164-Q525X in this experimental system (Figure 5H). CEP164 Directly Interacts with CCDC92 and TTBK2 NPHP-RC proteins are known to interact with other NPHP-RC proteins in the dynamic ‘‘NPHP-JS-MKS interaction network’’ (Sang et al., 2011).To identify novel direct interaction partners of CEP164, we performed yeast two-hybrid screening. We identified CCDC92 and TTBK2 as direct interactors of CEP164 (Figures S5C–S5J). Interactions between CEP164 and both partners were validated by GST pull-down (Figure S5D) and coimmunoprecipitation (Figures S5E–S5H). Immunofluorescence showed that CCDC92 fully colocalizes with CEP164 at the mother centriole (Figures S5I and S5J). CEP164 also interacted with NPHP3 and weakly with NPHP4 (Figures S6A–S6B), demonstrating that CEP164 is in a complex with other known NPHP-RC proteins (Figures S1A and S1B). The DDR protein DDB1 interacted with NPHP2 (Figures S6C and S6D). The disheveled protein (Dvl), which is a central component of the Wnt pathway, interacts with NPHP2/inversin targeting Dvl for proteasomal degradation, thereby triggering a switch from canonical to noncanonical Wnt signaling (Germino, 2005; Simons et al., 2005). We identify interaction between Dvl3 and CEP164 (Figures S6E–S6H). Immunocytochemistry reveals that endogenous Dvl3 and CEP164 share centrosomal localization (Figure S6E). We demonstrate that GST-CEP164 (aa 2–195) is sufficient to pull down endogenous Dvl3 from the cellular lysate (Figure S6F). Domain mapping for Dvl3 suggests that CEP164 interacts with the proline-rich region of Dvl3, because only mutants containing this sequence efficiently coimmunoprecipitate with CEP164-GFP (Figure S6G). Interestingly, only wildtype CEP164-mCherryRFP but not the NPHP-RC causing mutant CEP164-Q525X detected in family F59 (Table 1) can be efficiently immunoprecipitated with Dvl3 (Figure S6H), further supporting its pathogenic role. cep164 Loss of Function Causes NPHP-RC and DDR Activation in Zebrafish To test in a vertebrate model whether loss of cep164 function results in both, an NPHP-RC phenotype as well as DDR activation, we performed cep164 knockdown in zebrafish embryos using morpholino-oligonucleotides (MOs) (Figure 6). A p53 MO was injected to reduce off-target MO effects (Robu et al., 2007). At 28 hr postfertilization (hpf) we observed the ciliopathy phenotypes of ventral body axis curvature and cell death (Figure 6A–6C). Embryos showed increased expression of phosphorylated gH2AX (Figures 6D and 6E). At 48 hpf, cep164 morphants displayed the typical ciliopathy phenotype of abnormal heart looping (Figures 6F–6I). At 72 hpf, embryos developed further NPHP-RC phenotypes, including pronephric tubule cysts (Figures 6J and 6K), hydrocephalus, and retinal dysplasia (Figures 6L–6M). Depletion of CEP164 or ZNF423(Zfp423) Causes Sensitivity to DNA Damaging Agents To assess whether depletion of CEP164 causes sensitivity to DNA damage, Cep164 expression was stably suppressed in the mouse renal cell line IMCD3 (Figures 6P and 6Q). Cep164 knockdown resulted in a dose-dependent increase of gH2AX intensity levelsina FACS analysis,signifyingincreasedradiationsensitivity to IR and perturbed DDR. Cellular sensitivity to IR was also seen in cells depleted of CEP164 using a multicolor competition assay (MCA) (Smogorzewska et al., 2007) (Figures S7A and S7B). To test whether ZNF423(Zfp423) affects DDR, we examined P19 cells, which express high levels of endogenous Zfp423 (Figures 6R–6T). Replicate cultures infected with lentivirus expressing either scrambled control or Zfp423-targeted shRNA were exposed to 0–10 Gy of X-irradiation and imaged for Zfp423 and nuclear gH2AX foci (Figure 6R). Quantification showed significantly increased gH2AX intensities in Zfp423-depleted cells at lower (0.5 and 1.0 Gy) exposures (Figure 6S), but the effect was nonsignificant when corrected for the number of exposures. To determine whether sensitivity to lower dose is reproducible, we exposed 32 additional cultures at 1.0 Gy (Figure 6T). Normalized gH2AX fluorescence in Zfp423 knockdown had both higher mean (9.6 versus 4.7) and median (6.6 versus 5.2) values than control (Figure 6T). These data replicate the radiation sensitivity with high significance (p = 0.018, Mann-Whitney U test, 2 tails), indicating that P19 cells require Zfp423 for quantitatively normal DDR. These results are further confirmed by multicolor competition assays Figure S7C. DISCUSSION Disease Gene Identification Implicates NPHP-RC Proteins in DDR Many DDR signaling proteins localize to nuclear foci and to centrosomes. In addition, dual localization of proteins at distribution of gH2AX (activated H2AX phosphorylated at Ser139) in non-UV irradiated cells suggesting partial DDR activation (C). UV radiation caused enhanced gH2AX staining with a prominent nuclear foci pattern, characteristic of strong DDR activation (D). (C and D) Error bars denote SEM. (E and F) The effect of roscovitine on UV-triggered subcellular localization of CEP164 and Chk1. CEP164 and Chk1 proteins, along with nuclear marker Sam68 and cytoplasmic marker 14.3.3 were analyzed by Western blot. Roscovitine decreased the amount of CEP164 present in control and UV-irradiated cells (E). This was most likely due to translocation of CEP164 into the nucleus upon roscovitine treatment as shown by subcellular fractionation (F). As expected, UV radiation increased phosphorylation of Chk1 at Ser317 (p-Chk1) (E), and roscovitine decreased Chk1 protein expression and abrogated UV-induced p-Chk1 in both cytoplasm and nucleus (E-F). Proteins 14.3.3 and Sam68 serve as controls for cytoplasmic versus nuclear fraction, respectively. See also Figure S6. (G and H) Transient knockdown of Cep164 inhibits proliferation, which is rescued by wild-type but not mutant CEP164. In clonally selected and doxycycline (Dox)inducible mouse IMCD3 cells siRNA knockdown was performed. (G) IMCD3 cells depleted of murine Cep164 grew more slowly (siRNA, green line) than nondepleted cells (control, blue line) or the nondepleted cells induced to express human wild-type CEP164 (Dox, red line). Expression of WT Cep164 in siRNAdepleted cells rescued the slow growth phenotype of Cep164 depletion (siRNA+ dox, purple line). (H) As in (G), except mutant Cep164 cDNA (CEP164-Q525X) was expressed under doxycyclin control. Expression of this allele itself had a negative impact on cell growth (green line), suggesting a dominant negative effect. An even greater negative effect was seen when the endogenous Cep164 was depleted in cells expressing CEP164-Q525X (siRNA+ dox, purple line). The average counts are plotted with standard deviations. Asterisks indicate significant differences by unpaired Student’s t test (p < 0.05). Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 543 Figure 6. Knockdown of cep164 in Zebrafish Embryos Results in Ciliopathy Phenotypes, and Knockdown of Cep164 or Zfp423(Znf423) Causes Sensitivity to DNA Damage A morpholino-oligonucleotide (cep164 MO) targeting the exon 7 splice donor site of zebrafish cep164 was injected into fertilized eggs at the one to four-cells stage together with p53 MO (0.2 mM) to minimize nonspecific MO effects. 544 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. centrosomes and at nuclear foci has been demonstrated for multiple known DDR proteins related to ataxia or CVH. Individuals with mutations in the three NPHP-RC-causing genes that we identify here, MRE11, ZNF423, and CEP164, share the NPHP-RC phenotypes of CVH and ataxia. The first protein that strongly linked DDR signaling to the ataxia phenotype was the protein ATM (ataxia telangiectasia mutated) (Savitsky et al., 1995). Interestingly, we identify here in individuals with NPHPRC, CVH and ataxia, mutations in proteins that colocalize to nuclear foci with TIP60 and/or its interaction partner SC35. These proteins are ZNF423, CEP164, and the previously identified NPHP-RC protein SDCCAG8/NPHP10 (Otto et al., 2010). Our findings support the notion that many products of genes, which if mutated cause NPHP-RC and/or ataxia, play a role in DDR and are part of a dynamic protein complex. A DDR-Based Pathogenic Hypothesis of Dysplasia and Degeneration in NPHP-RC We here generate evidence that NPHP-RC proteins exhibit dual localization at centrosomes and in nuclear foci and that they play a role in DDR. We also demonstrate the parallel occurrence of DDR defects with an NPHP-RC phenotype upon cep164 knockdown in zebrafish. We therefore propose that defects in DDR may participate in the pathogenesis of NPHP-RC. Whereas multiple signaling pathways have been implicated in the pathogenesis of NPHP-RC (Hildebrandt et al., 2011), including noncanonical (Simons et al., 2005) and canonical Wnt signaling (Yu et al., 2009), Shh signaling (Huangfu et al., 2003), and mitotic spindle orientation (Fischer et al., 2006), none of them consistently explains the phenotypes observed. In particular, none of these mechanisms provides a model for the dichotomy of dysplasia phenotypes resulting from null-alleles of NPHP-RC genes versus degenerative phenotypes resulting from hypomorphic alleles of the same NPHP-RC genes. Based on our findings we here propose a pathogenic hypothesis for NPHP-RC that implicates DDR signaling as a relevant disease mechanism. Within this hypothesis, loss of function of NPHP-RC proteins with a dual role in DDR and centrosomal signaling, would cause disturbance of cell-cycle checkpoint control, which is particularly detrimental for embryonic and adult progenitor cell survival. This notion is in accordance with the orthologous mouse model for ZNF423 loss of function, the Zfp423À/À mouse, in which CVH with ataxia is caused by defective granule progenitor proliferation in the cerebellum (Alcaraz et al., 2006). Within this pathogenic hypothesis for NPHP-RC, a DDR signaling defect would lead to impairment of cell-cycle checkpoint control, which in turn would cause lack of progenitor cells. This hypothesis could lend a possible explanation to the following persisting conundrum of NPHP-RC pathogenesis: in certain NPHP genes (e.g., NPHP3, 6, or 8) null mutations cause severe, congenital-onset phenotypes of dysplasia and malformation in kidney, eye, CVH, and liver, whereas hypomorphic mutations in the same gene cause only late-onset degenerative phenotypes such as renal tubular degeneration and fibrosis (nephronophthisis), retinal degeneration (Senior-Loken syndrome), and liver fibrosis. However, the disease mechanisms of neither the degenerative nor the dysplastic phenotypes are understood. These findings suggest that null mutations act during morphogenesis in embryonic development causing dysplasia and malformation, whereas hypomorphic mutations act during the ‘‘chronic’’ processes of tissue maintenance and repair, which are spread out over months or years of the life of an organism. Because DDR signaling is in high demand under conditions of high proliferation during development (morphogenesis), causing high ‘‘replication stress’’ to progenitor cells, tissue dysplasia would be an expected pathogenic outcome. Conversely, during tissue maintenance, low replication stress would be expected, but persistent DDR impairment would allow (A–E) Whereas p53 MO injection (n = 67) did not produce any phenotype (A), coinjection of cep164 MO at 28 hpf caused the mild ciliopathy phenotype of ventral body axis curvature in 48% of embryos (60/125) (B). 50% of embryos (62/125) showed severe cell death throughout the body as judged by gray-appearing cells in the head region (C). Embryos with severe cell death also showed increased expression of phosphorylated gH2AX (D) compared to p53 MO control (E). Most embryos with massive cell death did not survive beyond 48 hpf. (F–I) At 48 hpf, surviving cep164 morphants displayed the ciliopathy phenotype of laterality defects. Whereas p53 MO did not cause any abnormal heart looping (F and G), cep164 MO caused inverted heart looping (H) or ambiguous heart looping (I). (A, atrium; L, left; V, ventricle). (J–M) At 72 hpf, cep164 morphant embryos developed further ciliopathy phenotypes. When compared to p53 MO controls (J), pronephric tubules (arrow heads) exhibited cystic dilation (K), asterisks) in 25% (7/28) of embryos, compared to p53 MO controls (J and L), 0% (0/67) of which showed kidney cysts, hydrocephalus (asterisk), or retinal dysplasia (brackets) (M). (N) At 0.2 mM, cep164 MO knockdown effectively altered mRNA processing as revealed by RT-PCR. The wild-type (WT) mRNA product is 339 bp. A shorter aberrantly spliced mRNA product appeared in cep164 morphants (asterisks), and the normal mRNA product was significantly reduced. p53 MO alone did not affect cep164 mRNA processing. (O) Quantification of g-H2AX levels in cep164 MO morphants. Whole-fish lysates were prepared from morphants injected with control MO (p53 0.2 mM) or cep164 MO (p53 0.2 mM, cep164 0.2 mM). Injection of cep164-targeting MO causes upregulation of g-H2AX in cep164 morphant embryos signifying perturbed DDR. gH2AX levels correlate with the phenotypic severity of the cep164 morphants (see A–C). Anti-a-tubulin antibody was used to show equal loading. (P and Q) Cep164-deficient IMCD3 cells exhibit radiation sensitivity. In IMCD3 cells transduced with shRNA retrovirus, Cep164 expression was suppressed by shRNA knockdown to about 20% of control as judged by qPCR (P). Cep164 knockdown resulted in a dose-dependent increase of gH2AX-positive cells in a FACS assay, signifying increased radiation sensitivity to IR and perturbed DDR. See also Figure S7. In (Q) the level of significance of two-tailed t test (p < 0.001) is indicated by an asterisk. Error bars denote SEM. (R–T) Zfp423(Znf423)-deficient P19 cells exhibit radiation sensitivity. P19 cells transduced with shRNA lentivirus were exposed to the indicated level X-irradiation. Zfp423 and gH2AX immunofluorescence was quantified in matched replicate cultures for each virus 2 hr after irradiation. (R) Representative images illustrate dose-responsiveness of gH2AX and effective knockdown of Zfp423 expression. (S) gH2AX intensity normalized to DAPI+ nuclei is increased following IR at 0.5 and 1.0 Gy, signifying increased IR sensitivity and perturbed DDR (2 fields from each of 6 replicate cultures per condition). Asterisks, uncorrected pair-wise p < 0.05, Mann-Whitney U test, 2 tails. (T) Histogram shows average gH2AX intensity per cell in 16 additional replicate cultures for each shRNA at 1.0 Gy exposure. p = 0.018, Mann-Whitney U test, 2 tails. See also Figure S7. Box plots delimit quartiles in (S). Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 545 slow accumulation of DNA damage with a phenotype of tissue degeneration. At least two related findings support this model: (1) In a mouse conditional knockout model of the cystic kidney disease gene Pkhd1, knockout of the gene before 2 weeks of postnatal life, up to which a high proliferation state prevailed, caused (dysplastic) kidney cysts, whereas knockdown after 2 weeks of postnatal life, when proliferation rate was shown to be dramatically reduced, only caused occasional cysts, the number of which increased when tissue injury was induced (Piontek et al., 2007). This phenomenon could be explained by different degrees of replication stress, and thereby DDR activation, under different proliferation rates. (2) In Seckel syndrome (primordial dwarfism), a progeria syndrome with CVH caused by mutation of the centrosomal and DDR proteins ATR, CEP152, or pericentrin, the degree of cerebellar impairment is dependent on cell proliferation state (Kalay et al., 2011; Murga et al., 2009; Rauch et al., 2008). A DDR-Based Pathogenic Hypothesis Might Explain Specific Organ Involvement in NPHP-RC Regarding the question why organ degeneration occurs in specific organs and at characteristic sites, it is tempting to speculate that the specific tissue regions or cell types affected in NPHP-RC are more strongly exposed to genotoxins. In the kidney, the distal convoluted tubule segment, around which most fibrotic changes occur, is more strongly exposed to genotoxins such as hydroxyurea. Retinal degeneration could be caused by postnatal accumulation of UV light-induced DNA damage. Most strikingly, bile duct-surrounding cholangiocytes in the liver are the one mammalian cell type that is most strongly exposed to genotoxins that are generated by the liver for excretion in bile. In summary, a testable pathogenic hypothesis of NPHP-RC that implicates DDR signaling, impaired cell-cycle checkpoint control with lack of progenitor cells might potentially explain some of the ill understood features of ciliopathies: (1) It might provide a mechanism for the dual phenotypes of degeneration/dysplasia seen in NPHP-RC in kidney, eye, cerebellum and liver. (2) It would implicate in the NPHP-RC pathogenesis, lack of response to replication stress-sensing as a functional basis for understanding the dualism of dysplasia that occurs in high-proliferation states during development/ morphogenesis or repair versus degeneration, which occurs during the low proliferation state of tissue mainte- nance. (3) It would characterize the degenerative phenotypes as diseases of ‘‘organ-specific premature aging,’’ thereby pointing in new directions for identification of small compounds for therapy including cyclin inhibitors. EXPERIMENTAL PROCEDURES Research Subjects We obtained human samples following informed consent from individuals with NPHP-RC. Approval for human subjects research was obtained from the University of Michigan Institutional Review Board and the other institutions involved. The diagnosis of NPHP-RC was based on published clinical criteria (Chaki et al., 2011). Linkage Analysis For genome-wide HM the GeneChip Human Mapping 250k StyI Array from Affymetrix was used. Nonparametric LOD scores were calculated using a modified version of the programs GENEHUNTER 2.1 (Kruglyak et al., 1996; Strauch et al., 2000) and ALLEGRO (Gudbjartsson et al., 2000) in order to identify regions of homozygosity as described (Hildebrandt et al., 2009; Sayer et al., 2006). Bioinformatics Genetic location is according to the February 2009 Human Genome Browser data (http://www.genome.ucsc.eduH). Immunoblotting and Immunoprecipitation Immunoblotting and immunoprecipitation were performed as previously described (Bryja et al., 2007). HEK293 cells were transfected with the indicated constructs and lysed 48 hr later. Samples were analyzed using SDS-PAGE and western blotting, or subjected to immunoprecipitation. The antibodies used for immunoprecipitation are described in Extended Experimental Procedures. cep164 Zebrafish Morpholino Oligo-Mediated Knockdown MOs were obtained from Gene Tools, LLC (Philomath, OR). MOs (cep164 at 0.1 mM, standard control at 0.2 mM, and p53 MO at 0.2 mM) were injected into zebrafish embryos at 1–4 cell stages. Embryos were fixed at 27 hpf with 4% PFA/PBS +1% DMSO overnight, permeablized with acetone at À20 C for 7 min, and stained with antibody against phosphorylated zebrafish gH2AX (1:1,000, gift from Amatruda lab at UT Southwestern) or antibody against cleaved Caspase-3 (1:200, BD Biosciences). Alex568-anti rabbit IgG was used at 1:2,000 and 1:1,000 respectively. The IF procedure followed standard protocol. Morpholinos were: cep164 MO: 50-TATATGCTCTTCTCCATC ACCTCAT; p53 MO: 50-GCGCCATTGCTTTGCAAGAATTG. For histological analyses, embryos were fixed at 72 hpf with 4% PFA/PBS and embedded in JB-4 resin (PolySciences) following the manufacturer’s protocol. Six millimeter sections were obtained using a Leica R2265 microtome and stained with hematoxylin-eosin following published procedures (Zhou et al., 2010). Statistical Analysis Student’s two-tailed nonpaired t tests and normal distribution two-tailed z tests were carried out using pooled standard error and S.D. values to determine the statistical significance of different cohorts. SUPPLEMENTAL INFORMATION Supplemental Information includes Extended Experimental Procedures, seven figures, and two tables and can be found with this article online at http://dx.doi. org/10.1016/j.cell.2012.06.028. ACKNOWLEDGMENTS We are grateful to the study individuals for their contributions. This research was supported by grants from the NIH to F.H. (DK068306, DK090917), B.A.H (NS054871, NS060109), N.K. (HD042601, DK075972, DK072301), and W.Z. (DK091405); the March of Dimes Foundation and the Center for Organogenesis of the University of Michigan to F.H.; the Netherlands Organization for Scientific Research to R.R. (NWO Vidi-91786396) and R.H.G. (NWO Vidi- 917.66.354), the Avenir-INSERM program, the Agence Nationale pour la Recherche, the Union Nationale pour les Aveugles et De´ ficients Visuels, RETINA France, Programme Hospitalier de Recherche National 2007, and the Association Bardet-Biedl, France to H.D. and C.S.; the European Community’s Seventh Framework Programme FP7/2009 under grant agreement no: 241955, SYSCILIA (to R.H.G., G.G.S., N.K., R.R., H.O., C.A.J. and G.W.); the Dutch Kidney Foundation (KJPB09.009 and IP11.58 to H.H.A); the Retina Research Foundation and the National Eye Institute (R01EY018571) to R.C.; 546 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. the NIH to H.W. (F32EY19430); and the DFG (SCHE1562 and SFB832 to B.S.; SFB829 to T.B.; SFB592 to H.O.). R.K.K. is supported by FFB-Canada, CIHR, FRSQ, and Reseau Vision. J.S.A. is supported by the Lundbeck Foundation. V.B. is supported by MSM0021622430 (Ministry of Education, Youth and Sports of the Czech Republic), 204/09/H058, and 204/09/0498 (Czech Science Foundation) and EMBO Installation Grant; I.C. is supported by the programme, Brno PhD Talent of South Moravian Center for International Mobility. S.S. is a laureate of the ‘‘Equipe FRM’’ (DEQ20071210558) and the Agence National de la Recherche (R09087KS, R11012KK). W.Z. is a Carl. W. Gottschalk research scholar of the American Society of Nephrology. J.A.S. is a GlaxoSmithKline clinician scientist. A.S. is supported by the Burroughs Wellcome Fund Career Award for Medical Scientists and the Doris Duke Charitable Foundation Clinical Scientist Development Awards and is a Rita Allen Foundation and an Irma T. Hirschl scholar. F.H. is an Investigator of the Howard Hughes Medical Institute, a Frederick G. L. Huetwell Professor, and a Doris Duke Distinguished Clinical Scientist. Received: September 4, 2011 Revised: February 1, 2012 Accepted: June 25, 2012 Published: August 2, 2012 REFERENCES Alcaraz, W.A., Gold, D.A., Raponi, E., Gent, P.M., Concepcion, D., and Hamilton, B.A. (2006). Zfp423 controls proliferation and differentiation of neural precursors in cerebellar vermis formation. Proc. Natl. Acad. Sci. USA 103, 19424–19429. Alcaraz, W.A., Chen, E., Valdes, P., Kim, E., Lo, Y.H., Vo, J., and Hamilton, B.A. (2011). Modifier genes and non-genetic factors reshape anatomical deficits in Zfp423-deficient mice. Hum. Mol. Genet. 20, 3822–3830. Ansley, S.J., Badano, J.L., Blacque, O.E., Hill, J., Hoskins, B.E., Leitch, C.C., Kim, J.C., Ross, A.J., Eichers, E.R., Teslovich, T.M., et al. (2003). Basal body dysfunction is a likely cause of pleiotropic Bardet-Biedl syndrome. Nature 425, 628–633. Attanasio, M., Uhlenhaut, N.H., Sousa, V.H., O’Toole, J.F., Otto, E., Anlag, K., Klugmann, C., Treier, A.C., Helou, J., Sayer, J.A., et al. (2007). Loss of GLIS2 causes nephronophthisis in humans and mice by increased apoptosis and fibrosis. Nat. Genet. 39, 1018–1024. Blachon, S., Gopalakrishnan, J., Omori, Y., Polyanovsky, A., Church, A., Nicastro, D., Malicki, J., and Avidor-Reiss, T. (2008). Drosophila asterless and vertebrate Cep152 Are orthologs essential for centriole duplication. Genetics 180, 2081–2094. Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007). Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J. Cell Sci. 120, 586–595. Bukanov, N.O., Smith, L.A., Klinger, K.W., Ledbetter, S.R., and IbraghimovBeskrovnaya, O. (2006). Long-lasting arrest of murine polycystic kidney disease with CDK inhibitor roscovitine. Nature 444, 949–952. Chaki, M., Hoefele, J., Allen, S.J., Ramaswami, G., Janssen, S., Bergmann, C., Heckenlively, J.R., Otto, E.A., and Hildebrandt, F. (2011). Genotype-phenotype correlation in 440 patients with NPHP-related ciliopathies. Kidney Int. 80, 1239–1245. Cheng, L.E., Zhang, J., and Reed, R.R. (2007). The transcription factor Zfp423/ OAZ is required for cerebellar development and CNS midline patterning. Dev. Biol. 307, 43–52. Ciccia, A., and Elledge, S.J. (2010). The DNA damage response: making it safe to play with knives. Mol. Cell 40, 179–204. Doxsey, S.J., Stein, P., Evans, L., Calarco, P.D., and Kirschner, M. (1994). Pericentrin, a highly conserved centrosome protein involved in microtubule organization. Cell 76, 639–650. Edmond, V., Moysan, E., Khochbin, S., Matthias, P., Brambilla, C., Brambilla, E., Gazzeri, S., and Eymin, B. (2011). Acetylation and phosphorylation of SRSF2 control cell fate decision in response to cisplatin. EMBO J. 30, 510–523. Elias, E., Lalun, N., Lorenzato, M., Blache, L., Chelidze, P., O’Donohue, M.F., Ploton, D., and Bobichon, H. (2003). Cell-cycle-dependent three-dimensional redistribution of nuclear proteins, P 120, pKi-67, and SC 35 splicing factor, in the presence of the topoisomerase I inhibitor camptothecin. Exp. Cell Res. 291, 176–188. Fischer, E., Legue, E., Doyen, A., Nato, F., Nicolas, J.F., Torres, V., Yaniv, M., and Pontoglio, M. (2006). Defective planar cell polarity in polycystic kidney disease. Nat. Genet. 38, 21–23. Germino, G.G. (2005). Linking cilia to Wnts. Nat. Genet. 37, 455–457. Graser, S., Stierhof, Y.D., Lavoie, S.B., Gassner, O.S., Lamla, S., Le Clech, M., and Nigg, E.A. (2007). Cep164, a novel centriole appendage protein required for primary cilium formation. J. Cell Biol. 179, 321–330. Griffith, E., Walker, S., Martin, C.A., Vagnarelli, P., Stiff, T., Vernay, B., Al Sanna, N., Saggar, A., Hamel, B., Earnshaw, W.C., et al. (2008). Mutations in pericentrin cause Seckel syndrome with defective ATR-dependent DNA damage signaling. Nat. Genet. 40, 232–236. Gudbjartsson, D.F., Jonasson, K., Frigge, M.L., and Kong, A. (2000). Allegro, a new computer program for multipoint linkage analysis. Nat. Genet. 25, 12–13. Hildebrandt, F., Heeringa, S.F., Ru¨ schendorf, F., Attanasio, M., Nu¨ rnberg, G., Becker, C., Seelow, D., Huebner, N., Chernin, G., Vlangos, C.N., et al. (2009). A systematic approach to mapping recessive disease genes in individuals from outbred populations. PLoS Genet. 5, e1000353. Hildebrandt, F., Benzing, T., and Katsanis, N. (2011). Ciliopathies. N. Engl. J. Med. 364, 1533–1543. Huangfu, D., and Anderson, K.V. (2005). Cilia and Hedgehog responsiveness in the mouse. Proc. Natl. Acad. Sci. USA 102, 11325–11330. Huangfu, D., Liu, A., Rakeman, A.S., Murcia, N.S., Niswander, L., and Anderson, K.V. (2003). Hedgehog signalling in the mouse requires intraflagellar transport proteins. Nature 426, 83–87. Janderova´ -Rossmeislova´ , L., Nova´ kova´ , Z., Vlasa´ kova´ , J., Philimonenko, V., Hoza´ k, P., and Hodny´, Z. (2007). PML protein association with specific nucleolar structures differs in normal, tumor and senescent human cells. J. Struct. Biol. 159, 56–70. Kalay, E., Yigit, G., Aslan, Y., Brown, K.E., Pohl, E., Bicknell, L.S., Kayserili, H., Li, Y., Tu¨ ysu¨ z, B., Nu¨ rnberg, G., et al. (2011). CEP152 is a genome maintenance protein disrupted in Seckel syndrome. Nat. Genet. 43, 23–26. Kruglyak, L., Daly, M.J., Reeve-Daly, M.P., and Lander, E.S. (1996). Parametric and nonparametric linkage analysis: a unified multipoint approach. Am. J. Hum. Genet. 58, 1347–1363. Ku, M.C., Stewart, S., and Hata, A. (2003). Poly(ADP-ribose) polymerase 1 interacts with OAZ and regulates BMP-target genes. Biochem. Biophys. Res. Commun. 311, 702–707. Liu, Q., Tan, G., Levenkova, N., Li, T., Pugh, E.N., Jr., Rux, J.J., Speicher, D.W., and Pierce, E.A. (2007). The proteome of the mouse photoreceptor sensory cilium complex. Mol. Cell. Proteomics 6, 1299–1317. Maude, S.L., and Enders, G.H. (2005). Cdk inhibition in human cells compromises chk1 function and activates a DNA damage response. Cancer Res. 65, 780–786. Murga, M., Bunting, S., Montan˜ a, M.F., Soria, R., Mulero, F., Can˜ amero, M., Lee, Y., McKinnon, P.J., Nussenzweig, A., and Fernandez-Capetillo, O. (2009). A mouse model of ATR-Seckel shows embryonic replicative stress and accelerated aging. Nat. Genet. 41, 891–898. Ng, S.B., Turner, E.H., Robertson, P.D., Flygare, S.D., Bigham, A.W., Lee, C., Shaffer, T., Wong, M., Bhattacharjee, A., Eichler, E.E., et al. (2009). Targeted capture and massively parallel sequencing of 12 human exomes. Nature 461, 272–276. O’Driscoll, M., Ruiz-Perez, V.L., Woods, C.G., Jeggo, P.A., and Goodship, J.A. (2003). A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nat. Genet. 33, 497–501. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. 547 Otto, E.A., Hurd, T.W., Airik, R., Chaki, M., Zhou, W., Stoetzel, C., Patil, S.B., Levy, S., Ghosh, A.K., Murga-Zamalloa, C.A., et al. (2010). Candidate exome capture identifies mutation of SDCCAG8 as the cause of a retinal-renal ciliopathy. Nat. Genet. 42, 840–850. Otto, E.A., Ramaswami, G., Janssen, S., Chaki, M., Allen, S.J., Zhou, W., Airik, R., Hurd, T.W., Ghosh, A.K., Wolf, M.T., et al; GPN Study Group. (2011). Mutation analysis of 18 nephronophthisis associated ciliopathy disease genes using a DNA pooling and next generation sequencing strategy. J. Med. Genet. 48, 105–116. Pan, Y.R., and Lee, E.Y. (2009). UV-dependent interaction between Cep164 and XPA mediates localization of Cep164 at sites of DNA damage and UV sensitivity. Cell Cycle 8, 655–664. Piontek, K., Menezes, L.F., Garcia-Gonzalez, M.A., Huso, D.L., and Germino, G.G. (2007). A critical developmental switch defines the kinetics of kidney cyst formation after loss of Pkd1. Nat. Med. 13, 1490–1495. Rauch, A., Thiel, C.T., Schindler, D., Wick, U., Crow, Y.J., Ekici, A.B., van Essen, A.J., Goecke, T.O., Al-Gazali, L., Chrzanowska, K.H., et al. (2008). Mutations in the pericentrin (PCNT) gene cause primordial dwarfism. Science 319, 816–819. Reinhardt, H.C., Cannell, I.G., Morandell, S., and Yaffe, M.B. (2011). Is posttranscriptional stabilization, splicing and translation of selective mRNAs a key to the DNA damage response? Cell Cycle 10, 23–27. Robu, M.E., Larson, J.D., Nasevicius, A., Beiraghi, S., Brenner, C., Farber, S.A., and Ekker, S.C. (2007). p53 activation by knockdown technologies. PLoS Genet. 3, e78. Sang, L., Miller, J.J., Corbit, K.C., Giles, R.H., Brauer, M.J., Otto, E.A., Baye, L.M., Wen, X., Scales, S.J., Kwong, M., et al. (2011). Mapping the NPHPJBTS-MKS protein network reveals ciliopathy disease genes and pathways. Cell 145, 513–528. Savitsky, K., Bar-Shira, A., Gilad, S., Rotman, G., Ziv, Y., Vanagaite, L., Tagle, D.A., Smith, S., Uziel, T., Sfez, S., et al. (1995). A single ataxia telangiectasia gene with a product similar to PI-3 kinase. Science 268, 1749–1753. Sayer, J.A., Otto, E.A., O’Toole, J.F., Nurnberg, G., Kennedy, M.A., Becker, C., Hennies, H.C., Helou, J., Attanasio, M., Fausett, B.V., et al. (2006). The centrosomal protein nephrocystin-6 is mutated in Joubert syndrome and activates transcription factor ATF4. Nat. Genet. 38, 674–681. Scha¨ fer, T., Pu¨ tz, M., Lienkamp, S., Ganner, A., Bergbreiter, A., Ramachandran, H., Gieloff, V., Gerner, M., Mattonet, C., Czarnecki, P.G., et al. (2008). Genetic and physical interaction between the NPHP5 and NPHP6 gene products. Hum. Mol. Genet. 17, 3655–3662. Simons, M., Gloy, J., Ganner, A., Bullerkotte, A., Bashkurov, M., Kro¨ nig, C., Schermer, B., Benzing, T., Cabello, O.A., Jenny, A., et al. (2005). Inversin, the gene product mutated in nephronophthisis type II, functions as a molecular switch between Wnt signaling pathways. Nat. Genet. 37, 537–543. Sivasubramaniam, S., Sun, X., Pan, Y.R., Wang, S., and Lee, E.Y. (2008). Cep164 is a mediator protein required for the maintenance of genomic stability through modulation of MDC1, RPA, and CHK1. Genes Dev. 22, 587–600. Smogorzewska, A., Matsuoka, S., Vinciguerra, P., McDonald, E.R., 3rd, Hurov, K.E., Luo, J., Ballif, B.A., Gygi, S.P., Hofmann, K., D’Andrea, A.D., and Elledge, S.J. (2007). Identification of the FANCI protein, a monoubiquitinated FANCD2 paralog required for DNA repair. Cell 129, 289–301. Stewart, G.S., Maser, R.S., Stankovic, T., Bressan, D.A., Kaplan, M.I., Jaspers, N.G., Raams, A., Byrd, P.J., Petrini, J.H., and Taylor, A.M. (1999). The DNA double-strand break repair gene hMRE11 is mutated in individuals with an ataxia-telangiectasia-like disorder. Cell 99, 577–587. Strauch, K., Fimmers, R., Kurz, T., Deichmann, K.A., Wienker, T.F., and Baur, M.P. (2000). Parametric and nonparametric multipoint linkage analysis with imprinting and two-locus-trait models: application to mite sensitization. Am. J. Hum. Genet. 66, 1945–1957. Szczerbal, I., and Bridger, J.M. (2010). Association of adipogenic genes with SC-35 domains during porcine adipogenesis. Chromosome Res. 18, 887–895. Valente, E.M., Silhavy, J.L., Brancati, F., Barrano, G., Krishnaswami, S.R., Castori, M., Lancaster, M.A., Boltshauser, E., Boccone, L., Al-Gazali, L., et al; International Joubert Syndrome Related Disorders Study Group. (2006). Mutations in CEP290, which encodes a centrosomal protein, cause pleiotropic forms of Joubert syndrome. Nat. Genet. 38, 623–625. Xiao, R., Sun, Y., Ding, J.H., Lin, S., Rose, D.W., Rosenfeld, M.G., Fu, X.D., and Li, X. (2007). Splicing regulator SC35 is essential for genomic stability and cell proliferation during mammalian organogenesis. Mol. Cell. Biol. 27, 5393–5402. Yu, J., Carroll, T.J., Rajagopal, J., Kobayashi, A., Ren, Q., and McMahon, A.P. (2009). A Wnt7b-dependent pathway regulates the orientation of epithelial cell division and establishes the cortico-medullary axis of the mammalian kidney. Development 136, 161–171. 548 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Supplemental Information EXTENDED EXPERIMENTAL PROCEDURES Whole-Exome Sequencing Whole-Exome Capture Exome enrichment was conducted largely following the manufacturer’s protocol for the NimbleGen SeqCap EZ Exome v2 (Roche NimbleGen, version 2). Briefly, three micrograms of genomic DNA was fragmented by sonication using the Covaris S2 system to achieve a uniform distribution of fragments with a mean size of 300 bp. The sonicated DNA was purified using AMPure XP Solid Phase Reversible Immobilization paramagnetic (SPRI) beads (Agencourt) followed by polishing of the DNA ends by removing the 30 overhangs and filling in the 50 overhangs resulting from sonication using T4 DNA polymerase and Klenow fragment (New England Biolabs). Following end polishing, a single ‘A’-base was added to the 30 end of the DNA fragments using Klenow fragment (30 to 50 exo minus). This prepares the DNA fragments for ligation to specialized adaptors that have a ‘T’-base overhang at their 30 ends. The end-repaired DNA with a single ‘A’-base overhang was ligated to paired-end adaptors (Illumina) in a standard ligation reaction using T4 DNA ligase and 2–4 mM final adaptor concentration, depending on the DNA yield following purification after the addition of the ‘A’-base (a 10-fold molar excess of adaptors is used in each reaction). Following ligation, the samples were purified using SPRI beads, amplified by six cycles of PCR to maintain complexity and avoid bias due to amplification and quality controlled by library size assessment on the Agilent Bioanalyzer and quantitation using PicoGreen reagent (Invitrogen). One microgram of amplified, purified DNA (DNA library) was prepared for hybridization by adding COT1 DNA and blocking oligonucleotides to the DNA library, desiccating the DNA completely and resuspending the material in NimbleGen hybridization buffer. The resuspended material was denatured at 95 C prior to addition of the exome capture library bait material. The DNA library and biotinlabeled capture library were hybridized by incubation at 47 C for 68 hr. Following hybridization, streptavidin coated magnetic beads were used to purify the DNA:DNA hybrids formed between the capture library and sequencing library during hybridization. The purified sequencing library was amplified directly from the purification beads using 8 cycles of PCR using Pfx DNA polymerase (Invitrogen). The libraries were purified following amplification and the library size was assessed using the Agilent Bioanalyzer. A single peak between 350-400 bp indicates a properly constructed and amplified library ready for sequencing. Final quantitation of the library was performed using the Kapa Biosciences Real-time PCR assay and appropriate amounts loaded onto the Illumina flowcell for sequencing by paired-end 50 nt sequencing on the Illumina HiSeq2000 sequencer. Sequencing Sequencing was performed largely as described in (Bentley et al., 2008). Following dilution to 10 nM final concentration based on the real-time PCR and bioanalyzer results, the final library stock is then used in paired-end (PE) cluster generation at a final concentration of 6-8 pM to achieve a cluster density of 600,000/mm2 (on the Illumina HiSeq2000 instrument, v2.5 reagents). Following cluster generation, 100nt paired-end sequencing was performed using the standard Illumina protocols. Sequencing each sample in a single lane at paired-end 100 nt conditions on the Illumina HiSeq (v2.5 reagents) generated an average of 19.1Gb of pass-filter sequence data with 94.96% aligning to the genome. This amount of sequence corresponded to an average coverage of 181x over the 44 Mb of capture region. 80.41% of sequencing reads fell with 200 bp of the target region. 98.39% of the capture region was sequenced to a depth of at least 1x, 95.6% to at least 8x, 92.17% to at least 20x, and 89.42% to at least 30x. Data Analysis Raw sequencing data for each individual were mapped to the human reference genome (build hg19) using the Burrows-Wheeler Aligner (BWA v0.5.81536)14. The BWA aligned sequencing reads were processed by Picard (http://picard.sourceforge.net/) to label the PCR duplicates. The Genome Analysis Toolkit (GATK, version 5091) was then used to remove duplicates, perform local realignment and map quality score recalibration to produce a ‘‘cleaned’’ BAM file for each individual. SNP calls were made by the Unified Genotyper module in GATK using the ‘‘cleaned’’ BAM files in batch fashion (90 samples per batch). The resulting Variant Call Format (VCF, version 4.0) files were annotated using the GenomicAnnotator module in GATK to identify and label the called variants that are within the targeted coding regions and overlap with known and likely benign SNPs reported in dbSNP v132 (ftp://ftp.ncbi.nlm.nih. gov/snp/organisms/human_9606/VCF/v4.0/00-All.vcf.gz) with allele frequencies > 5%. The annotated VCF files were then filtered using the GATK variant filter module with a hard filter setting and a custom script for initial filtering. Variant calls that failed to pass the following filters were eliminated from the call set: (1) MQ0 > = 4 && ((MQ0 / (1.0 * DP)) > 0.1); (2) QUAL < 30.0 jj QD < 5.0 jj HRun > 5 jj SB > 0.00; (3) Cluster size 10; (4) not contain dbSNP id; (5) inside the targeted regions. Combined VCF files were then split into individual files and remaining variants were tested for recessive segregation in the respective families. Topline statistics on 32 whole human exomes that we evaluated, which include the exomes that we evaluated in this paper, are available from the authors (unpublished data). For each of the three families with mutation of MRE11, ZNF423 or CEP164 there was only one gene in each family in any of the homozygous regions for which the homozygous variant passed Polyphen 1 and 2 scores, segregated within the family, and was absent from >270 controls, thus validating the pathogenic relevance of the mutation (unpublished data). CEP164 Mutation Analysis We performed exon-PCR and Sanger sequencing of all 31 coding exons for one affected individual in each of 856 different NPHP-RC families. The number of families examined (by increasing number of organs involved) was: isolated nephronophthisis (5), isolated retinal degeneration (100), Senior-Loken syndrome (168), Joubert syndrome (240), Bardet-Biedl syndrome (195), Meckel syndrome Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S1 (52), and other severe unsolved NPHP-RC cases (96). We also examined 480 individuals from additional NPHP-RC families using massively-parallel sequencing of exon-PCR from pooled DNA samples (Otto et al., 2010). Yeast-Two-Hybrid Screening with ZNF423 A subclone of human CEP290 was prepared using high fidelity Taq polymerase, spanning 1770 bp, (562 amino acids, 1917-2479) and including wild-type stop codon and cloned into pENTR-TOPO as previously described in (Sayer et al., 2006) (clone named ‘JAS2’). The CEP290 subclone insert was switched to destination vector pDEST32 (binding domain containing yeast-2-hydrid vector, ‘‘bait’’) (Invitrogen). CEP290 (JAS2) was used as bait, fused to the GAL4 DNA binding domain in the pDEST32 vector, and a human fetal brain expression library was screened and cloned into pEXPAD22 GAL4 activation domain fusion vector (Invitrogen). Approximately 1 3 106 clones were screened after cotransforming plasmids into competent MaV203 yeast cells (lithium acetate method) and plating onto His, Leu and Trp deficient medium containing 25 mM 3-aminotriazole. Colonies were replica plated on restrictive media and surviving colonies were used for cDNA extraction. Five ml cultures were grown at 30 C overnight. cDNA was extracted using RPM yeast plasmid isolation kitÔ (Bio 101 systems). cDNA was transformed into E. coli, purified and directly sequenced using vector specific primers. Sequence analysis allowed prediction of amino acid sequences (ORF Finder), which were then identified by BLAT analysis (http://genome.ucsc.edu). Direct yeast-2-hybrid interaction experiments allowed colony growth to be compared to 2 negative controls (respective plasmids without insert, data not shown) and 5 positive control yeast strains for different interaction strength. Coimmunoprecipitation with ZNF423 Full length human ZNF423 was subcloned into pcDNA3.1-NT-GFP using long range PCR (Expand, Roche) and IMAGE clone 100072717 as template. Clones were sequenced to confirm orientation and fidelity. A partial length human CEP290 cDNA spanning 1770 bp as described above (JAS2) was subcloned into a DEST-V5 vector. HEK293 cells were transiently cotransfected with full length GFP-ZNF423 and partial length CEP290-V5 using Lipofectamine 2000. After 24 hr cells were washed in phosphate-buffered saline (PBS), pelleted, and lysed in NP-40 buffer (150 mM sodium chloride, 0.5% NP-40, 50 mM Tris pH 7.4, phenylmethanesulphonylfluoride, and protease inhibitors). The lysate was centrifuged for 20 min at 10,000 g at 4 C and the supernatant was precleared with protein G sepharose beads (GE Healthcare, Giles, Buckinghamshire, UK). After removal of protein G the supernatant was then incubated overnight with protein G and 1 mg of either anti-V5 (Invitrogen) or mouse IgG (Sigma) at 4 C. The beads were washed extensively in lysis buffer and bound proteins resolved by 10% SDS-polyacrylamide gel electrophoresis as described above. GFP-tagged proteins were detected with anti-GFP-conjugated to HRP (Santa Cruz) 1:5,000 and V5-tagged proteins were detected with anti-V5 conjugated to HRP (Invitrogen) 1:5,000. Spheroid Assays IMCD3 Cell Culture IMCD3 is a mouse inner medullary collecting duct cell line. The cells were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM):F12 (1:1) (GlutaMAX, GIBCO), supplemented with 10% Fetal Calf Serum (FCS) and penicillin and streptomycin (1% P/S). Cells were incubated at 37 C in 5% carbon dioxide (CO2) to approximately 90% confluence. Transfections Before exposing the cells to different experimental conditions, cells were seeded in 6-, 24- or 48-well plates, depending on the experimental setup. After at least 6 hr, the cells were transfected with Lipofectamine RNAimax (Invitrogen, 13778-075), according to the supplier’s protocol. Opti-MEM (Invitrogen, 31985-062) was used to dilute the ON-TARGETplus siRNA SMARTpools (Thermo Scientific Dharmacon) for Nontargeting pool (D-001810-10) or Cep164 (L-057068-01). IMCD3 Spheroid Growth Assay Cells were trypsinized 24 hr post-transfection and resuspended cells were then mixed 1:1 with growth factor-depleted matrigel (BD Bioscience). After the matrigel polymerized for 20 min at 37 C, warm medium was dropped over the matrix until just covered. The IMCD3 cells formed spheroids with cleared lumens 3 days later. Medium was removed by pipetting and the gels were washed three times for 10 min with warm PBS supplemented with calcium and magnesium. The gels were then fixed in fresh 4% PFA for 30 min RT. After washing three times in PBS after fixation, the cells were permeabilized for 15 min in gelatin dissolved in warm PBS (350 mg/50 mL) and 0.5% Triton X-100. Primary antibody (mouse anti-acetylated tubulin, Sigma, at 1:20,000) was diluted in the permeabilization gelatin buffer and incubated at 4 C overnight. After washing the spheroids 3 times for 30 min in permeabilization buffer, goat anti-mouse Alexa-555 secondary antibody (Invitrogen) was diluted 1:500 in permeabilization buffer and incubated overnight at 4 C. The next day, spheroids were washed 3 times in permeabilization buffer and mounted with Dapi (1:2,000) in Fluoromount-G (Cell Lab, Beckman Coulter). Images were taken with a Zeiss LSM510 confocal microscope and 50 spheroids per condition were scored. GraphPad Prism 5.0 was used to perform two-tailed Student’s t tests. RT-qPCR for Cep164 IMCD3 cells were transfected with nontargeting siControl or siCep164 oligonucleotides (Dharmacon, ON-TARGETplus SMARTpool Cep164) using Lipofectamine RNAimax (Invitrogen) in 6 well plates seeded with cells at 50% confluency. Cells were lysed 24, 48 and 72 hr after transfection and total RNA was isolated (RNeasy Mini Kit, QIAGEN, 74106) and measured (NanoDrop spectrophotometer ND-1000, Thermo Fischer Scientific Inc.). cDNA was synthesized from 500 ng RNA template using the iScript cDNA Synthesis Kit (Bio-Rad, 170-8891) according to the supplier’s protocol. Dilutions were made for RT-QPCR analysis to determine the expression of Cep164, normalized against reference gene RPL27. The primers (Sigma) used are mCep164 forward 50 -AGAGTGACAACCA S2 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. GAGTGTCC, mCep164 reverse 50 -GGAGACTCCTCGTACTCAAAGTT, mRPL27 forward 50 - CGCCCTCCTTTCCTTTCTGC and mRPL27 reverse 50 -GGTGCCATCGTCAATGTTCTTC. The iQ SYBR Green Supermix (Bio-Rad, 170-8880) was used to multiply and measure the cDNA with a CFX96 Touch Real-Time PCR Detection System (Bio-Rad). All samples were run in triplicate in 20 ml reactions. The following PCR program was used: 95 C for 3 min, followed by 40 cycles of 10 s at 95 C, 30 s at 51.5/61 C and 30 s at 72 C, then 10 s at 95 C followed by a melt of the product from 65 C-95 C.The DDCT method was used for statistical analysis to determine gene expression levels. Immunofluorescence and Confocal Microscopy For immunostaining IMCD3 cells, cells were grown on coverslips and fixed for 30 min in 4%PFA followed by a 15 min permeabilization step in 0.5%Triton X-100/1%BSA/PBS. Primary antibody incubations (human anti-CREST at 1:1,000, mouse anti-acetylated tubulin at 1:20,000) were performed overnight in 1% BSA/PBS. Secondary antibody incubations were performed for 1 hr at RT. DAPI incubations were performed for 10 min at RT. Coverslips were mounted in Fluoromount G (Cell Lab, Beckman Coulter). Confocal imaging was performed using Zeiss LSM510 Confocal laser microscope and images were processed with the LSM software. Approximately 250 events per condition were scored. GraphPad Prism 5.0 was used to perform two-tailed Student’s t tests. Cell-Cycle Studies Generation of IMCD3 and hTERT-RPE Cell Lines that Are Doxycycline-Inducible for N-GFP-CEP164 Constructs Both, IMCD3 and hTERT-RPE cells were stably transfected with N terminally GFP tagged human CEP164 constructs in a retroviral vector for doxycycline (Dox)-inducible expression (pRetroX-Tight-Pur). Following selection on puromycin for 2 weeks, cells were induced with 10 ng/ml of doxycyclin for 20 hr. Based on the GFP expression levels upon immunofluorescence, clonal cells were generated for IMCD3 cells. Cell-Cycle Studies IMCD3 cells Dox-inducible for human N-GFP-Cep164-WT (wild-type) or N-GFP-Cep164-Q525X (mutant) were transfected with either nontargeting control siRNA (50 nM, Dharmacon D-001206-14-20) or mouse Cep164 siRNA (50 nM, Smartpool Dharmacon L-057068-01-0020) using Polyplus transfection reagents. Cells were then re-plated and treated for double thymidine (2 mM) blocks beginning at 24 hr post transfection. At the same time, cells were also induced with doxycycline (10 ng/ml) for N-GFP-Cep164-WT or N-GFP-Cep164-Q525X. Cells were released from second thymidine block for 6 hr and fixed with 2% PFA and stained with PI/RNase staining solution (BD Biosciences). FACS analysis for cell-cycle histograms was performed and data were then analyzed using Modfit software. Mean and SD of % DNA amount for different phases (triplicate samples) were calculated and plotted as bar diagram. Cell-Proliferation Studies IMCD3 cells Dox-inducibly expressing human N-GFP-CEP164-WT or N-GFP-CEP164-Q525X were transfected with either nontargeting control siRNA (50 nM, Dharmacon D-001206-14-20) or mouse Cep164 siRNA (50 nM, Smartpool Dharmacon L-057068-01-0020) using Polyplus transfection reagents. Cells were then re-plated in triplicate of 10,000 cells for each group and treated for thymidine (2 mM) blocks beginning at 24 hr post transfection. At the same time, cells were also induced with doxycycline (10 ng/ml) for N-GFP-CEP164-WT or N-GFP-CEP164-Q525X expression. Cells were counted at 48, 72, and 96 hr post transfection. Roscovitine Studies Cell Culture and Cell Irradiation Sh-Neg-IMCD3 cells were maintained in DMEM:F12 supplemented with 10% FCS, 1% Penicillin/Streptomycin and 5 mg/ml Puromycin (all from GIBCO/Invitrogen, Carlsbad, CA) and maintained in a 5% CO2 humidified atmosphere at 37 C. Roscovitine (Chem Partners, China) was incubated at 80 mM for 24 hr. Ultraviolet (UV) irradiation was performed using a UV StratalinkerÔ (Stratagene/Agilent, Santa Clara, CA) at 30 J/m2 dose; cells were left to recover for 1 hr at 37 C before further analyses. Immunoblotting, Immunofluorescence, and Subcellular Fractionation Cells were lysed as previously described (Bukanov et al., 2006). Subcellular fractionation was performed using Nuclear Extract Kit according to the manufacturer’s instructions (Active Motif, Carlsbad, CA). Protein concentration was determined by BCA protein assay (Pierce/Thermofisher, Rockford, IL). Proteins were resolved on SDS PAGE 5%–12% gradient and transferred using iBlot system (Invitrogen). The following antibodies were used: CEP164 (Novus Biologicals, Littleton, CO), Phospho-Chk1 (Ser317) and phospho-H2AX (Ser139) (Cell Signaling, Danvers, MA), Chk1 (Upstate/Millipore, Billerica, MA), H2AX and 13.3.3 (AbCam, Cambridge, MA) and Sam-68 (Santa Cruz Biotech. Santa Cruz, CA). Primary antibodies were detected with anti-rabbit or anti-mouse horseradish peroxidase-conjugated secondary antibodies (Promega, Madison, WI) and revealed by ECL (Amersham, Little Chalfont, Buckinghamshire, UK) as previously described (Bukanov et al., 2006). For immunofluorescence, cells were fixed for 20 min in 4% paraformaldehyde (Electron Microscopy Science, Hatfield, PA) in PBS (pH 7.2) and were permeabilized with 0.1% Triton X-100 in PBS. Immunofluorescence was then performed as previously described (Smith et al., 2006). Gene Expression Analysis Gene expression analysis was performed as described before (Smith et al., 2006, 17-pp2821-2831) using TAQ-Man gene expression assays (Applied Biosystems/Invitrogen, Carslbad, CA). Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S3 Interaction Studies of CEP164 with TTBK2 and CCDC92 Yeast Two-Hybrid A GAL4-based yeast two-hybrid system was used to screen for binary CEP164 interactors by using methods described by Letteboer and Roepman (Letteboer and Roepman, 2008). Bait constructs expressing full length CEP164 (CEP164fl ) and several fragments thereof (CEP1641-550 , CEP164551-1100 and CEP1641101-1460 ) were used to screen two retinal cDNA libraries: a bovine library of randomly primed retina cDNA and a human library of oligo-dT primed retina cDNA (Letteboer and Roepman, 2008). GST Pull-Down Full length 3xFlag-CCDC92 and 3xFlag-TTBK2 were expressed in COS-1 cells. CEP164fl , CEP1641-550 , CEP164551-1100 and CEP1641101-1640 were cloned in pDEST15 and the resulting expression constructs were transformed in BL21DE3 to express glutathione S-transerase (GST) fusion proteins. The GST pull-down was performed as described by Coene et al. (Coene et al., 2011). The results were assessed by immunoblotting, and Flag-tagged proteins were detected using anti-Flag primary antibodies (Sigma-Aldrich) and goat-anti-mouse IRDye800 secondary antibodies (Rockland, Immunochemicals, Gilbertsville, PA, USA). A Li-Cor Odyssey 2.1 scanner was used for imaging. Coimmunoprecipitation 3xHA-CEP164fl was expressed in combination with either 3xFlag-CCDC92fl or 3xFlag-TTBK2fl in COS-1 cells (and vice versa with swapped tags). 3xFlag-dNp63 or 3xHA-dNp63 was used as a control for specificity. The coimmunoprecipitation was performed as described by Coene et al. Transfected COS-1 cells were lysed and incubated with a-M2-agarose from mouse (Sigma-Aldrich) or a-HA affinity matrix from rat (Roche) for 2 hr at 4 C. After washing, sample buffer was added to the beads and the sample was heated. Beads were then precipitated by centrifugation and the supernatant was analyzed by immunoblotting to assess if TTKB2 and CCDC92 indeed coimmunoprecipitated with CEP164. Interaction Studies of CEP164 with DVL3 Immunoprecipitation Coupled to Mass Spectrometry Mouse embryonal fibroblasts were grown in DMEM supplemented with 10% fetal bovine serum at 37 C to confluency and cell pellets are frozen at À80 C. Cells were then lysed in lysis buffer (0.5% NP40; 150mM NaCl; 50 mM Tris pH 7.4) and subjected to immunoprecipitation with antibodies against Dvl3 (sc-8027); Dvl2 (sc-8026) or control IgG (sc-2025; Santa Cruz Biotechnology). Proteomics grade trypsin (Sigma) was added to the beads directly (10 ng/ml) and left at 37 C for 12 hr. Tryptic digested peptides were desalted using ZipTip C18 column (Millipore) based on the manufacturer’s protocol. LC-MS/MS was performed on a NanoAcquity UPLC (Waters) on-line coupled to an ESI Q-TOF Premier (Waters) mass spectrometer. Trypsin digested peptides eluted from ZipTip column were diluted in autoclaved M.Q. water (Millipore) and loaded onto a 180 mm x 20 mm nanoAcquity UPLC Symmetry trap column (Waters) packed with 5 mm BEH C-18 beads. After 1 min of trapping, peptides were eluted through a 75 mm x 150 mm nanoAcquity (Waters) analytical column packed with 1.7 mm BEH C-18 beads at a flow rate of 400 nL/min using a gradient of 3 – 40% acetonitril with 0.1% formic acid for 35 min at a temperature of 35 C. Effluent was directly fed into the ESI source of the mass spectrometer. Raw data was acquired in data independent MSE Identity (Waters) mode. Precursor ion spectra were acquired with collision energy 5 V and fragment ion spectra with a collision energy 20-35 V ramp in alternating 1 s scans. Raw data was then subjected to a database search using species specific Uniprot and NCBI mouse protein database by the PLGS2.3 software (Waters). Acetyl N-terminal, Deamidation N and Q, Carbamidomethyl C and Oxidation M were set as variable modifications. Peptide accuracy and MS/MS fragment mass accuracy was less than 20 ppm. Western Blotting and Immunoprecipitation Immunoblotting and sample preparations were performed as previously described (Bryja et al., 2007b). HEK293 cells were transfected with the indicated constructs and lysed 48 hr later. Samples were analyzed using SDS-PAGE and Western blotting, or subjected to immunoprecipitation. The antibodies used include the following: Dvl3 (sc-8027), Dvl2 (sc-8026), mouse IgG (sc-2025), from Santa Cruz Biotech, CEP164 (4533.00.02) from Sdix, HA.11 (MMS-101P) from Covance, GFP (3H9), RFP (3F5) from Chromotek and FLAG M2 (F1804), GST (G1160) from Sigma. Immunoprecipitation was performed as previously described (Bryja et al., 2007a). The antibodies used for immunoprecipitation were RFP Trap from Chromotek, HA.11 (MMS-101P) from Covance, anti FLAG (F1804; Sigma), anti Dvl3 (sc-8027), anti Dvl2 (sc- 8026) (all Santa Cruz Biotechnology). GST Pull-Down Production of recombinant GST-tagged fragments of CEP164 (Sivasubramaniam et al., 2008) was induced by adding 0.2 mM IPTG and grown for 4 hr at 37 C. The bacteria were spun down at 4000 g/4 C/10 min and the pellet was resuspended in 10 ml of GST lysis buffer (50 mM Tris-Cl pH 7.5, 150 mM NaCl, 5mM MgCl2, protease inhibitors (Roche)) and stored at À80 C. Then the solution was thawed, sonicated for 3x30 s and spun down 15000 g/4 C/15 min. The supernatant was then used for incubation with GST beads for 2 hr at rotator (100 ml of beads per 100 ml of original bacterial culture). After incubation beads were washed 3 times with 1 ml of GST lysis buffer and frozen at À80 C as 25% slurry in GST lysis buffer +20% glycerol. HEK293T cells were transfected according to the scheme, grown for 24 or 48 hr and lysed in 0.5% NP 40 lysis buffer (0.5% NP40, 50 mM Tris pH 7.4, 1 mM EDTA, 150 mM NaCl) with added protease inhibitors (Roche) and phosphatase inhibitors (Calbiochem). The lysate was spun down at 16100 g/4 C/10 min and the supernatant was used for overnight incubation with 15 ml of solid GST beads containing GST recombinant proteins on the rotator. After the incubation samples were washed with 800 ml 0.5% NP 40 lysis buffer, S4 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. the GST beads were collected at 0.1 g/4 C/1 min, the supernatant was aspirated out and this washing was repeated 6 times. Proteins were eluted with 45 ml of 2x Laemmli buffer. Immunohistochemistry HEK293 cells were seeded at approx. 2x105 cells/well on collagen coated coverslips in 24-well plates. Cells were fixed in fresh 4% paraformaldehyde, permeabilized with 0.05% Triton X-100, blocked with PBTA (3% BSA, 0,25% Triton, 0,01% NaN3) for 1 hr and incubated overnight with primary antibodies CEP164 (4533.00.02) from Sdix and Dvl2 (sc-8026 from Santa Cruz Biotechnology). Next day, coverslips were washed in PBS, and incubated with secondary antibodies: Alexa 488, Alexa 568 (Invitrogen), washed with PBS, stained with DAPI (1:5000) and mounted on coverslips. Cells were visualized using a Leica TCS SP-5 confocal microscope. Multicolor Competition Assay MCA was performed as described previously (Smogorzewska et al., 2007). Briefly, 1.7 3 105 U2OS-GFP were reverse transfected with 64 pmol siRNA duplex using Lipofectamine RNAiMAX (Invitrogen) as per manufactures instructions. 48 hr post transfection siRNA treated U2OS-GFP were mixed with anti-Luciferase siRNA treated U2OS-RFP cells at a 1:1 ratio. 72 hr post transfection cells were treated with DNA damaging agents as depicted. Cells were harvested for FACS analysis 7 days post treatment. siRNA duplex sequences were as follows: Luc: CGTACGCGGAATACTTCGA (siRNA targeting firefly luciferase mRNA), CEP164#1: GGACCATC CATGTGACGAA; CEP164#2: GGCTGGAACGTGTCAAGAA; CEP164#3: GAGTTGGAGTCTCAACAGA; ZNF423#1: GGAGAACCA CAAGAACATT; ZNF423#2: CGAGTGCAGTGTCAAGTTT; ZNF423#3: GCATCAACCACGAGTGTAA, all from Ambion; BRCA2: (Stealth siRNA, Invitrogen) used as a combination of three siRNAs: GGAACCAAATGATACTGATCCATTA, GGAGGACTCCTTATGTC CAAATTTA, GAGCGCAAATATATCTGAAACTTC;ATM (Stealth siRNA, Invitrogen) used as a combination of three siRNAs: GCGCA GTGTAGCTACTTCTTCTATT, GGGCCTTTGTTCTTCGAGACGTTAT, GCAACATTTGCCTATATCAGCAATT; anti-CEP164 antibody for immunoblot was from Novus (cat #45330002). Quantitative RT-PCR Intron spanning primer pairs were designed for quantitative RT-PCR of ZNF423 and beta-actin as control. RNA was extracted from the siRNA treated U2OS cells using QIAGEN RNeasy Plus Mini Kit. First-strand cDNAs were then synthesized using the Invitrogen SuperScript III Reverse Transcriptase kit. Quantitative RT-PCR was performed using Platinum SybrGreen Super Mix (Invitrogen) according to the manufacturer’s instructions. Sequences of primer pairs were: Beta-actin forward: GCTACGAGCTGCCTGACG, Beta-actin reverse: GGCTGGAAGAGTGCCTCA, ZNF423 forward: GTCTCTGGCAGACCTGACG, ZNF423 reverse: AGAGTTGTGGG TCGTCATCA. ZNF423 DNA Damage Response Lentivirus constructs were obtained in a pLKO vector (Sigma) modified to express Emerald GFP in place of the Puromycin marker. Transduction efficiencies were $70% by GFP fluorescence. For DDR assays, transduced cells were plated at 1x105 per well in 12-well plate containing poly-L-lysine coated glass cover slides one day before irradiation with a calibrated source (Moores UCSD Cancer Center). Two hr after irradiation, cells were fixed with 4% PFA. Mouse anti-gH2AX (Upstate, clone JBW30) and rabbit anti-ZNF423 (Y-W Cho and B.A.H. unpublished data) antibodies were added at room temperature for 2 hr. Donkey anti-mouse (Alexa fluor 555) and anti-rabbit (Alexa fluor 488) antibodies (Jackson Immunoresearch) were incubated at room temperature for 1 hr. Images were captured on an Olympus FV1000 confocal microscope. Signal intensities of gH2AX were quantified from image files in ImageJ software. Distribution of foci in Zfp423-knockdown cells did not meet the conditions of a t test and were analyzed by the Mann-Whitney U (Wilcoxon rank sum) test implemented in R 2.8.1. FACS Analysis of gH2AX Both control and Cep164 stable knockdown IMCD3 cells were irradiated with indicated doses (Figure 6Q). Cells were then fixed 60 min post-irradiation with 2% paraformaldehyde, permeabilized with 0.1% SDS and stained with rabbit gH2AX (Cell Signaling). Alexa-fluor-488 conjugated secondary anti-rabbit antibody was used to analyze the mean fluorescent intensity (MFI) of samples in triplicate in the FACS facility (Cancer Center, University of Michigan). Results shown are representation of mean and SD of triplicate samples. Statistical Analysis Student’s two-tailed nonpaired t tests and normal distribution two-tailed z tests were carried out using pooled standard error and s.d. values to determine the statistical significance of different cohorts. SUPPLEMENTAL REFERENCES Bentley, D.R., Balasubramanian, S., Swerdlow, H.P., Smith, G.P., Milton, J., Brown, C.G., Hall, K.P., Evers, D.J., Barnes, C.L., Bignell, H.R., et al. (2008). Accurate whole human genome sequencing using reversible terminator chemistry. Nature 456, 53–59. Bryja, V., Gradl, D., Schambony, A., Arenas, E., and Schulte, G. (2007a). Beta-arrestin is a necessary component of Wnt/beta-catenin signaling in vitro and in vivo. Proc. Natl. Acad. Sci. USA 104, 6690–6695. Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007b). Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1dependent mechanism. J. Cell Sci. 120, 586–595. Bukanov, N.O., Smith, L.A., Klinger, K.W., Ledbetter, S.R., and Ibraghimov-Beskrovnaya, O. (2006). Long-lasting arrest of murine polycystic kidney disease with CDK inhibitor roscovitine. Nature 444, 949–952. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S5 Coene, K.L., Mans, D.A., Boldt, K., Gloeckner, C.J., van Reeuwijk, J., Bolat, E., Roosing, S., Letteboer, S.J., Peters, T.A., Cremers, F.P., et al. (2011). The ciliopathy-associated protein homologs RPGRIP1 and RPGRIP1L are linked to cilium integrity through interaction with Nek4 serine/threonine kinase. Hum. Mol. Genet. 20, 3592–3605. Giorgio, G., Alfieri, M., Prattichizzo, C., Zullo, A., Cairo, S., and Franco, B. (2007). Functional characterization of the OFD1 protein reveals a nuclear localization and physical interaction with subunits of a chromatin remodeling complex. Mol. Biol. Cell 18, 4397–4404. Habbig, S., Bartram, M.P., Mu¨ ller, R.U., Schwarz, R., Andriopoulos, N., Chen, S., Sa¨ gmu¨ ller, J.G., Hoehne, M., Burst, V., Liebau, M.C., et al. (2011). NPHP4, a cilia-associated protein, negatively regulates the Hippo pathway. J. Cell Biol. 193, 633–642. Letteboer, S.J., and Roepman, R. (2008). Versatile screening for binary protein-protein interactions by yeast two-hybrid mating. Methods Mol. Biol. 484, 145–159. Liebau, M.C., Ho¨ pker, K., Mu¨ ller, R.U., Schmedding, I., Zank, S., Schairer, B., Fabretti, F., Ho¨ hne, M., Bartram, M.P., Dafinger, C., et al. (2011). Nephrocystin-4 regulates Pyk2-induced tyrosine phosphorylation of nephrocystin-1 to control targeting to monocilia. J. Biol. Chem. 286, 14237–14245. Otto, E.A., Schermer, B., Obara, T., O’Toole, J.F., Hiller, K.S., Mueller, A.M., Ruf, R.G., Hoefele, J., Beekmann, F., Landau, D., et al. (2003). Mutations in INVS encoding inversin cause nephronophthisis type 2, linking renal cystic disease to the function of primary cilia and left-right axis determination. Nat. Genet. 34, 413–420. Otto, E.A., Ramaswami, G., Janssen, S., Chaki, M., Allen, S.J., Zhou, W., Airik, R., Hurd, T.W., Ghosh, A.K., Wolf, M.T., et al.; GPN Study Group (2011). Mutation analysis of 18 nephronophthisis associated ciliopathy disease genes using a DNA pooling and next generation sequencing strategy. J. Med. Genet. 48, 105–116. Sayer, J.A., Otto, E.A., O’Toole, J.F., Nurnberg, G., Kennedy, M.A., Becker, C., Hennies, H.C., Helou, J., Attanasio, M., Fausett, B.V., et al. (2006). The centrosomal protein nephrocystin-6 is mutated in Joubert syndrome and activates transcription factor ATF4. Nat. Genet. 38, 674–681. Sivasubramaniam, S., Sun, X., Pan, Y.R., Wang, S., and Lee, E.Y. (2008). Cep164 is a mediator protein required for the maintenance of genomic stability through modulation of MDC1, RPA, and CHK1. Genes Dev. 22, 587–600. Smith, L.A., Bukanov, N.O., Husson, H., Russo, R.J., Barry, T.C., Taylor, A.L., Beier, D.R., and Ibraghimov-Beskrovnaya, O. (2006). Development of polycystic kidney disease in juvenile cystic kidney mice: insights into pathogenesis, ciliary abnormalities, and common features with human disease. J. Am. Soc. Nephrol. 17, 2821–2831. Smogorzewska, A., Matsuoka, S., Vinciguerra, P., McDonald, E.R., 3rd, Hurov, K.E., Luo, J., Ballif, B.A., Gygi, S.P., Hofmann, K., D’Andrea, A.D., and Elledge, S.J. (2007). Identification of the FANCI protein, a monoubiquitinated FANCD2 paralog required for DNA repair. Cell 129, 289–301. Zhou, W., Dai, J., Attanasio, M., and Hildebrandt, F. (2010). Nephrocystin-3 is required for ciliary function in zebrafish embryos. Am. J. Physiol. Renal Physiol. 299, F55–F62. S6 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Figure S1. Links of Identified Nephronophthisis-Related Ciliopathy (NPHP-RC) Proteins to DDR Signaling and Cell-Cycle Control DDR pathways function to sense and repair DNA damage and, if repair is incomplete, arrest cell cycle at checkpoints to avoid propagation of DNA replication errors. Columns depict two pathways of DDR signaling, the ATM (ataxia-telangiectasia mutated) pathway (A), and the ATR (ATM-and-Rad-related) pathway (C). Rows depict stages of DDR signaling including DNA damage sensing and repair (A, C), as well as outcomes regarding checkpoint activation, cell-cycle arrest, and apoptosis (D). Some known components of DDR are shown in gray. Proteins relevant to this study are shown in color. Gene products identified in this study as mutated in individuals with NPHP-RC are encircled with a red dashed oval. Each of the 3 NPHP-RC proteins that we discovered most recently, MRE11, ZNF423 and CEP164 as well as others that we discovered recently, including ATXN10 (Sang et al., 2011) and SDCCAG8/NPHP10 (Otto et al., 2010) exhibit a functional connection to DDR pathways as follows: (A) Within the ATM-Chk2 pathway ionizing radiation (IR) can induce DNA double-strand breaks (DSB) activating PARP1. PARP1 may interact with ZNF423 (Figure 2E) (Ku et al., 2003), which we found here to be defective in individuals with NPHP-RC (Figures 1C and 1D). PARP1 recruits the MRN/ATM complex to DSBs. Activation of the ATM kinase activity by MRN (a heterotrimer of MRE11, Rad50 and Nbs) and TIP60 leads to i) induction of the gH2AX-dependent signaling cascade, ii) phosphorylation of CHK2 and p53 and, iii) to the recruitment of a large number of other DDR factors (not shown). We here detect an MRE11 mutation in an individual with NPHP-RC (Figures 1A and 1B). The ATM activator TIP60 is part of the ‘TIP60 protein complex’ (B). (B) Relationships of TIP60 with other proteins are shown within a blue inset, including gene products that are mutated in NPHP-RC of humans (red), mouse (orange rim), and zebrafish (yellow). Indirect protein interactions are depicted as black lines, direct interaction as black dots. Protein-protein interactions identified within this study are shown as purple lines or dots. Proteins that we detected as defective in NPHP-RC in this study are encircled by a red dotted line. TIP60 activates SC35/SRSF2 (Edmond et al., 2011) and directly interacts with the NPHP-RC protein OFD1 (Giorgio et al., 2007). TIP60 colocalizes to nuclear foci with the proteins encircled by a black dashed line (see Figure 4). The proteins marked with a lightning have known functions in DDR signaling. (C) Within the ATM-to-Chk1 pathway DNA lesions induced by UV light or replication stress (denoted by blue oval) cause replication fork stalling and accumulation of RPA-coated ssDNA regions, which recruit the ATR/ATRIP and the RAD17/RFC2-5 complexes. RAD17/RFC2-5 loads the 9-1-1 complex and stimulates ATR kinase activity by the 9-1-1-associated protein TOPBP1. ATR then activates Chk1 phosphorylation (modified from Ciccia and Elledge, 2010). CEP164 interacts with ATRIP through its N-terminal interaction domain and is phosphorylated by ATR in response to DNA damage by UV light (Sivasubramaniam et al., 2008). (D) The DDR pathways are important to sense and repair DNA damage and, if repair is incomplete, arrest cell cycle at checkpoints. If DNA repair is successful, cell cycle progresses. If DNA repair fails, apoptosis may result or, through activation of checkpoint proteins (Chk1 and Chk2) and inhibition of the CDC25s, cell-cycle arrest in G1/S or G2/M may result. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S7 Figure S2. Alternative Transcripts, Knockdown Targets, and Human Mutations of CEP164 with Interacting Domains, Interaction Partners, and Antigens of the Cep164 Protein (A) The CEP164 gene extends over 85.4 kb, contains 33 exons (vertical hatches) and an alternative exon 5a used in isoform 2. (B) Exon structure of human CEP164 cDNA. Positions of start codon (ATG) and of stop codon (TGA) are indicated. For the mutations detected (see (F) arrows indicate positions in relation to exons and protein domains (see (C). Positions of exon-targeting shRNAs are shown for knockdown in human (orange) and mouse (blue). (C) Domain structure of the CEP164 protein. One N-terminal globular domain (WW), a lysine-rich repeat (LR), a glutamate-rich region (Glu), and five coiled-coil domains (CC1-5) are predicted. (D) Minimal segment to which CEP164 binds the interaction partner ATRIP has been mapped. (E) Extent of antigens used for a-CEP164 antibody production. (F) CEP164 mutations detected in 4 families with NPHP-RC. Family number and predicted translational change are indicated (see Table 1). Sequence traces are shown for mutations above normal controls. Mutated nucleotides are indicated by arrow heads in traces of normal controls. (H), homozygous mutation; (h), heterozygous mutation (see also Figure 1). S8 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. G 50 64 hTERT-RPE No doxycyclin Doxycyclin 10 ng/ml D hTERT-RPE -Tubulin hTERT-RPE GFPcontrol IMCD3 GFPcontrol F * E -Tubulin C A B Figure S3. Subcellular Localization of CEP164 (A) Characterization of anti-CEP164 antibodies by immunoblotting. IMCD3 cells that doxycycline-inducibly express GFP-CEP164 isoform 1 were induced with (+) or without (-) 10 ng/ml doxycyclin for 20 hr. 20 mg of RIPA extracted cell lysates were subjected to SDS-PAGE and blotted with the indicated antibodies. Molecular weight is shown in kDa. Note that anti-CEP164 antibodies recognize a band compatible with the size of EGFP-CEP164 isoform 1 (191 kDa; arrow heads). Loading control (on right) uses an anti-aÀtubulin antibody. (B) Characterization of a-CEP164 antibodies by overexpression of N terminally GFP-labeled human isoform 1 wild-type construct EGFP-CEP164-WT and immunofluorescence in cell lines. Cells were costained with antibodies a-CEP164-NR, -ENR, and -SR (red), respectively, demonstrating colocalization (Merge) with N-EGFP-hCEP164-WT (green). The secondary antibody alone (lower panels) does not result in a signal. Scale bars, 5 mm. (C) CEP164 in GFP-Centrin-2 mouse photoreceptors. In photoreceptors of centrin-2-GFP mice, immunofluorescence using anti-CEP164-ENR antibody detects Cep164 at basal bodies/mother centrioles (visible as a single red dot), whereas centrin-2-GFP is expressed at both centrioles, mother and daughter. Inset shows enlargement of a representative centrosome at 3-fold higher magnification. Nuclei are stained with DAPI. Scale bars, 5 mm. ONL, outer nuclear layer; PRL, photoreceptor layer; RPE, retinal pigment epithelium. (D–G) Expression of mutant CEP164 in hTERT-RPE cells abrogates localization to centrosomes. (D) Doxycyclin (Dox)-inducible overexpression of N terminally GFP-tagged human full-length CEP164 wild-type isoform 1 (NGFP-hCEP164-WT) in hTERT-RPE (human retinal pigment epithelium) cells demonstrates, in addition to a cytoplasmic expression pattern, localization at centrosomes. (E) In contrast, the signal at centrosomes is abrogated upon overexpression of an N terminally GFP-tagged truncated CEP164 construct (NGFP-CEP164-Q525X) representing the mutation p.Q525X occurring in NPHP-RC family F59. Similar data was obtained upon CEP164 expression in IMCD3 cells (see also Figures 3B–3D). (F) Expression of GFP alone as a negative control yielded no centrosomal expression in IMCD3 or hTERT-RPE cells. hTERT-RPE cells were stably transfected with the CEP164 constructs in a retroviral vector for doxycyclin-inducible expression (pRetroX-Tight-Pur). Following selection with puromycin for 2weeks, cells were induced with10 ng/ml of doxycyclin for 20 hr. Cells were then costained with g-tubulin (red) for the colocalization with NGFP-hCEP164-WT fusion proteins (green) (see also Figures 3B–3D). Scale bars, 10 mm. (G) Immunoblot of IMCD3 and hTERT-RPE cells expressing inducible GFP-CEP164 constructs confirmed doxycyclin-inducible NGFP-CEP164 fusion protein expression ($191 kDa) both with a-CEP164-NR and a-GFP antibody blotting (arrow heads). The band at $85 kDa (asterisk) may represent the endogenous isoform 2 of CEP164 (84 kDa). IMCD3 and hTERT-RPE cells were stably transfected with the N terminally tagged-GFP-CEP164 construct in a retroviral vector (pRetroX-Tight-Pur) for doxycyclin-inducible expression. Cells were selected on puromycin for 2 weeks. Both cell lines were plated and treated with the indicated doses of doxycycline and lyzed with RIPA buffer 24 hr after induction. A total of 50 mg of cell lysates were loaded onto SDS-PAGE and blotted with the antibodies indicated. The same membranes were reprobed with anti-a-tubulin antibody as an indicator of equal loading (GFP-blotted membrane shown) (see also Figure 3). See also Figure 3. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S9 Figure S4. Subcellular Localization of CEP164 and SDCCAG8 by Immunofluorescence (A) CEP164 colocalizes with the mother centriole (labeled with g-tubulin), the mitotic spindle poles (actetylated tubulin), and the midbody throughout the cell cycle in hTERT-RPE cells. Whereas in the centriole-engaged state (upper panel) g-tubulin antibody labels both centrioles, the a-CEP164 antibody only labels one centriole (the mother centriole) in hTERT cells. Immunofluorescence of endogenous CEP164 was performed in hTERT cells. Cells were fixed (4% PFA), permeabilized (0.1% SDS) and immuno-stained with antibody anti-CEP164-NR (red) and costained with anti-g-tubulin or anti-acetylated tubulin antibodies (green). DAPI was used to label DNA (Blue). Scale bars, 2.5 mm. (B)–(D) Upon immunofluorescence (IF) in hTERT-RPE cells antibody a-SDCCAG8-CG recognizes nuclear foci that are absent upon transient SDCCAG8/NPHP10 knockdown. Left panels show transfected hTERT-RPE cells, middle panels show IF with a-SDCCAG8-CG, right panels show merge of left and middle panels. Antibody a-SDCCAG8-CG recognizes nuclear foci in hTERT-RPE cells (middle and right panels). (B) hTERT cells transiently transfected with GFP-labeled negative control shRNA construct exhibit nuclear foci upon IF with a-SDCCAG8-CG (arrow heads in right panel), whereas in cells transfected with GFP-labeled SDCCAG8 shRNA knockdown constructs pGIPZ-259547 (C) or pGIPZ-90858 (D) nuclear foci are absent (asterisks in right panels of (C) and (D), demonstrating specificity of the nuclear foci signal detected by a-SDCCAG8-CG. Scale bars, 5 mm. See also Figure 4. S10 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Figure S5. (A–B) Immuofluorescence Imaging of the NPHP-RC Protein SDCCAG8/NPHP10 with Other Proteins of Nuclear Foci in hTERT-RPE Cells, and (C–J) Identification of CCDC92 and TTKB2 as Direct Interaction Partners of CEP164 (A–B) SDCCAG8 labeled with antibody a-SDCCAG8-CG exihibits a similar number and size of nuclear foci in comparison to signals from antibodies against the nuclear foci markers promyelocytic leukemia protein (PML) (A) and centromere protein C (CENP-C) (B). However, these proteins do not colocalize with SDCCAG8. Scale bars, 5 mm. (C–J) Identification of CCDC92 and TTKB2 as direct interaction partners of CEP164. (C) Four baits, BD-CEP164fl , BD-CEP1641-550 , BD-CEP164551-1100 , and BD- CEP1641101-1640 were used to screen two retinal cDNA libraries, a human oligo-dT primed library and a bovine randomly primed library, employing a GAL-4 based yeast two-hybrid system. CEP164 was found to interact with two proteins, coiled coil domain containing protein 92 (CCDC92) and tau tubulin kinase 2 (TTBK2). The CEP164, CCDC92 and TTBK2 fragments are indicated with amino acids numbering in superscript. (D) Lysates from COS-1 cells transfected with 3xFlag- CCDC92fl or 3xFlag-TTBK2fl were used in a pull-down assay of GST fusion proteins GST-CEP164fl (191kDa), GST-CEP1-550 (87kDa), GST-CEP164551-1100 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S11 (92kDa), and GST-CEP1641101-1640 (68kDa). 3xFlag-CCDC92fl preferentially binds GST-CEP164fl and GST-CEP1641-550 , while 3xFlag-TTBK2fl binds all four GST fusion proteins. The results are further confirmed by unbound GST (26kDa), which shows no interaction with 3xFlag-CCDC92fl nor 3xFlag-TTBK2fl . (E–F) For coimmunoprecipitation 3xHA-CEP164fl in combination with 3xFlag-CCDC92fl were overexpressed in COS-1 cells (10% protein input shown). In an anti-HA immunoprecipitation 3xHA-CEP164fl coprecipitated with 3xFlag-CCDC92fl (E), whereas the coprecipitation with the unrelated 3xHA-dNp63 is very limited. This was confirmed in the reciprocal coIP using an anti-Flag antibody against 3xFlag-CCDC92fl (F). (G–H) For coimmunoprecipitation 3xHA-CEP164fl in combination with 3xFlag-TTBK2fl were overexpressed in COS-1 cells (10% protein input shown). In an antiHA immunoprecipitation of 3xHA-CEP164fl , this protein coprecipitated with 3xFlag-TTBK2fl (G), whereas the coprecipitation with the unrelated 3xHA-dNp63 was negative. This was confirmed in the reciprocal coIP using an anti-Flag antibody to immunoprecipitate 3xFlag-TTBK2fl (H). (I–J) (I) The a-CCDC92 antibody signal fully colocalizes with a-CEP164-M26 at the mother centriole upon immunofluorescence imaging in hTERT-RPE cells. (J) TTBK2 weakly colocalizes with a-CEP164-M26 at one of the centrioles, but yields a strong signal at the mid body in dividing hTERT-RPE cells. See also Figure 4. S12 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Figure S6. Molecular Interaction of NPHP-RC Gene Products with DDR Proteins. (A and B) Interaction of NPHP3 with CEP164. (A) To determine whether CEP164 interacts with known NPHP proteins, HEK293T cells were cotransfected with expression constructs for N terminally V5-tagged human full-length CEP164 and FLAG-tagged human full length NPHP1-NPHP5, NPHP8, NPHP9, or the control protein CD2AP as indicated. Comparable amounts of CEP164 were present in all lysates (middle panel). NPHP proteins were precipitated, using anti-Flag M2 beads (bottom panel). NPHP3 immobilized V5-tagged CEP164; a weak interaction was also detectable for NPHP1, 2 and 4 (top panel). (B) The reciprocal interaction confirmed an interaction between CEP164 and NPHP3. Flag-tagged CEP164 or CD2AP were coexpressed with V5-tagged NPHP1-NPHP4. Precipitation of NPHP3, but not the control protein CD2AP immobilized V5-tagged NPHP3. A weak, but reproducible interaction was also detected for NPHP4 (top panel). (C and D) Interaction of NPHP2 with DDB1. (C) HEK293T cells were transiently transfected with the Flag-tagged NPH-proteins NPHP1, NPHP2, NPHP4 or a control protein (Flag-EPS). Following immunoprecipitation with Flag-antibody Western blot analysis revealed that endogenous DDB1 coprecipitated with NPHP2 and to a lesser extent with NPHP4 but not with NPHP1 or the control protein. (D) Control experiment for protein expression using anti-FLAG antibody. (E and H) Cep164 and Dvl3 are in a precipitable complex. (E) CEP164 and Dvl colocalize at centrosomes. Whereas CEP164 is known to localize to the mother centriole only, Dvl3 is noted to label both centrioles (upper panel) or to the daughter centriole only (lower panel). HEK293 cells were fixed, labeled with antibody a-CEP164-PR and a-Dvl3. (F) Domain mapping of the CEP164 interaction with Dvl3. Series of bacterially produced CEP164-GST fragments were incubated with HEK293 cell lysates, precipitated using GST pull-down, and probed for presence of Dvl3 by immunoblotting against Dvl3 and GST, respectively. Endogenous Dvl3 interacts with a region within the first 194 amino acids of CEP164, which corresponds to the ATRIP binding domain. (G) CEP164 interacts with the proline-rich region of Dvl3. HEK293 cells were transfected with GFP-tagged full-length human CEP164 and a panel of truncation mutants of Flag-Dvl3 as indicated, and coimmunoprecipitation was performed with an anti-GFP antibody. Note that CEP164 only interacts with Dvl3 fragments that contain the proline-rich region (Pro). (H) Full length CEP164 but not the mutant CEP164-Q525X interacts with Dvl3. HEK293 cells were transfected with the indicated plasmids and subjected to immunoprecipitation using antibodies against FLAG-Dvl3 (upper panel) or CEP164-mCherryRFP (middle panel). Immunoblot analysis demonstrates strong interaction with Dvl3 for wild-type (wt) CEP164-mCherryRFP, whereas interactions is strongly reduced for CEP164-Q525X to Dvl3. Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. S13 Figure S7. Depletion of CEP164 or ZNF423(Zpf423) Results in Sensitivity to DNA Damaging Agents (A) Multicolor competition assay (MCA) (Smogorzewska et al., 2007). GFP-U2OS cells were transfected with the indicated siRNAs against CEP164 and ZNF423. Seventy-two hours after after transfection they were mixed with RFP-U2OS cells transfected with Luciferase (Luc) and DNA damage was induced using the indicated agents. After 7 days, percent of gfp-positive cells was measured using fluorescence-activated cell sorting and the resistance to DNA damage was calculated in comparison to untreated cells. DNA damage resistance was set at 100% for GFP-U2OS cells transfected with an siRNA against Luciferase (Luc). Cells transfected with siRNAs against ATM were used as a positive control. Experiments were done in triplicate. Standard deviations are indicated. Depletion of CEP164 (B) led to sensitivity to camptothecin (CPT) and gamma irradiation (IR). In addition, depletion with siRNA #3 led to sensitivity to mitomycin C (MMC). Furthermore, transfection of two separate ZNF423 siRNAs (#2 and #3), that resulted in ZNF423 mRNA depletion as assessed by RT-qPCR (C), caused sensitivity to MMC, CPT, and IR (L). An siRNA that did not result in ZNF423 mRNA depletion (C), siRNA #1 did not result in increased sensitivity to DNA damage. (B) Western blot analysis of CEP164 expression after siRNA transfection in cells used in the MCA assay shown in (A) (C) RT-qPCR in U2OS cells transfected with the three siRNAs against ZNF423 used in the MCA assay shown in (A). Reactions were performed in triplicate. Standard deviations are indicated. See also Figure 6. S14 Cell 150, 533–548, August 3, 2012 ª2012 Elsevier Inc. Vítězslav Bryja, 2014    Attachments      #14      L. Čajánek1 , R. Sri Ganji1 , C. Henriques‐Oliveira, P. Koník, S. Theofilopoulos, V. Bryja*, E.  Arenas*  (2013):  Tiam1  regulates  the  Wnt/Dvl/Rac1  signaling  pathway  and  the  differentiation of midbrain dopaminergic neurons. Mol. Cell. Biol. 33(1):59‐70.  1  equal contribution, *corresponding authors      Impact factor (2012): 5.372  Times cited (without autocitations, WoS, Feb 21st 2014): 4  Significance:  In  the  non‐canonical  Wnt  pathway,  Wnts  are  known  to  activate  small  GTPase Rac1 via Dishevelled. Dvl, however, cannot activate Rac1 directly but has  to act via guanine exchange factor (GEF). Here, we identify Tiam1, a Rac1‐GEF, as  a  long  searched  protein  responsible  for  Dvl‐mediated  activation  of  Rac1  in  the  non‐canonical Wnt pathway.  Contibution  of  the  author/author´s  team:  Identification  and  validation  of  Dvl‐Tiam1  interaction, and characterization of Tiam1 function in non‐canonical Wnt pathway                Tiam1 Regulates the Wnt/Dvl/Rac1 Signaling Pathway and the Differentiation of Midbrain Dopaminergic Neurons Lukáš Cˇajánek,a * Ranjani Sri Ganji,b Catarina Henriques-Oliveira,a Spyridon Theofilopoulos,a Peter Koník,c Víteˇzslav Bryja,b,d Ernest Arenasa Molecular Neurobiology Unit, Department of Medical Biochemistry and Biophysics, Karolinska Institute, Stockholm, Swedena ; Institute of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republicb ; Laboratory of Structural Biology, Faculty of Science, JCU, Ceske Budejovice, Czech Republicc ; Department of Cytokinetics, Institute of Biophysics, Academy of Sciences of the Czech Republic, Brno, Czech Republicd Understanding the mechanisms that drive the differentiation of dopaminergic (DA) neurons is crucial for successful development of novel therapies for Parkinson’s disease, in which DA neurons progressively degenerate. However, the mechanisms underlying the differentiation-promoting effects of Wnt5a on DA precursors are poorly understood. Here, we present the molecular and functional characterization of a signaling pathway downstream of Wnt5a, the Wnt/Dvl/Rac1 pathway. First, we characterize the interaction between Rac1 and Dvl and identify the N-terminal part of Dvl3 as necessary for Rac1 binding. Next, we show that Tiam1, a Rac1 guanosine exchange factor (GEF), is expressed in the ventral midbrain, interacts with Dvl, facilitates Dvl-Rac1 interaction, and is required for Dvl- or Wnt5a-induced activation of Rac1. Moreover, we show that Wnt5a promotes whereas casein kinase 1 (CK1), a negative regulator of the Wnt/Dvl/Rac1 pathway, abolishes the interactions between Dvl and Tiam1. Finally, using ventral midbrain neurosphere cultures, we demonstrate that the generation of DA neurons in culture is impaired after Tiam1 knockdown, indicating that Tiam1 is required for midbrain DA differentiation. In summary, our data identify Tiam1 as a novel regulator of DA neuron development and as a Dvl-associated and Rac1-specific GEF acting in the Wnt/ Dvl/Rac1 pathway. Wnt ligands are known to activate a number of highly evolutionary conserved pathways that regulate various essential aspects of embryo development and adult tissue homeostasis. The Wnt/␤-catenin pathway, also referred to as the canonical Wnt signaling pathway, is the most intensively studied and best defined pathway (1, 2). In this study, we focus on Wnt signaling pathways independent of ␤-catenin, in particular on Wnt-mediated activation of small GTPases (3, 4). These pathways are also involved in many aspects of embryogenesis, such as proper morphogenesis and differentiation of neuronal tissue (5), including the ventral midbrain (6, 7), the area where dopaminergic (DA) neurons are born. We and others have previously shown that Wnt5a activates a ␤-catenin-independent signaling in DA cells (8) and promotes DA differentiation of primary ventral midbrain precursor cells, as well as of mouse neural and embryonic stem cells (9–12). To date, the precise molecular mechanisms underlying the prodifferentiation effects of Wnt5a and the signaling pathways activated by Wnt5a in DA cells, as well as in other cell types, are only beginning to emerge. We recently demonstrated that the prodifferentiation effects of Wnt5a require the activity of the Rac1 GTPase (6). Rac1 belongs to the Rho family of small GTPases, and its activity is controlled by guanosine exchange factors (GEFs), GTPase activity-activating proteins (GAPs), and guanine nucleotide exchange inhibitors (GDIs) (13, 14). While Wnt ligands are known to trigger activation of various small GTPases from the Rho and Ras families in various cell types (4), evidence for the employment of specific GEFs in Wnt5a/Rac1 signaling is missing, and the mechanism of small GTPase activation in context of Wnt signaling is poorly understood. In the present study, we aimed to elucidate the mechanism of Rac1 activation in the Wnt5a/Rac1 pathway. Earlier work proposed that Rac1 interacts with Dishevelled (Dvl) (15, 16), a cytosolic protein with three homologs in mammals. Dvl is a critical component of Wnt-driven signaling, acts as a scaffolding protein associated with many different intracellular proteins (17, 18), and is phosphorylated by several kinases, such as casein kinase 1 (CK1) (9, 19). In our study, we first characterize the Dvl-Rac1 interaction and show that the N-terminal part of Dvl3 is required for complex formation between Rac1 and Dvl3, as well as for the activation of Rac1. We next searched for Dvl interactors and found that the Rac1 GEF T-cell lymphoma invasion and metastasis 1 (Tiam1) (20) is a novel binding partner of Dvl. Moreover, using loss-offunction (LOF) approaches, we demonstrated that Tiam1 is functionally required for the activation of Rac1 by Wnt5a or Dvl, as well as for generation of DA neurons from DA progenitors in neurosphere cultures. In sum, we hereby provide biochemical and functional evidence that Tiam1 is a novel regulator of Wnt/Dvl/ Rac1 signaling that is required for midbrain DA neuron differen- tiation. MATERIALS AND METHODS Cell culture, transfection, and treatments. The mouse SN4741 DA cell line was propagated as described before (21). For the purpose of transient Received 5 June 2012 Returned for modification 10 July 2012 Accepted 11 October 2012 Published ahead of print 29 October 2012 Address correspondence to Vitezslav Bryja, bryja@sci.muni.cz, or Ernest Arenas, ernest.arenas@ki.se. * Present address: Lukáš Cˇajánek, Biozentrum, University of Basel, Switzerland. L.C. and R.S.G. contributed equally to this work. Copyright © 2013, American Society for Microbiology. All Rights Reserved. doi:10.1128/MCB.00745-12 January 2013 Volume 33 Number 1 Molecular and Cellular Biology p. 59–70 mcb.asm.org 59 onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom gene overexpression, cells were transfected with Superfect (Qiagen) according to the manufacturer’s instructions. The following constructs were used: MYC-Dvl2 (22), FLAG-Dvl3, hemagglutinin (HA)-Dvl3 (23), Dvl2enhanced green fluorescent protein (EGFP) (24), HA-Dvl2 (wild type [wt] and M1, M2, and M4 mutants) (25), FLAG-Tiam1 (26), and MYCRac1 (27). For experiments involving small interfering RNA (siRNA), the SN4741 cells were transfected with 50 to 100 nM siGENOME Tiam1 siRNA (SMARTpool M-047-08-01) or nontargeting siRNA (both from Dharmacon) by using Lipofectamine 2000 (Invitrogen) according to the manufacturer’s recommendations. The efficiency of the silencing was assessed by Western blotting or quantitative PCR (qPCR). For analysis of cellular signaling, the cells were stimulated with 100 ng/ml of mouse Wnt5a (R&D Systems) for 2.5 h. Control stimulations were done with vehicle (0.1% bovine serum albumin [BSA], 0.05% CHAPS in phosphatebuffered saline [PBS]). HEK293A cells, T98G cells, U78MG cells, or mouse embryonic fibroblasts (MEFs) were grown in Dulbecco’s modified Eagle’s medium (DMEM) containing 10% fetal calf serum (FCS), 2 mM L-glutamine, 50 units/ml penicillin, and 50 units/ml streptomycin. Transient transfections of HEK293A cells were carried out with polyethylimin (PEI) with a ratio of 1 ␮g DNA per 3 ␮l of PEI in 100 ul serum-free medium, incubating the mixture for 10 min at room temperature, and adding the mixture dropwise to HEK293A cells in their culture medium. The following constructs, in addition to those listed above, were used: FLAG-Dvl3 deletion mutants (23). Inhibitors of CK1 (D4476, Calbiochem; PF-670462, Tocris Bioscience) were used as described previously (28, 29), where indicated. Immunoprecipitation, Western blotting, and densitometry analyses. HEK293A cells (24 to 36 h after transfection), SN4741 cells, T98G cells, or U78MG cells (endogenous immunoprecipitation [IP]) were harvested for total cell extract in RIPA buffer (150 mM NaCl, 50 mM Tris-Cl [pH 7.4], 1 mM EDTA, 0.5% NP-40, 1ϫ protease inhibitor cocktail [Roche]) or RIPA plus 0.05% SDS buffer (for Rac1 immunoprecipitations). Whole adult brain extract has been prepared by mechanical dissociation of brain tissue and subsequent lysis in complete RIPA buffer. Cell extracts were cleared from cell debris by centrifugation (16,000 ϫ g for 15 min at ϩ4°C), and supernatants were incubated with protein G-Sepharose (GE Healthcare) (1 h at ϩ4°C in an orbital shaker) to pull down nonspecific interactors. Subsequently, precleared extracts obtained after centrifugation (200 ϫ g for 5 min at ϩ4°C) were incubated with 1 ␮g (10 ␮g in case of IPs for mass spectrometry [MS] analyses) of the following antibodies: mouse monoclonal anti-Rac1 (clone 23A8; Millipore), rabbit polyclonal anti-Tiam1 (sc-872), mouse monoclonal anti-MYC (sc-40), mouse monoclonal anti-Dvl2 (sc-8026), mouse monoclonal anti-Dvl3 (sc-8027) (Santa Cruz Biotechnology), rabbit polyclonal anti-FLAG (F7425; Sigma) antibodies, for 30 min at ϩ4°C in an orbital shaker and subsequently with protein G-Sepharose. Following overnight incubation, samples were rigorously washed with RIPA buffer and subsequently analyzed by Western blotting. Sample preparation and Western blotting was done as described before (30). Luminescence was detected by either film exposure (Agfa), in the case of Fig. 3G and H and 4B to E, or the ChemiDoc XRS system (Bio-Rad). Nonsaturated images, within dynamic range of the charge-coupled-device (CCD) camera, were chosen as representative and used for densitometric analyses. The following antibodies were used, in addition to those described before: mouse monoclonal antiFLAG antibody (M2; Sigma), rabbit polyclonal anti-Dvl2 antibody (sc-13974; Santa Cruz Biotechnology), horseradish peroxidase (HRP)conjugated anti-mouse secondary antibody (GE Healthcare), and HRP-conjugated anti-rabbit secondary antibody (Sigma). GTPase pulldown assay. Cells were harvested in GTPase lysis buffer (150 mM NaCl, 10 mM Tris-Cl [pH 7.4], 10 mM MgCl2, 1% Triton X-100, 0.1% SDS, 1 mM dithiothreitol [DTT], 1ϫ protease inhibitor cocktail [Roche]). The pulldowns were performed with recombinant GST-PAK-CRIB as described before (31). For experiments with Tiam1 siRNA, the SN4741 cells were seeded out (750,000 cells/6-well plate), transfected first with either control or Tiam1 siRNA 24 h after seeding, and then transfected with the indicated expression constructs after an additional 24 h, processed 24 h after the last transfection, and subsequently analyzed by Western blotting. In experiments involving Dvl overexpression, MYC-Rac1 was cotransfected, and changes in Rac1 activity were subsequently measured for MYC-Rac1. A densitometry analysis was performed using Image J software, where indicated. The level of activated Rac1 (GTP-Rac1) was normalized to the level of Rac1 in total cell lysate (TCL). LC-MS/MS and data mining. Following the coimmunoprecipitation (extracts from 5- by 15-cm plates of 80 to 90% confluent MEFs were used per one condition), samples were washed in detergent-free RIPA buffer and subjected to proteomics grade trypsin digest (10 ng/␮l at 37°C for 12 h) (Sigma). Liquid chromatography-tandem mass spectrometry (LC-MS/ MS) was performed on a NanoAcquity ultraperformance liquid chromatograph (UPLC; Waters) coupled to an ESI Q-Tof Premier (Waters) mass spectrometer. Digested peptides were first desalted using a ZipTip C18 column (Millipore), diluted in MQ water, and loaded onto a nanoAcquity UPLC symmetry trap column (Waters) packed with 5 ␮m BEH C18 beads. Peptides were eluted through a nanoAcquity (Waters) analytical column packed with 1.7 ␮m BEH C18 beads at a flow rate of 400 nl/min using a gradient of 3 to 40% acetonitrile with 0.1% formic acid for 35 min. Effluent was directly fed into the ESI source of the MS instrument. Raw data were acquired in MSE Identity mode and later subjected to a search using UniProt and NCBI mouse protein database by the PLGS2.3 software (Waters). Acetyl N terminus, deamidation N and Q, carbamidomethyl C, and oxidation M were set as variable modifications. Peptide accuracy and MS/MS fragment mass accuracy was set to less than 20 ppm. Generated hit lists were manually curated to remove common contaminants and hits found in IgG IP samples. Immunocytochemistry and confocal microscopy. SN4741 cells (20,000 to 40,000 cells/well of a 24-well plate) were grown overnight on glass coverslips and transfected with the indicated plasmids. Twenty-four hours posttransfection, cells were fixed in 4% paraformaldehyde (15 min at ϩ4°C), serum blocked, and incubated in the appropriate primary and, subsequently, secondary antibodies as previously described (32). The following antibodies were used: mouse monoclonal anti-Rac1 (1:100; Millipore), rabbit polyclonal anti-FLAG (1:1,000; Sigma), rabbit polyclonal anti-MYC, mouse monoclonal anti-MYC (both 1:1,000; Santa Cruz Biotechnology), Alexa Fluor 488–goat anti-mouse or anti-rabbit, and Alexa Fluor 555–donkey anti-mouse or anti-rabbit (all 1:1,000; Molecular Probes, Invitrogen) antibodies. Fluorescent labeling was examined using a Zeiss LSM5 exciter inverted confocal scanning laser microscope. Precursor cultures. Ventral midbrain (VM) precursor cultures were grown as neurospheres, as previously described (12). Briefly, VMs were dissected from E10.5 mouse embryos, dissociated with collagenase/dispase (30 min at 37°C on a rocking platform), followed by mechanical trituration. Next, cells were plated at a density of 100,000 cells/cm2 in 1 ml of N2 medium (supplemented with 250 ng/nl Shh, 25 ng/ml fibroblast growth factor 8 [FGF8], and 20 ng/ml FGF2) and grown as neurospheres for 7 days. A total of 0.5 ml of fresh medium was added every second day. At day 7, spheres were collected and broken into small cell clusters (collagenase/dispase treatment plus mechanical trituration), which were then seeded at a density of 100,000 cells/cm2 on poly-D-lysine–laminin-precoated plates in N2 medium (no growth factors) and transfected with appropriate siRNA (50 to 100 nM) using Lipofectamine 2000. Fresh N2 medium was added 24 h after transfection (supplemented with brainderived neurotrophic factor (BDNF) and glial cell-derived neurotrophic factor (GDNF) to a final concentration 20 ng/ml of each), and cells were harvested 3 days after the transfection. For experiments with lentiviral particles, the VM tissue was dissected and plated as described above. The cells were transduced immediately after dissection in N2 medium containing Shh, FGF2, FGF8, and polybrene (2 ␮g/ml). After 24 h, the medium was replaced with a fresh one. Cells were grown as neurospheres for 5 days, with the addition of fresh medium every 2 to 3 days. At day 5, spheres were collected and seeded at a Cˇajánek et al. 60 mcb.asm.org Molecular and Cellular Biology onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom density of 50 spheres/cm2 on PDL-laminin-precoated plates in N2 medium (with BDNF and GDNF, 20 ng/ml each) and differentiated for 3 days. The efficiency of the silencing was determined by qPCR. For immunocytochemistry experiments, the following antibodies were used: rabbit polyclonal anti-tyrosine hydroxylase (anti-TH) (1:500; Pel-Freeze), mouse monoclonal anti-␤III tubulin (Tuj, 1:1,000; Promega), mouse monoclonal anti-TH (1:400; Sigma), mouse anti-GFP (1:75; Millipore), and rabbit anti-GFP (1:800; Invitrogen) antibodies. Two or three wells per condition were analyzed in each experiment. The DA neuron numbers were counted as either the number of THϩ cells per area of Tujϩ cell clusters (10 randomly selected observation fields per well, siRNA experiments) or the number of THϩ cells per area of DAPIϩ (4=,6-diamidino- 2-phenylindole) clusters (all spheres within the well analyzed, short hairpin RNA [shRNA] experiments). The area of DAPIϩ clusters was measured in square pixels and subsequently converted into square micrometers (1 square pixel ϭ 0.6209 square ␮m, for 10ϫ objective and 2ϫ zoom). CD1 mice (Charles River) were housed, fed, and sacrificed according to Karolinska Institute guidance for animal experiments and the ethical permits to E.A. Production of lentiviral shRNA particles. Third-generation replication-incompetent lentiviral vectors (pLKO.1-CMV-tGFP) encoding either nontargeting shRNA or Tiam1 shRNA (TRCN0000042594 and TRCN0000042595) were obtained from Sigma. The viral particles, vesicular stomatitis virus glycoprotein (VSVG) pseudotyped, were produced using the four-plasmid system as previously described (33). Forty-eight hours later, after transfection of HEK293T cells, the supernatant was collected and filtered. High-titer stocks were obtained by ultracentrifugation (50,000 ϫ g for 2 h), and the pellets were resuspended in PBS with 1% BSA and stored at Ϫ80°C until further use. The titers used for primary cell transduction were in the range of 1 ϫ 107 to 1.5 ϫ 107 transducing units per ml, and the multiplicity of infection was 1 to 2. Reverse transcription (RT)-PCR and quantitative PCR. Total RNA from either SN4741 or differentiated VM cultures was extracted using the RNeasy extraction kit (Qiagen) according to the manufacturer’s instructions. Preparation of cDNA, quantitative PCR (qPCR), and primers used in this study were described previously (10), with the exception of the following: Tiam1 primers (forward, GAG CCA GAG GGA GGC GTG GA; reverse, TGG CAT CCT GAG GGG ACG GG), Rac1 primers (forward, CGT CCC CTC TCC TAC CCG CA; reverse, CTT TCG CCA TGG CCA GCC CC). 18S was used as an internal control. Relative mRNA expression was calculated as a fold change versus the control. Statistical analyses. Statistical analyses (Student’s t test, Mann-Whitney test, one-way analysis of variance [ANOVA] with Neumann-Keuls multiple comparison test, and Kolmogorov-Smirnov test) were performed using Prism 4 (GraphPad Software). P values of Ͻ0.05 were considered statistically significant differences. Results are presented as means Ϯ standard errors of the means (SEM). RESULTS Characterization of the interaction between Rac1 and Dvl. We previously found that Wnt5a treatment activates the small GTPase Rac1 in DA cells (6). Indeed, stimulation of SN4741 dopaminergic (DA) cells with recombinant Wnt5a increased the level of Rac1GTP (activated Rac1) compared to vehicle-treated cells (Fig. 1A and B). It has been shown that Dishevelled (Dvl), a critical component of Wnt signaling machinery, is sufficient to increase Rac1 activity (15) and that it coimmunoprecipitates with Rac1 in hippocampal neurons (16). We thus decided to examine and characterize the possible interaction between Rac1 and Dvl in DA cells. As Dvl2 and Dvl3 are both involved in Wnt5a-mediated signaling in DA cells (9), we first analyzed the cellular distribution of Dvl2-MYC or Dvl3-FLAG after overexpression in SN4741 cells. Dvl, as a scaffolding protein, undergoes highly dynamic polymerization. This was reflected by formation of dynamic protein complexes in the cytosol (34), referred to as Dvl dots/puncta (35, 36). Interestingly, Dvl2-MYC/Dvl3-FLAG immmunoreactive puncta colocalized with endogenous Rac1 in cells where either Dvl2MYC or Dvl3-FLAG was overexpressed (Fig. 1C and D), suggesting that endogenous Rac1 is recruited to the Dvl-containing protein complexes. We next performed coimmunoprecipitation experiments in HEK293A cells and observed that endogenous Rac1 coimmunoprecipitates with both Dvl2-MYC (Fig. 1E) and Dvl3-FLAG (Fig. 2A) with similar efficiency. To get further insight into the interaction between Dvl and Rac1 and its possible functional consequences for Wnt signaling, the region of Dvl that mediates binding FIG 1 Dvl and Rac1 form a complex. (A) Wnt5a treatment (100 ng/ml; 2.5 h) triggers the activation of Rac1 (increase in Rac1-GTP) in SN4741 DA cells. (B) Quantification of the effect of Wnt5a on Rac1 activity, compared to vehicle treatment. The levels of Rac1-GTP were normalized to the total amount of Rac1 under each condition. (C) Colocalization of Rac1 (Alexa 555, red) and Dvl2-MYC (Alexa 488, green). Note the recruitment of endogenous Rac1 into immunoreactive puncta by expression of Dvl2-MYC. (D) Colocalization of Dvl3-FLAG (Alexa 555, red) and Rac1 (Alexa 488, green). (E) Dvl2-MYC coimmunoprecipitates with endogenous Rac1 in HEK293A cells. The right panel shows total cell lysate (TCL). *, P Ͻ 0.05 by Student t test. Tiam1 Regulates Noncanonical Wnt Signaling January 2013 Volume 33 Number 1 mcb.asm.org 61 onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom to Rac1 was determined using the FLAG-tagged deletion constructs of Dvl3 (23). The full-length Dvl3-FLAG and the three C-terminal deletion Dvl3 mutants coimmunoprecipitated with endogenous Rac1 (Fig. 2A and B). Importantly, however, deletion mutants lacking the N-terminal part, including the DIX domain, failed to be detected after Rac1 immunoprecipitation (see lanes 2 to 4 in Fig. 2A). Note that Dvl3 mutant 5, which lacks the C-terminal part, including the PDZ and DEP domains, showed somewhat weaker coimmunoprecipitation with Rac1 than the remaining Dvl mutants. These results indicated that the DIX domain of Dvl is required for its interaction with Rac1 and that this complex is possibly stabilized by interaction(s) with the C-terminal part of Dvl. Since the DIX domain of Dvl is required for the formation of Dvl-containing polymers (25), we tested whether the failure to interact with Rac1 could be due to impaired polymerization of Dvl. Several amino acids (Y27, F43, and V67 together with K68) within the DIX domain have been demonstrated as critical for Dvl polymerization and Wnt/␤-catenin signaling (25). While Dvl2 DIX domain mutant M4 (Y27D) retained some of the capacity of Dvl2 to form large cytosolic complexes in SN4741 cells, mutants M1 (F43S) and M2 (V67A and K68A) completely failed to form such complexes (Fig. 2C) (25). Interestingly, M1 and M2 mutants retained both (i) the ability to coimmunoprecipitate with Rac1 (Fig. 2D), even if the M1 mutant was somewhat less efficient in coimmunoprecipitation, and (ii) the potential to induce Rac1 acFIG 2 The N-terminal part of Dvl is required to interact with Rac1. (A) Deletion mapping of the Dvl3 domain mediating the interaction with Rac1 in HEK293A cells. The FLAG-Dvl3 expression constructs used in coimmunoprecipitation experiments are showed in panel B (total cell lysates [TCL] shown to the right). Note that FLAG-Dvl3 pulldown after Rac1 immunoprecipitation (IP) was lost in Dvl3 mutants lacking the N-terminal part, including the DIX domain. Arrows point to the Dvl3-FLAG variants coimmunoprecipitating with endogenous Rac1. (B) Schematic representation of the constructs used in panel A and the results obtained. ϩ or Ϫ symbols indicate the efficiency of coimmunoprecipitation for the different mutants. (C) HEK293A cells were transfected with the indicated expression constructs of Dvl2 harboring point mutations in the DIX domain. Subcellular localization of individual Dvl2 mutants has been determined by immunocytochemical staining using anti-HA antibody. (D) Coimmunoprecipitation of all Dvl2 constructs was detected after Rac1 immunoprecipitation but not if control IgG was used instead of an anti-Rac1 antibody. (E) HEK293A cells were transfected with the indicated Dvl constructs and Rac1-MYC. Rac1 activity was determined by the level of Rac1-GTP pulldown. Note that Rac1 was not activated by the Dvl3 ⌬N-FLAG construct. (F) Summary of experiments shown in panel E. *, P Ͻ 0.05; **, P Ͻ 0.01 by Student t test (compared to mock). Cˇajánek et al. 62 mcb.asm.org Molecular and Cellular Biology onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom tivation, as shown by the increased levels of Rac1-GTP (Fig. 2E and F). On the other hand, Dvl3-⌬N (construct 2 in Fig. 2B), which lacks the whole DIX domain, failed to activate Rac1 (Fig. 2E and F). These results indicate that the N terminus of Dvl, including the DIX domain (amino acids 1 to 81), is required for the Dvl-Rac1 interaction and for the Dvl-mediated Rac1 activation. In contrast, structural integrity of the DIX domain, which is absolutely crucial for DIX-mediated polymerization and Dvl function in Wnt/␤-catenin signaling, is dispensable for activation of Rac1. These findings prompted us to examine alternative mechanisms of Rac1 activation by Dvl. Tiam1, a Rac1 GEF, interacts with Dvl. We next used a proteomic approach to identify endogenous interactors of Dvl that could be involved in the activation of Rac1. The PDZ domaincontaining proteins have been previously found to act as critical scaffolds for the activation of GTPases from the Rho family (37, 38). In our experiments, we performed immunoprecipitation of Dvl3 in mouse embryonic fibroblasts (MEFs) and subsequent LS-MS/MS. In addition to previously described Dvl binding partners such as nucleoredoxin (NXN) (17) or Ror2 (17), we found T-cell lymphoma invasion and metastasis-inducing protein 1 (Tiam1) (Fig. 3A), which is a Rac1 GEF known to activate Rac1 (20). In order to verify the Dvl-Tiam1 interaction, we first overexpressed Tiam1-FLAG and Dvl2-EGFP in SN4741 DA cells and subsequently analyzed their subcellular localization by confocal microscopy. In cells transfected with Tiam1-FLAG only, Tiam1FLAG was localized both in cytosol and at the membrane (Fig. 3B). Upon Dvl2-EGFP coexpression, the membrane-localized pool of Tiam1-FLAG was not significantly mobilized, but the cytosolic pool was efficiently relocated and colocalized with Dvl2EGFP in cytosolic puncta. These findings thus supported the possibility of a Dvl-Tiam1 protein-protein interaction in the cytosolic puncta (Fig. 3C). To further confirm that Dvl and Tiam1 are part of the same protein complex, we performed coimmunoprecipitations in HEK293A cells. We found that Tiam1 interacted with both Dvl2-HA (Fig. 3D) and Dvl3-FLAG (Fig. 3E), as detected by their coimmunoprecipitation with Tiam1-FLAG. Importantly, analysis of adult mouse brain lysates showed that endogenous Tiam1 could be pulled down both with Dvl3 (Fig. 3F) and Dvl2 antibodies (Fig. 3G). These results thus show that both endogenous Dvl2 FIG 3 Tiam1 is a novel Dvl binding partner. (A) Tiam1 is present in Dvl3-associated complexes, identified by LC-MS/MS after endogenous Dvl3 coimmunoprecipitation. (B) Tiam1-FLAG protein (red) localizes both in the cytosol and at the membrane (arrow) upon overexpression in SN4741. Topro3 was used to counterstain the cell nucleus. (C) Coexpression of Dvl2-EGFP (green) and Tiam1-FLAG (red) in SN4741 cells leads to a cytoplasmic colocalization of Dvl2-EGFP and Tiam1-FLAG in puncta. Arrowheads show that a small pool of Tiam1-FLAG that did not colocalize with Dvl2-EGFP remained in the membrane. (D) Dvl2-HA was found to form a complex with Tiam1-FLAG, as assessed by immunoprecipitation with Tiam1-FLAG in HEK293A cells (first panel, last lane). The next panel shows total cell lysates (TCL). (E) Dvl3-FLAG was also immunoprecipitated by Tiam1-FLAG in HEK293A cells. The complex formed by Tiam1-FLAG and Dvl3-FLAG was detected by immunoprecipitation with either anti-Dvl3 or anti-Tiam1 antibodies. (F and G) Endogenous Dvl3 (F) or endogenous Dvl2 (G) were immunoprecipitated from whole adult mouse brain lysates, and the presence of Tiam1 has been detected with an anti-Tiam1 antibody. IgG, control antibody. Tiam1 Regulates Noncanonical Wnt Signaling January 2013 Volume 33 Number 1 mcb.asm.org 63 onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom and Dvl3 form complexes with endogenous Tiam1. Detailed mapping of the interaction domains using Dvl3 deletion constructs described in Fig. 2B showed that the proline-rich domain between PDZ and DEP is essential for Tiam1 binding (Fig. 4A and B). Our domain mapping of Dvl suggests that while Rac1 binds at the N-terminal DIX domain, Tiam1 binds at the C-terminal prolinerich domain of Dvl, and hence Dvl can serve as a scaffold for the Rac1-Tiam1 interaction. Tiam1 interaction with Dvl is promoted by Wnt5a and CK1 inhibition. We next examined whether the interaction between Dvl and Tiam1 could be modulated by casein kinase 1 (CK1), a critical regulator known to phosphorylate several components of the Wnt pathway, including Dvl (9, 19). Importantly, the prolinerich domain of Dvl, which we found to be required for Tiam1 binding, contains many residues phosphorylated by CK1 (39, 40). We previously identified CK1ε as a negative regulator of Dvlinduced Rac1 activation and found that pharmacological inhibition of CK1 by D4476 increases the level of Rac1-GTP in mouse embryonic fibroblasts (31). As shown in Fig. 5A, D4476 (50 ␮M, 2 h) led to increased Rac1 activity also in SN4741 cells. We therefore examined whether the interaction between Dvl and Tiam1 is modified by CK1ε overexpression or pharmacological inhibition of CK1. Overexpression of CK1ε completely abrogated the interFIG 4 Proline-rich region of Dvl3 is required for interaction with Tiam1. (A) Schematic representation of the domain mapping of the interaction between Tiam1-HA and Dvl3-FLAG deletion mutants. ϩ or Ϫ symbols indicate the efficiency of coimmunoprecipitation for the different mutants used in panel B. The top panel shows detection of Dvl3-FLAG mutants after immunoprecipitation of Tiam1-HA by anti-HA antibody. Note the lack of signal from FLAG antibody in lanes with Dvl3 mutants 496-716 and 1-246. Bottom panel shows total cell lysate (TCL). FIG 5 CK1 and Wnt5a regulate Tiam1-Dvl3 interaction. (A) Inhibition of CK1 with 50 ␮M D4476 (for 2 h) induced the activation of Rac1 in SN4741 cells. Activation was determined by pulldown of Rac1-GTP and subsequent Western blotting. The efficiency of the inhibition of CK1 was monitored by the disappearance of the hyperphosphorylated form of Dvl2 after D4476 treatment. (B) CK1 negatively regulates the interaction between Tiam1-FLAG and Dvl3-HA as assessed by coimmunoprecipitation of Dvl3-HA by FLAG antibodies and subsequent Western blotting. Note that the coexpression of CK1ε abrogated the Tiam1-Dvl3 interaction (compare lanes 4 and 7), while pharmacological inhibition of CK1 increased the coimmunoprecipitation of Dvl3-HA (lanes 5 and 6) and rescued the effect of CK1ε overexpression (lane 8). CK1 inhibitors D4476 (50 ␮M) and PF-670462 (5 ␮M) were added to the medium 2 to 3 h before cells were harvested for immunoprecipitation. (C to E) Wnt5a and CK1 inhibitor (PF-670462) promoted the interaction between endogenous Dvl3 and Tiam1 in dopaminergic SN4741 (C), glioma T98G (D), and U78G (E) cell lines. Immunoprecipitation with anti-Tiam1 antibodies was followed by Western blot detection of Dvl3 interacting with Tiam1. Cˇajánek et al. 64 mcb.asm.org Molecular and Cellular Biology onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom action between Tiam1-FLAG and Dvl3-HA (compare lanes 4 and 7 in Fig. 5B). Moreover, both endogenous and overexpressed CK1 was found to coimmunoprecipitate with Tiam1-FLAG (lanes 5 to 8 in Fig. 5B). On the other hand, the inhibition of CK1 by either D4476 (50 ␮M) or PF-670462 (5 ␮M) for 2 to 3 h led to a robust increase in the amount of Dvl3-HA that coimmunoprecipitated with Tiam1-FLAG. Importantly, the negative effect of CK1 overexpression was rescued by CK1 inhibition (compare the last two lanes in Fig. 5B). These results demonstrate that the interaction between Dvl and Tiam1 is under the dynamic control of CK1 and suggest that CK1 may also regulate the interaction between Dvl and Tiam1 in the context of endogenous Wnt signaling. In order to test this possibility, we have analyzed the effects of Wnt5a treatment and CK1 inhibition on the interaction of endogenous Dvl3 and Tiam1 in dopaminergic SN4741 cells. Our results show that while the interaction between Tiam1 and Dvl3 is almost negligible in control conditions, it can be clearly induced by Wnt5a treatment or by the CK1 inhibition by PF-670462 (Fig. 5C). These findings were also confirmed in other neural cell lines, such as U78MG (Fig. 5D) and T98G (Fig. 5E). Thus, our results indicate that Wnt5a and CK1 inhibition promote the interaction between endogenous Dvl and Tiam1 in different neural types, including midbrain DA neurons. Does Tiam1 regulate Rac1 function and the formation of Dvl-Rac1 complexes? The results obtained until this point indicated that Dvl2 and Dvl3 interact and colocalize not only with Rac1 but also with Tiam1 and that Tiam1-Dvl interaction is under tight control of CK1 and Wnt5a. We thus decided to examine whether Tiam1 can regulate the formation of the Dvl-Rac1 complex and the activity of Rac1. We found that the coimmunoprecipitation of Dvl2-MYC with Rac1 was enhanced when Tiam1FLAG was overexpressed (Fig. 6A). Quantification of these results (n ϭ 5) indicated that the expression of Tiam1 increased the Dvl2Rac1 coimmunoprecipitation by 46%, compared to the condition without Tiam1 expression (Fig. 6B). Thus, our results suggest that the presence of Tiam1 facilitates the Dvl-Rac1 interaction. We next examined whether Tiam1 also acts as a Rac1 GEF and regulates the activity of Rac1. As expected, Tiam1 overexpression was sufficient to induce Rac1 activation in both HEK293A and SN4741 cells (Fig. 6C). In light of these data, we found it plausible that Tiam1 could be functionally required for Rac1 activation in context of the Wnt/Dvl/Rac1 pathway. In order to test this possibility, we knocked down Tiam1 protein in SN4741 DA cells using a Tiam1 siRNA. While Dvl2-MYC or Dvl3-FLAG overexpression increased the level of activated MYC-tagged Rac1-GTP in control siRNA cells, they did not activate Rac1 when Tiam1 expression was knocked down (Fig. 6D and data not shown). Moreover, the relative increases in Rac1-GTP levels after Dvl2-MYC or Dvl3FLAG overexpression were significantly decreased to almost basal levels after Tiam1 knockdown (Fig. 6E and data not shown). Next, we tested whether Tiam1 was similarly required for Wnt5a liganddependent activation of Rac1 in SN4741 DA cells. We found that the ability of Wnt5a to increase Rac1-GTP levels was impaired when Tiam1 expression was knocked down by siRNA (Fig. 6F and G). In sum, our results show that Tiam1 is sufficient for Rac1 activation in DA cells and, more importantly, that Tiam1 is required for the activation of Rac1 by either Dvl overexpression or acute stimulation by Wnt5a. Thus, our results identify Tiam1 as a novel regulator of the noncanonical Wnt/Dvl/Rac1 signaling pathway. FIG 6 Tiam1 functionally interacts with Dvl to regulate Rac1 activation. (A) Dvl2-MYC was pulled down by Rac1 (first lane), and this coimmunoprecipitation was enhanced by coexpression of Tiam1-FLAG (third lane). (B) Densitometric quantification of the relative amount of Dvl2-MYC coimmunoprecipitating with Rac1 in control or Tiam1 cotransfected cells. Values were normalized to the total level of Dvl-MYC in total cell lysate (TCL). (C) SN4741 and HEK293A cells were transfected with a Tiam1-FLAG or mock plasmid and analyzed for the level of GTP-Rac1 24 h after transfection. Tiam1-FLAG was sufficient to increase GTP-Rac1 levels, compared to that of the control, in both cell lines. (D) SN4741 cells were transfected with control or Tiam1 siRNAs and subsequently with Rac1-MYC and Dvl3-FLAG or mock vector. Dvl3-FLAG increased the levels of MYC-tagged Rac1-GTP compared to those of mocktransfected SN4741 cells, but not in cells where Tiam1 expression was knocked down by siRNA. (E) Densitometric quantification of the relative level of Rac1GTP (normalized to total MYC-tagged Rac1) induced by Dvl3-FLAG in control or Tiam1 siRNA conditions. (F) SN4741 cells transfected with control or Tiam1 siRNA and 24 h after they were stimulated with Wnt5a (100 ng/ml) or vehicle (PBS-CHAPS) for 2.5 h and subsequently analyzed for Rac1 activation (level of Rac1-GTP). (G) Densitometric quantification of the relative level of Rac1-GTP (normalized to total Rac1) induced by Wnt5a stimulation in control siRNA versus Tiam1 siRNA (second versus fourth lane). *, P Ͻ 0.05 by Mann-Whitney test. Tiam1 Regulates Noncanonical Wnt Signaling January 2013 Volume 33 Number 1 mcb.asm.org 65 onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom Cˇajánek et al. 66 mcb.asm.org Molecular and Cellular Biology onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom Tiam1 is required for the differentiation of midbrain DA precursors into neurons. As we hereby report that Tiam1 regulates the function of Rac1 in the context of the Wnt/Dvl/Rac1 pathway and we have previously reported that Wnt5a and Rac1 are involved in differentiation of primary ventral midbrain DA neurons (6), we decided to examine the possible contribution of Tiam1 to this process. First, we analyzed the expression of Tiam1 mRNA in mouse ventral midbrain (VM) at different embryonic (E) stages, from E10.5 to E15.5, by qPCR. We found that the expression of Tiam1 mRNA progressively increased during embryonic development, reaching maximal levels at E15.5 (Fig. 7A). Moreover, when we examined the levels of Tiam1 protein in the developing VM, we found that Tiam1 is present as early as at E10.5 (Fig. 7B). These results indicated a possible function of Tiam1 in midbrain development, at the indicated developmental stages. We therefore tested whether Tiam1 mRNA interference impaired the generation of midbrain DA neurons in midbrain neural stem/progenitor cells expanded as neurospheres in the presence of Shh, basic FGF, and FGF8 for 7 days (12). Primary neurospheres were gently dissociated to smaller cell clusters, transfected with either control siRNA or Tiam1 siRNA, and differentiated for 3 days. Transfection of differentiated progenitor cultures with Tiam1 siRNA reduced the expression of Tiam1 mRNA by 50% (Fig. 7D) at day 3 of differentiation. Tiam1 gene knockdown led to significant decreases in the mRNA expression of beta III tubulin (Tuj), a neuronal marker, and tyrosine hydroxylase (TH), the rate-limiting enzyme in dopamine synthesis and marker of DA neurons (Fig. 7D). Similarly, Rac1 siRNA, which efficiency reduced Rac1 mRNA levels in SN4741 DA cells by 50 to 75% (data not shown), caused a decrease in the expression of TH mRNA (and Rac1 mRNA) in VM neurospheres differentiated for 3 days (Fig. 7C). Interestingly, Tiam1 gene knockdown also decreased the mRNA expression of Pitx3 (paired-like homeodomain transcription factor 3), a specific marker of VM DA neurons, but did not affect the expression of Nurr1, a transcription factor expressed both in DA neurons and postmitotic neuroblasts (Fig. 7D). These results suggested a role of Tiam1 in the last stages of differentiation, in the acquisition of two DA neuron markers, Pitx3 and TH, by Nurr1ϩ neuroblasts. Indeed, immunocytochemical analysis of differentiated ventral midbrain neurospheres showed that Tiam1 gene knockdown reduced by 50% the number of THϩ DA neurons in Tujϩ clusters normalized by area (Fig. 7E and F). These results indicated that Tiam1 is required for the generation of midbrain DA neurons from neural stem/progenitor cell preparations. Finally, since our biochemical data identified Tiam1 as a downstream component of Wnt5a-mediated signaling, we examined whether Tiam1 knockdown can also block the Wnt5a-induced DA differentiation in neurosphere cultures. In our experiments, Wnt5a treatment (100 ng/ml) increased the number of THϩ DA neurons per area by 58%, from 59.2 Ϯ 3.8 THϩ cells/mm2 to 93.6 Ϯ 8.3 THϩ cells/mm2 (Fig. 7G and H). These effects were efficiently ablated by lentiviral delivery of Tiam1 shRNA (61.9 Ϯ 6.5 THϩ cells/mm2 ). In sum, our data suggest that Rac1 and its upstream regulator Tiam1 are involved in the differentiation of midbrain DA neurons and that Tiam1 works downstream of Wnt5a in DA differentiation. DISCUSSION In this study, we investigated the mechanisms of activation of the small GTPase, Rac1, in the context of Wnt signaling in DA cells and found that the Rac1 GEF Tiam1 regulates the Wnt/Dvl/Rac1 signaling pathway and the differentiation of midbrain dopaminergic (DA) neurons. We previously reported that Wnt5a promotes the differentiation of DA neurons via a Wnt/␤-cateninindependent pathway (6, 8, 9) that involves Rac1 (31). The activation of Rac1 by Wnt/␤-catenin-independent signaling has been reported in various cellular systems (4), and Rac1 has been found to play a role in cytoskeleton rearrangements, cell polarity, and cell migration (13, 41), as well as in neuronal development and differentiation (42–44). Surprisingly, however, little is known about upstream mechanisms involved in the regulation of small GTPases in the context of the Wnt/␤-catenin independent signaling. Previous studies have shown that Dvl is both required and sufficient for the activation of Rac1 (15, 31, 45). Our results further suggest that the activation of Rac1 by Dvl requires the formation of protein complexes containing Dvl and Rac1, as Dvl mutants lacking the ability to interact with Rac1 also fail to induce Rac1 activation. We identified the N-terminal region of Dvl3 (1 to 81 amino acids) as critical for mediating the Dvl3-Rac1 interaction and the activation of Rac1 by Dvl3. Our results differ from the study by Habas and colleagues (15), who reported that the DEP domain of Dvl2 is sufficient to bind and activate Rac1 and failed to detect activation of Rac1 by Dvl3. In our study, we found that both Dvl2 and Dvl3 are sufficient to induce Rac1 activation and that the DEP domain of Dvl3 is not sufficient for Rac1 interaction. It is possible that methodological and cell-type-specific factors account for the discrepancy. However, data in the literature indirectly support an interaction of Rac1 with the N-terminal part of Dvl, as the N-terminal part of Dvl1 is sufficient for association with PAK, a direct effector and binding partner of Rac1 (46). Evidence for the employment of specific GEFs in Wnt/␤catenin-independent signaling has only recently begun to emerge. WGEF, p114-RhoGEF, and GEF-H1 have been proposed as GEFs for RhoA (47, 48). However, GEFs involved in the Wnt5a/Dvlmediated activation of Rac1 have not been described until now. Our data show first evidence for the involvement of a specific Rac1 FIG 7 Tiam1 is expressed in the developing VM and is required for Wnt5a-induced DA neuron differentiation. (A) qPCR analysis showed that Tiam1 mRNA is expressed between E10.5 and E15.5 in mouse ventral midbrain (VM). *, P Ͻ 0.05 by Mann-Whitney test (compared to E10.5). (B) Western blot analyses confirmed the presence of Tiam1 protein in both ventral midbrain and dorsal midbrain (DM) samples at E10.5. The floor plate marker Foxa2 was used to confirm the identity of the VM sample. (C) Rac1 siRNA of Rac1 and TH mRNA in differentiating ventral midbrain precursors, 3 days after transfection. (D) Tiam1 siRNA decreased the expression of Tiam1 mRNA, as well as Tuj, Th, and Pitx3 mRNAs but not Nurr1 mRNA in differentiating ventral midbrain precursors 3 days after transfection. These results suggested an effect of Tiam1 in DA differentiation, similar to that of Wnt5a. (E) Immunostaining of control or Tiam1 siRNA VM neurospheres showed similar numbers of Tuj1ϩ cells (green) but lower numbers of THϩ dopamine neurons (red) after 3 days of differentiation. (F) Quantification of the number of THϩ cells per area of Tujϩ neurospheres revealed that Tiam1 siRNA significantly reduced the number of dopamine neurons. (G) Treatment with Wnt5a (100 ng/ml, 3 days) increased the number of THϩ neurons in neurosphere cultures treated with control shRNA lentiviruses but not in Tiam1 shRNA lentiviruses. DAPI was used for nuclei counterstaining. (H) Quantification of the number of THϩ cells per area (mm2 ) of sphere, summary of 4 independent experiments. *, P Ͻ 0.05; **, P Ͻ 0.01; ***, P Ͻ 0.001 by Student t test or one-way analysis of variance (ANOVA) (Fig. 7H). Tiam1 Regulates Noncanonical Wnt Signaling January 2013 Volume 33 Number 1 mcb.asm.org 67 onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom GEF, Tiam1, in the noncanonical Wnt/Dvl/Rac1 pathway. This conclusion is based on the following lines of evidence. First, we found that Tiam1 interacts with Dvl and enhances the Dvl-Rac1 interaction. Second, we showed that the Dvl-Tiam1 interaction is regulated by CK1. Third, siRNA knockdown experiments demonstrated that Tiam1 is functionally required for the activation of Rac1 by Wnt5a or Dvl. These findings are in full agreement with the known negative role of CK1 in Wnt/Dvl/Rac1 signaling and provide a possible molecular mechanism by which CK1 regulates DA differentiation, namely by regulating both the interaction between Dvl and Tiam1 and the activation of Rac1. We previously reported that inhibition of CK1 or overexpression of dominant negative CK1ε increases the level of Rac1-GTP, in a Dvl-dependent manner, while overexpression of CK1 blocks the activation of Rac1 mediated by Dvl (31). Here, we show that inhibition of CK1 facilitates, whereas CK1 overexpression blocks, the interaction between Dvl and Tiam1. Moreover, we could detect CK1 present in Tiam1 immunoprecipitates. Interestingly, a recent proteomic study also found CK1 as a Tiam1-associated protein (49). Combined, our results suggest that CK1 controls changes in proteinprotein interactions between Dvl and Tiam1, which result in the activation of Rac1 by Tiam1. Surprisingly, however, we were unable to detect the interaction between Tiam1 and Rac1. One possible explanation for this is the instability of the trimeric complex formed by Tiam1-Rac1-GTP (50). Evidence in the literature supports the view that the activity of small GTPases is regulated by protein-protein interactions between GEFs and scaffolding proteins (37, 38, 51). This possibility is in agreement with our observation of a Dvl-Tiam1 interaction and a negative effect of CK1 on Dvl-Tiam1 binding and Wnt/Dvl/Rac1 signaling. We thus propose a model where CK1 activity determines the amount of Tiam1 associated with the scaffolding protein Dvl, which is hence available to bind and activate Dvl-bound Rac1. In this model, the Wntinduced activation of CK1 (9, 52) may thus represent a negative feedback mechanism to fine-tune the level of Rac1-GTP. Future experiments should aim at addressing whether the activation of Rac1 by Wnt5a/Dvl is exclusively controlled by Tiam1 or is redundantly controlled by other GEFs. Another pending issue is the identity of the signaling component(s) downstream of the Wnt5a/Dvl/Rac1 pathway. It has been previously shown that cJun acts downstream of Tiam1/Rac1 (53), opening the intriguing possibility that the Wnt5a/Dvl/Rac1 pathway directly regulates transcription via cJun. However, we found that Wnt5a did not activate a cJun-luciferase reporter in DA cells (data not shown). Thus, our results suggest that (i) cJun may not be downstream of Rac1 in the context of Wnt5a-mediated signaling and that (ii) an alternative factor(s) mediates the Wnt5a/Dvl/Rac1 pathway in DA neurons. At a functional level, Tiam1 has been previously shown to regulate neurite and axon outgrowth in neuroblastoma cells (54, 55) as well as in hippocampal neurons (56), migration of cortical neurons (57), and the formation of dendritic spines (26). Previous studies have also shown that the attenuation of Tiam1 and/or Rac1 activity impairs cytoskeleton remodeling and leads to loss of cell polarity, which in context of neuronal differentiation can be reflected by impaired neurogenic cell division, migration, and/or neuritogenesis (43, 58–60). Moreover, Wnt5a/Dvl-mediated signaling events have been implicated in the regulation of polarity in differentiating neurons (4, 6, 61). Thus, it is plausible that cell polarity underlies the requirement of Wnt5a/Dvl/Rac1 signaling for neuronal differentiation. Importantly, the role of Tiam1 in the ventral midbrain (VM) and its possible implication in the differentiation/development of DA neurons has not been addressed before. Here, we show that Tiam1 is expressed in the VM at the time when DA neurons are born. Moreover, we found that Tiam1 knockdown impairs the generation of DA neurons in VM progenitor cultures. Thus, it is plausible to expect that Tiam1 regulates a signaling pathway(s) relevant to DA development and differentiation (62). Interestingly, Wnt5a and Rac1 activation had previously been shown to promote the differentiation of Nurr1ϩ THϪ postmitotic neuroblasts into Nurr1ϩ THϩ Pitx3ϩ DA neurons (6, 10). In this regard, our biochemical and functional data suggest that Tiam1 mediates Wnt5a-induced DA neuron differentiation. Furthermore, our results indicating that Tiam1 knockdown decreases TH, Pitx3, and Tuj mRNA levels but does not affect the expression of Nurr1 are fully in agreement with the idea that Wnt5a promotes the transition of Nurr1ϩ DA neuroblasts into mature DA neurons. Thus, combined, our results suggest that Tiam1 plays a role in the differentiation of Nurr1ϩ DA precursors and is a key modulator of the Wnt/Dvl/Rac1 pathway. In summary, we hereby show that Rac1 interacts with its upstream regulator Dvl. We identified Tiam1, a Rac1-specific GEF, as a binding partner of Dvl and a critical regulator of Rac1 activation in the context of the Wnt5a/Dvl/Rac1 pathway. Finally, we provide functional evidence that Tiam1 is required for the Wnt5amediated generation of DA neurons in ventral midbrain progenitor cultures. Hence, our data highlight the important role of the Wnt5a/Dvl/Rac1 signaling pathway in DA neuron differentiation and pinpoint Tiam1, the first Rac1 GEF identified in noncanonical Wnt signaling. ACKNOWLEDGMENTS We thank R. J. Lefkowitz (HHMI, Duke University, Durham, NC) for Dvl2-EGFP, S. Yanagawa (Kyoto University) for Dvl2-MYC, R. T. Moon (HHMI, University of Washington, Seattle, WA) for Dvl3 constructs, Alan Hall (Memorial Sloan-Kettering Cancer Center, New York, NY) for Rac1-MYC, M. Bienz (LMB, MRC, Cambridge, United Kingdom) for Dvl DIX mutants (M1 to M4), Pontus Aspenström (KI, Stockholm, Germany) for GST-PAK CRIB, K. Tolias (Baylor College of Medicine, Houston, TX) for Tiam1-FLAG constructs, and Ondrˇej Slabý for U78MG and T98G cells. We acknowledge Johnny Söderlund, Nad’a Bílá, Jakub Harnoš, and Alessandra Nanni for excellent assistance. This work was supported by grants from the European Union (Neurostemcell), Swedish Foundation for Strategic Research (SRL Program and CEDB project), Swedish Research Council (VR2008:2811, VR2011: 3116, and DBRM), Norwegian Research Council, and Karolinska Institute to E.A., as well as grants from the Ministry of Education, Youth, and Sports of the Czech Republic (MSM0021622430, CZ.1.07/2.3.00/ 20.0180), EMBO Installation Grant, and Czech Science Foundation (204/ 09/0498; 204/09/H058) to V.B. L.C. was supported by a KID Ph.D. student fellowship (6110/06-225) from Karolinska Institute. S.T. was supported by a VR fellowship for visiting scientists and short-term fellowship from the Onassis Foundation. We declare no competing interests. REFERENCES 1. Angers S, Moon RT. 2009. Proximal events in Wnt signal transduction. Nat. Rev. Mol. Cell Biol. 10:468–477. 2. MacDonald BT, Tamai K, He X. 2009. Wnt/beta-catenin signaling: components, mechanisms, and diseases. Dev. Cell 17:9–26. 3. Lai SL, Chien AJ, Moon RT. 2009. Wnt/Fz signaling and the cytoskeleton: potential roles in tumorigenesis. Cell Res. 19:532–545. Cˇajánek et al. 68 mcb.asm.org Molecular and Cellular Biology onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom 4. Schlessinger K, Hall A, Tolwinski N. 2009. Wnt signaling pathways meet Rho GTPases. Genes Dev. 23:265–277. 5. Montcouquiol M, Crenshaw EB, III, Kelley MW. 2006. Noncanonical Wnt signaling and neural polarity. Annu. Rev. Neurosci. 29:363–386. 6. Andersson ER, Prakash N, Cajanek L, Minina E, Bryja V, Bryjova L, Yamaguchi TP, Hall AC, Wurst W, Arenas E. 2008. Wnt5a regulates ventral midbrain morphogenesis and the development of A9-A10 dopaminergic cells in vivo. PLoS One 3:e3517. doi:10.1371/journal.pone. 0003517. 7. Inestrosa NC, Arenas E. 2010. Emerging roles of Wnts in the adult nervous system. Nat. Rev. Neurosci. 11:77–86. 8. Schulte G, Bryja V, Rawal N, Castelo-Branco G, Sousa KM, Arenas E. 2005. Purified Wnt-5a increases differentiation of midbrain dopaminergic cells and dishevelled phosphorylation. J. Neurochem. 92:1550–1553. 9. Bryja V, Schulte G, Rawal N, Grahn A, Arenas E. 2007. Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1dependent mechanism. J. Cell Sci. 120:586–595. 10. Castelo-Branco G, Wagner J, Rodriguez FJ, Kele J, Sousa K, Rawal N, Pasolli HA, Fuchs E, Kitajewski J, Arenas E. 2003. Differential regulation of midbrain dopaminergic neuron development by Wnt-1, Wnt-3a, and Wnt-5a. Proc. Natl. Acad. Sci. U. S. A. 100:12747–12752. 11. Hayashi H, Morizane A, Koyanagi M, Ono Y, Sasai Y, Hashimoto N, Takahashi J. 2008. Meningeal cells induce dopaminergic neurons from embryonic stem cells. Eur. J. Neurosci. 27:261–268. 12. Parish CL, Castelo-Branco G, Rawal N, Tonnesen J, Sorensen AT, Salto C, Kokaia M, Lindvall O, Arenas E. 2008. Wnt5a-treated midbrain neural stem cells improve dopamine cell replacement therapy in parkinsonian mice. J. Clin. Invest. 118:149–160. 13. Jaffe AB, Hall A. 2005. Rho GTPases: biochemistry and biology. Annu. Rev. Cell Dev. Biol. 21:247–269. 14. Rossman KL, Der CJ, Sondek J. 2005. GEF means go: turning on RHO GTPases with guanine nucleotide-exchange factors. Nat. Rev. Mol. Cell Biol. 6:167–180. 15. Habas R, Dawid IB, He X. 2003. Coactivation of Rac and Rho by Wnt/ Frizzled signaling is required for vertebrate gastrulation. Genes Dev. 17: 295–309. 16. Rosso SB, Sussman D, Wynshaw-Boris A, Salinas PC. 2005. Wnt signaling through Dishevelled, Rac and JNK regulates dendritic development. Nat. Neurosci. 8:34–42. 17. Gao C, Chen YG. 2010. Dishevelled: the hub of Wnt signaling. Cell Signal. 22:717–727. 18. Wallingford JB, Habas R. 2005. The developmental biology of Dishevelled: an enigmatic protein governing cell fate and cell polarity. Development 132:4421–4436. 19. Cong F, Schweizer L, Varmus H. 2004. Casein kinase Iepsilon modulates the signaling specificities of Dishevelled. Mol. Cell. Biol. 24:2000–2011. 20. Habets GG, Scholtes EH, Zuydgeest D, van der Kammen RA, Stam JC, Berns A, Collard JG. 1994. Identification of an invasion-inducing gene, Tiam-1, that encodes a protein with homology to GDP-GTP exchangers for Rho-like proteins. Cell 77:537–549. 21. Bryja V, Cajanek L, Grahn A, Schulte G. 2007. Inhibition of endocytosis blocks Wnt signalling to beta-catenin by promoting Dishevelled degradation. Acta Physiol. (Oxf.) 190:55–61. 22. Lee JS, Ishimoto A, Yanagawa S. 1999. Characterization of mouse Dishevelled (Dvl) proteins in Wnt/Wingless signaling pathway. J. Biol. Chem. 274:21464–21470. 23. Angers S, Thorpe CJ, Biechele TL, Goldenberg SJ, Zheng N, Maccoss MJ, Moon RT. 2006. The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wnt-beta-catenin pathway by targeting Dishevelled for degradation. Nat. Cell Biol. 8:348–357. 24. Chen W, ten Berge D, Brown J, Ahn S, Hu LA, Miller WE, Caron MG, Barak LS, Nusse R, Lefkowitz RJ. 2003. Dishevelled 2 recruits betaarrestin 2 to mediate Wnt5A-stimulated endocytosis of Frizzled 4. Science 301:1391–1394. 25. Schwarz-Romond T, Fiedler M, Shibata N, Butler PJ, Kikuchi A, Higuchi Y, Bienz M. 2007. The DIX domain of Dishevelled confers Wnt signaling by dynamic polymerization. Nat. Struct. Mol. Biol. 14:484–492. 26. Tolias KF, Bikoff JB, Burette A, Paradis S, Harrar D, Tavazoie S, Weinberg RJ, Greenberg ME. 2005. The Rac1-GEF Tiam1 couples the NMDA receptor to the activity-dependent development of dendritic arbors and spines. Neuron 45:525–538. 27. Ridley AJ, Paterson HF, Johnston CL, Diekmann D, Hall A. 1992. The small GTP-binding protein Rac regulates growth factor-induced membrane ruffling. Cell 70:401–410. 28. Badura L, Swanson T, Adamowicz W, Adams J, Cianfrogna J, Fisher K, Holland J, Kleiman R, Nelson F, Reynolds L, St Germain K, Schaeffer E, Tate B, Sprouse J. 2007. An inhibitor of casein kinase I epsilon induces phase delays in circadian rhythms under free-running and entrained conditions. J. Pharmacol. Exp. Ther. 322:730–738. 29. Rena G, Bain J, Elliott M, Cohen P. 2004. D4476, a cell-permeant inhibitor of CK1, suppresses the site-specific phosphorylation and nuclear exclusion of FOXO1a. EMBO Rep. 5:60–65. 30. Cajanek L, Ribeiro D, Liste I, Parish CL, Bryja V, Arenas E. 2009. Wnt/beta-catenin signaling blockade promotes neuronal induction and dopaminergic differentiation in embryonic stem cells. Stem Cells 27: 2917–2927. 31. Bryja V, Schambony A, Cajanek L, Dominguez I, Arenas E, Schulte G. 2008. Beta-arrestin and casein kinase 1/2 define distinct branches of noncanonical WNT signalling pathways. EMBO Rep. 9:1244–1250. 32. Parish CL, Parisi S, Persico MG, Arenas E, Minchiotti G. 2005. Cripto as a target for improving embryonic stem cell-based therapy in Parkinson’s disease. Stem Cells 23:471–476. 33. Dull T, Zufferey R, Kelly M, Mandel RJ, Nquyen M, Trono D, Naldini L. 1998. A third-generation lentivirus vector with a conditional packaging system. J. Virol. 72:8463–8471. 34. Yokoyama N, Golebiewska U, Wang HY, Malbon CC. 2010. Wntdependent assembly of supermolecular Dishevelled-3-based complexes. J. Cell Sci. 123:3693–3702. 35. Schwarz-Romond T, Merrifield C, Nichols BJ, Bienz M. 2005. The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J. Cell Sci. 118:5269– 5277. 36. Smalley MJ, Signoret N, Robertson D, Tilley A, Hann A, Ewan K, Ding Y, Paterson H, Dale TC. 2005. Dishevelled (Dvl-2) activates canonical Wnt signalling in the absence of cytoplasmic puncta. J. Cell Sci. 118:5279– 5289. 37. Garcia-Mata R, Burridge K. 2007. Catching a GEF by its tail. Trends Cell Biol. 17:36–43. 38. Marinissen MJ, Gutkind JS. 2005. Scaffold proteins dictate Rho GTPasesignaling specificity. Trends Biochem. Sci. 30:423–426. 39. Bernatik O, Ganji R, Dijksterhuis J, Konik P, Cervenka I, Polonio T, Krejci P, Schulte G, Bryja V. 2011. Sequential activation and inactivation of Dishevelled in the Wnt/beta-catenin pathway by casein kinases. J. Biol. Chem. 286:10396–10410. 40. Klein TJ, Jenny A, Djiane A, Mlodzik M. 2006. CKIepsilon/discs overgrown promotes both Wnt-Fz/beta-catenin and Fz/PCP signaling in Drosophila. Curr. Biol. 16:1337–1343. 41. Etienne-Manneville S, Hall A. 2002. Rho GTPases in cell biology. Nature 420:629–635. 42. de Curtis I. 2008. Functions of Rac GTPases during neuronal development. Dev. Neurosci. 30:47–58. 43. Govek EE, Newey SE, Van Aelst L. 2005. The role of the Rho GTPases in neuronal development. Genes Dev. 19:1–49. 44. Tahirovic S, Hellal F, Neukirchen D, Hindges R, Garvalov BK, Flynn KC, Stradal TE, Chrostek-Grashoff A, Brakebusch C, Bradke F. 2010. Rac1 regulates neuronal polarization through the WAVE complex. J. Neurosci. 30:6930–6943. 45. Ishida-Takagishi M, Enomoto A, Asai N, Ushida K, Watanabe T, Hashimoto T, Kato T, Weng L, Matsumoto S, Asai M, Murakumo Y, Kikuchi A, Takahashi M. 2012. The Dishevelled-associated protein Daple controls the noncanonical Wnt/Rac pathway and cell motility. Nat. Commun. 3:89. 46. Luo ZG, Wang Q, Zhou JZ, Wang J, Luo Z, Liu M, He X, WynshawBoris A, Xiong WC, Lu B, Mei L. 2002. Regulation of AChR clustering by Dishevelled interacting with MuSK and PAK1. Neuron 35:489–505. 47. Tanegashima K, Zhao H, Dawid IB. 2008. WGEF activates Rho in the Wnt-PCP pathway and controls convergent extension in Xenopus gastrulation. EMBO J. 27:606–617. 48. Tsuji T, Ohta Y, Kanno Y, Hirose K, Ohashi K, Mizuno K. 2010. Involvement of p114-RhoGEF and Lfc in Wnt-3a- and Dishevelledinduced RhoA activation and neurite retraction in N1E-115 mouse neuroblastoma cells. Mol. Biol. Cell 21:3590–3600. 49. Woodcock SA, Jones RC, Edmondson RD, Malliri A. 2009. A modified tandem affinity purification technique identifies that 14-3-3 proteins inTiam1 Regulates Noncanonical Wnt Signaling January 2013 Volume 33 Number 1 mcb.asm.org 69 onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom teract with Tiam1, an interaction which controls Tiam1 stability. J. Proteome Res. 8:5629–5641. 50. Worthylake DK, Rossman KL, Sondek J. 2000. Crystal structure of Rac1 in complex with the guanine nucleotide exchange region of Tiam1. Nature 408:682–688. 51. Buchsbaum RJ, Connolly BA, Feig LA. 2003. Regulation of p70 S6 kinase by complex formation between the Rac guanine nucleotide exchange factor (Rac-GEF) Tiam1 and the scaffold spinophilin. J. Biol. Chem. 278: 18833–18841. 52. Swiatek W, Tsai IC, Klimowski L, Pepler A, Barnette J, Yost HJ, Virshup DM. 2004. Regulation of casein kinase I epsilon activity by Wnt signaling. J. Biol. Chem. 279:13011–13017. 53. Fanto M, Weber U, Strutt DI, Mlodzik M. 2000. Nuclear signaling by Rac and Rho GTPases is required in the establishment of epithelial planar polarity in the Drosophila eye. Curr. Biol. 10:979–988. 54. Leeuwen FN, Kain HE, Kammen RA, Michiels F, Kranenburg OW, Collard JG. 1997. The guanine nucleotide exchange factor Tiam1 affects neuronal morphology; opposing roles for the small GTPases Rac and Rho. J. Cell Biol. 139:797–807. 55. Tanaka M, Ohashi R, Nakamura R, Shinmura K, Kamo T, Sakai R, Sugimura H. 2004. Tiam1 mediates neurite outgrowth induced by ephrin-B1 and EphA2. EMBO J. 23:1075–1088. 56. Kunda P, Paglini G, Quiroga S, Kosik K, Caceres A. 2001. Evidence for the involvement of Tiam1 in axon formation. J. Neurosci. 21:2361–2372. 57. Kawauchi T, Chihama K, Nabeshima Y, Hoshino M. 2003. The in vivo roles of STEF/Tiam1, Rac1 and JNK in cortical neuronal migration. EMBO J. 22:4190–4201. 58. Mertens AE, Pegtel DM, Collard JG. 2006. Tiam1 takes PARt in cell polarity. Trends Cell Biol. 16:308–316. 59. Minobe S, Sakakibara A, Ohdachi T, Kanda R, Kimura M, Nakatani S, Tadokoro R, Ochiai W, Nishizawa Y, Mizoguchi A, Kawauchi T, Miyata T. 2009. Rac is involved in the interkinetic nuclear migration of cortical progenitor cells. Neurosci. Res. 63:294–301. 60. Schlessinger K, McManus EJ, Hall A. 2007. Cdc42 and noncanonical Wnt signal transduction pathways cooperate to promote cell polarity. J. Cell Biol. 178:355–361. 61. Zhang X, Zhu J, Yang GY, Wang QJ, Qian L, Chen YM, Chen F, Tao Y, Hu HS, Wang T, Luo ZG. 2007. Dishevelled promotes axon differentiation by regulating atypical protein kinase C. Nat. Cell Biol. 9:743–754. 62. Sasaki N, Kurisu J, Kengaku M. 2010. Sonic hedgehog signaling regulates actin cytoskeleton via Tiam1-Rac1 cascade during spine formation. Mol. Cell. Neurosci. 45:335–344. Cˇajánek et al. 70 mcb.asm.org Molecular and Cellular Biology onDecember11,2012byguesthttp://mcb.asm.org/Downloadedfrom Vítězslav Bryja, 2014    Attachments      #15      A. Soldano, Z. Okray, P. Janovská, K. Tmejová, E. Reynaud, A. Claeys, J. Yan, B. De  Strooper, J.‐M. Dura, V. Bryja, B. A. Hassan (2013): The Amyloid Precursor Proteins are  conserved modulators of the Wnt/PCP pathway required for robustness of axonal  outgrowth. PLoS Biol. 11(5):e1001562.       Impact factor (2012): 12.690  Times cited (without autocitations, WoS, Feb 21st 2014):  0  Significance: This paper for the first identifies endogenous role of the family of amyloid  precursor proteins (APP). APP and APP‐like proteins are well characterized in the  pathogenesis  of  Alzheimer  disease,  where  they  significantly  contribute  to  the  formation of amyloid plaques. Here we showed that proteins from APP family are  regulators  of  non‐canonical  Wnt  pathway  and  are  required  for  proper  signal  transduction between the Wnt ligand and Dishevelled.  Contibution of the author/author´s team: Biochemical and functional analysis of the  role of APP/APLP2 in the mammalian model system.                The Drosophila Homologue of the Amyloid Precursor Protein Is a Conserved Modulator of Wnt PCP Signaling Alessia Soldano1,2,3 , Zeynep Okray1,2,3 , Pavlina Janovska4 , Katerˇina Tmejova´4 , Elodie Reynaud5,6 , Annelies Claeys1,2 , Jiekun Yan1,2 , Zeynep Kalender Atak2 , Bart De Strooper1,2,3 , Jean-Maurice Dura5,6 , Vı´teˇzslav Bryja4,7 , Bassem A. Hassan1,2,3 * 1 VIB Center for the Biology of Disease, Leuven, Belgium, 2 Center for Human Genetics, University of Leuven School of Medicine, Leuven, Belgium, 3 Doctoral Program in Molecular and Developmental Genetics, University of Leuven Group Biomedicine, Leuven, Belgium, 4 Institute of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republic, 5 Institut de Ge´ne´tique Humaine/Centre National de la Recherche Scientifique UPR1142, Montpellier, France, 6 Laboratoire Neuroge´ne´tique et Me´moire, De´partement Ge´ne´tique et De´veloppement, Montpellier, France, 7 Institute of Biophysics of the Academy of Sciences of the Czech Republic, Brno, Czech Republic Abstract Wnt Planar Cell Polarity (PCP) signaling is a universal regulator of polarity in epithelial cells, but it regulates axon outgrowth in neurons, suggesting the existence of axonal modulators of Wnt-PCP activity. The Amyloid precursor proteins (APPs) are intensely investigated because of their link to Alzheimer’s disease (AD). APP’s in vivo function in the brain and the mechanisms underlying it remain unclear and controversial. Drosophila possesses a single APP homologue called APP Like, or APPL. APPL is expressed in all neurons throughout development, but has no established function in neuronal development. We therefore investigated the role of Drosophila APPL during brain development. We find that APPL is involved in the development of the Mushroom Body ab neurons and, in particular, is required cell-autonomously for the baxons and non-cell autonomously for the a-axons growth. Moreover, we find that APPL is a modulator of the Wnt-PCP pathway required for axonal outgrowth, but not cell polarity. Molecularly, both human APP and fly APPL form complexes with PCP receptors, thus suggesting that APPs are part of the membrane protein complex upstream of PCP signaling. Moreover, we show that APPL regulates PCP pathway activation by modulating the phosphorylation of the Wnt adaptor protein Dishevelled (Dsh) by Abelson kinase (Abl). Taken together our data suggest that APPL is the first example of a modulator of the Wnt-PCP pathway specifically required for axon outgrowth. Citation: Soldano A, Okray Z, Janovska P, Tmejova´ K, Reynaud E, et al. (2013) The Drosophila Homologue of the Amyloid Precursor Protein Is a Conserved Modulator of Wnt PCP Signaling. PLoS Biol 11(5): e1001562. doi:10.1371/journal.pbio.1001562 Academic Editor: Konrad Basler, University of Zurich, Switzerland Received December 17, 2012; Accepted April 2, 2013; Published May 14, 2013 Copyright: ß 2013 Soldano et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Funding: This work was supported by VIB (BAH and BDS), Concerted Research Action (GOA) and Methusalem grants from the KU Leuven (BDS and BAH), Fonds Wetenschappelijke Oderzoeks (FWO) grants G.0543.08, G.0680.10, G.0681.10, and G.0503.12 (BAH), grant ANR-07-NEURO-034 from the Agence Nationale pour la Recherche (JMD) and the Czech Science Foundation (grants 204/09/0498, 301/11/0747), Ministry of Education, Youth and Sports of the Czech Republic (grant MSM0021622430), Academy of Sciences of the Czech Republic (AVOZ50040507, AVOZ50040702), and an EMBO Installation Grant (VB). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript Competing Interests: The authors have declared that no competing interests exist. Abbreviations: Abl, Abelson kinase; AD, Alzheimer’s disease; APP, Amyloid precursor protein; APP-C99, C-terminus of human APP; APPL, APP Like; ApplDC, APPL delta C-terminal domain; dKO, double knock-out; Dsh, Drosophila Dishevelled; Dvl, mammalian Dishevelled; Dvl2, Dishevelled2; FasII, FascilinII; Fmi, Flamingo; Fz, Frizzled; Fzd5, Frizzled 5; MB, Mushroom Bodies; MEF, mouse embryonic fibroblast; PCP, Wnt Planar Cell Polarity; sAPPL, soluble APPL; Vangl2, Van Gogh2; Vang/ Stbm, Van Gogh/Strabismus. * E-mail: Bassem.Hassan@cme.vib-kuleuven.be Introduction The Wnt Planar Cell Polarity (PCP) pathway is a highly conserved regulator of cellular orientation within the plane of an epithelium [1,2]. Genetic and molecular studies in Drosophila indicate Disheveled (Dsh), a cytoplasmic transducer of Wnt signaling; Frizzled (Fz), a seven-transmembrane receptor for Wnt ligands; and Van Gogh (Vang), a four-pass transmembrane protein, as core Wnt-PCP proteins. Intriguingly, the Wnt-PCP pathway regulates axon outgrowth rather than neuronal polarity during brain development of both vertebrates and Drosophila [3–5]. The Amyloid Precursor Protein (APP) is a member of a highly conserved family of type I transmembrane proteins that includes APP, APLP1, and APLP2 [6] in mammals and APP-Like, or APPL, in Drosophila melanogaster [7]. APP proteins show not only structural but also functional conservation, as exemplified by the ability of human APP to rescue behavioral phenotypes of APPL null flies [8]. APP is the subject of intense research because of genetic and biochemical links to Alzheimer’s disease (AD), whereby the proteolytic processing of APP generates the Amyloid Beta peptide whose accumulation in the brain is widely thought to induce neurodegeneration [9–12]. Despite these efforts, the normal physiological function of APP in vivo in the nervous system remains elusive and highly controversial. This is due to the lack of a consensus over the neuronal phenotypes in null mutant animals and the mechanism of APP action in vivo. APP knock-out mice are viable and show developmental neuronal deficits, namely cortical migration and agenesis of the corpus callosum, at variable penetrance depending on genetic background [13–15]. In contrast, another in vivo report, based mainly on gain of function PLOS Biology | www.plosbiology.org 1 May 2013 | Volume 11 | Issue 5 | e1001562 and RNA interference experiments, suggests that APP may be required for developmental axonal degeneration [16], although it is unclear whether APP knock-out mice show these phenotypes. Finally, initial findings proposed an axonal transport function for APP [17] but later studies strongly questioned the presence of these defects in APP knock-out mice [18]. Mechanistically, there is also disagreement on whether APP acts cell autonomously or non autonomously. For example, an extensive network of molecular interactions has been described for the intracellular domain of APP [19], yet a knock-in of APP lacking the intracellular domain appears to rescue physiological and learning deficits reported in APP knock-out mice, suggesting that the intracellular domain is dispensable [20]. Because current models based on Amyloid toxicity do not provide a complete explanation for the onset of neuronal dysfunction in AD, it has long been argued that greater attention needs to shift towards understanding the normal physiological function of APP in order to assess its potential contribution to AD pathology [21,22]. Therefore, a mechanistic understanding of the in vivo physiological function of APP proteins is of paramount importance. To elucidate the function and mechanism of action of APP proteins in vivo, we first investigated Drosophila APPL, the APP homologue in the fruit fly, as a model system. We show that APPL is a novel neuronal-specific modulator of the PCP pathway required for the robustness of axonal outgrowth during the development of the Mushroom Bodies (MB), a Drosophila center for learning and memory. APPL carries out this function through facilitating the PCP-specific phosphorylation of the Wnt adaptor protein Dishevelled (Dsh/Dvl) by the Abelson kinase (Abl). Furthermore, we show that APPL is part of the membrane complex formed by WntPCP core proteins. Finally, biochemical and cell biological analyses show that human APP immunoprecipitates mammalian PCP proteins and that APP proteins are necessary for Dvl phosphorylation in response to the PCP ligand Wnt5a. Therefore, the APP proteins represent a novel and conserved family of neuronal modulators of Wnt-PCP signaling required for the robustness of brain wiring during development. Results APPL Is a Robustness Factor Required Cell-Autonomously for MB b-Lobe Development Drosophila APPL is a neuronal-specific protein expressed in most, if not all, neurons throughout development and adult life. In particular, APPL is highly expressed in the developing Drosophila MB, especially the so-called ab neurons (Figure 1A–D). Flies null for Appl (henceforth Appl2/2 ) are viable, fertile, and reported to show no gross structural defects in the brain [8]. While a requirement for APPL in learning and memory specifically in adult flies has been shown [23], the function of its pan-neuronal expression throughout development remains unknown, as does the in vivo mechanism of its action(s). We began addressing the function of APPL in neuronal development by carefully examining the development of the Drosophila MB in Appl2/2 mutant flies. The MB derives from two groups of four neuroblasts, one in each hemisphere, that sequentially generate three subsets of neurons: the c-, a9b9 and ab neurons, where APPL is highly expressed. Each ab neuron projects an axon that branches into a dorsal ‘‘a branch’’ and a medial ‘‘b branch.’’ The fascicles generated by each of these branches are referred to as the a lobe and b lobe. The a and b lobes can be easily visualized using the anti-FascilinII (FasII) antibody. The lobes were present and morphologically normal in 97 adult control animals (Figure 1E) examined. In contrast, 26% of Appl2/2 brains examined (n = 101) showed axonal defects (Figure S1). Specifically, 14% of the brains show a-lobe loss (Figure 1F), whereas 12% of the brains show b-lobe loss (Figure 1G). These defects are developmental in origin as they can be observed during ab lobe formation at 48 h of pupal development (Figure 1H–J). To ascertain whether APPL acts cell autonomously to regulate MB axonal outgrowth, we generated GFP-marked single Appl2/2 cell clones using the MARCM technique [24]. While none of the control clones showed any defects (Figure 2A), 10% of the mutant clones showed lack of blobe growth (Figures 2B and S2A), similar to the penetrance observed in null mutant brains. However, none of the mutant clones showed loss of a-lobe growth. Taken together, these data indicate that APPL is required for normal MB axonal outgrowth and that it is required cell-autonomously for the growth of the blobe and non-cell autonomously for the growth of a lobe. To verify the specificity of the Appl2/2 phenotype, we rescued the defects by restoring APPL expression in ab neurons. Expression of full-length membrane-bound APPL, but not a secreted form or a form lacking the intracellular domain (ApplDC), strongly suppresses the b-lobe loss phenotype (Figures 2C–H and S2B). These results indicate that APPL is required as a full-length, membrane-tethered protein and that the intracellular signaling downstream of APPL is necessary for normal b-lobe outgrowth. Interestingly, secreted APPL strongly reduces the loss of the a lobe (Figure S2C and S2D), confirming that APPL acts non-autonomously in a-lobe outgrowth. All together, the results suggest that APPL is a robustness factor for an unknown axon growth signal whereby Appl2/2 ab neurons are at a phenotypic threshold that causes them to fail to grow in approximately 26% of the cases. Abelson Kinase Is a Downstream Effector of APPL Required for MB Axon Outgrowth To unravel the mechanism by which APPL supports MB axon outgrowth, we chose to focus on the cell-autonomous function of APPL in the b lobe. A previous study using APPL gain of function indicates that APPL overexpression induces axonal outgrowth that is dependent on Abl kinase activity [25]. We asked whether Abl Author Summary Wnt Planar Cell Polarity (PCP) signaling is a universal regulator of polarity in epithelial cells, but in neurons it regulates axon outgrowth, suggesting the existence of axonal modulators of Wnt-PCP activity. The Amyloid Precursor Proteins (APPs) are intensely investigated because of their link to Alzheimer’s disease (AD). APP’s in vivo function in the brain and the mechanisms underlying it remain unclear and controversial. In the present work we investigate the role of the Drosophila neuron-specific APP homologue, called APPL, during brain development. We find that APPL is required for the development of ab neurons in the mushroom body, a structure critical for learning and memory. We find that APPL is a modulator of the Wnt-PCP pathway required for axonal outgrowth, but not for cell polarity. Molecularly, both human APP and fly APPL are found in membrane complexes with PCP receptors. Moreover, we show that APPL regulates PCP pathway activation through its downstream effector Abelson kinase (Abl), which modulates the phosphorylation of the Wnt adaptor protein Dishevelled (Dsh) and the subsequent activation of Wnt-PCP signaling. Taken together our data suggest that APPL is the first example of a neuron-specific modulator of the Wnt-PCP pathway. APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 2 May 2013 | Volume 11 | Issue 5 | e1001562 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 3 May 2013 | Volume 11 | Issue 5 | e1001562 kinase also acts downstream of APPL during MB b-lobe growth. First, we tested if APPL genetically interacts with Abl. To this end we analyzed the adult MB morphology of Appl2/2 ; Abl2/+ flies. Loss of one copy of Abelson causes a dramatic increase (up to 51%) in complete (41%) or partial (10%) b-lobe loss, compared to Appl2/2 alone (Figure 3A, 3B and 3E, 3F). As controls we analyzed the siblings heterozygous for both Appl and Abl (Appl2/+ ; Abl2/+ ) and observed no phenotypes (Figure 3F). Similarly to Appl2/2 mutants alone, the phenotypes arise at early developmental stages (Figure S3A). To further confirm that Abl is the downstream mediator of APPL signaling during b-lobe growth, we tested if overexpression of Abl specifically in MB ab neurons rescues the Appl null phenotype. We find that wild-type Abl, but not a Kinase Dead (Abl-KD) form of Abl, rescues the Appl null phenotype (Figure 3C–F) to the same extent as MB ab expression of APPL itself (Figure S2B). Taken together these data indicate that Abl is the effector of APPL required for the b-lobe growth. Next, we further characterized the downstream pathway involved. APPL Is a Novel Neuronal-Specific Modulator of PCP Signaling It has been recently shown that Abl phosphorylates Disheveled (Dsh), a core intracellular component of the Wnt pathway, on the Tyrosine 473. This modification is required for the efficient activation of the PCP signaling pathway in epithelial cells [26]. Interestingly, in the nervous system the Wnt-PCP pathway is required for robust axonal outgrowth both in Drosophila [27] and mouse [28,29]. More recently, several PCP pathway components, like the Wnt receptor Frizzled (Fz), Flamingo (Fmi), Strabismus (Stbm or Vang), and Dsh, have been shown to play a role in the correct targeting and bifurcation of MB axons [4,30,31]. Indeed, we observe that flies harboring a PCP-specific mutation in Dsh (dsh1 ) [32] show the same MB developmental defects observed in Appl2/2 flies (Figure 4A). Together, these observations prompted us to speculate that APPL acts to facilitate Wnt-PCP pathway activation during MB development by mediating Dsh phosphorylation by Abl. To verify this hypothesis we tested if the phosphorylation of Dsh on Y473 is required for MB development. We analyzed dsh1 flies harboring one of two genomic rescue constructs: a wild-type construct (dshDshGFP-wt) or a Tyrosine 473 phospho-mutant construct (dshDshGFP-Y473F). Whereas the restoration of wild-type Dsh fully rescues the dsh1 MB phenotype (Figures 4B and S4A), DshY473F completely fails to do so (Figures 4C and S4A). To exclude that DshY473F failure to rescue the phenotype is due to a perturbation in the expression pattern resulting from the point mutation, we performed anti-GFP staining on adult brains of both transgenic lines. As shown in Figure 4D and 4E, both wild type and mutant Dsh are expressed in the MB lobes. These results clearly indicate that Abl phosphorylation of Dsh on Tyrosine 473 and the subsequent activation of Wnt-PCP signaling are required for the b-lobe growth. Collectively, the data suggest the exciting possibility that APPL may be a neuronal-specific component of the Wnt-PCP pathway. To address this issue, we first asked if APPL interacts genetically with core members of the Wnt-PCP pathway like the classical Wnt-PCP receptor Fz and the canonical Wnt-PCP protein Van Gogh/Strabismus (Vang/Stbm). We first analyzed the b-lobe of Appl2/2 ; Fz2/+ flies. Reduction of Fz in the APPL null background increases the frequency of the b lobe loss up to 21%, whereas no phenotype is observed in control siblings (Figures 5B, E and S5A). This interaction is specific to the MB because expression of a dominant negative form of Fz (Fz-DN) in Appl2/2 ab neurons yields similar results (Figures 5C, 5E and S5A). In both experiments described, the increase in b-lobe loss is relatively mild compared to the dramatic increase in b-lobe loss due to APPL-Abl epistasis for example, suggesting that APPL and Fz may act together for Wnt-PCP activation. To clarify this, we tested if inhibition of Fz alone is sufficient to induce the b phenotype. Overexpression of Fz-DN in ab neurons of wild-type flies did not cause any morphological defects (Figure S5A and S5B). These data were further confirmed by the analysis of Fz mutant ab clones of different sizes. None of the analyzed MB clones showed morphologically aberrant axons (Figure S5C). To rule out compensation by Fz2, we examined Fz2 expression in the brain and found that it is not detectable in MB (unpublished data). Furthermore, MARCM clones null for Fz2 alone or Fz and Fz2 together show no aberrant morphology (Figure S5D and S5E). Therefore, APPL function is a critical determinant of the role of PCP in the outgrowth of MB b axons. To further ascertain the interaction with the Wnt-PCP pathway, we analyzed if APPL interacts with the Wnt-PCP four-pass transmembrane protein Vang/Stbm. Reduction of Van Gogh in the Appl null background (Appl2/2 ;vang2/+ ) increases the frequency of the b-lobe loss phenotype to 33% (Figures 5D,E and S5A), whereas no phenotype is observed in control siblings. APPL is also expressed in the developing fly retina (Figure S5F), where the PCP pathway regulates the polarity of photoreceptor cells. However, we did not observe defects in photoreceptor polarity (Figure S5G–I), suggesting that the role of APPL in Wnt-PCP signaling is specific to axonal outgrowth. Together, the data above identify APPL as the first neuronal-specific modulator of the Wnt-PCP pathway’s role in axonal outgrowth. Next, we analyzed if the expression pattern of Vang and APPL overlaps during MB development. For this purpose, we used a line that expresses a YFP tagged form of Vang under the control of the Actin promoter. As shown in Figure 5F, during the development of the b lobe both APPL and Vang are expressed at a high level in the growing axons. Interestingly, in adult stage, APPL expression is reduced in the rest of the brain and enriched in the ab neurons while Vang levels are strongly reduced Figure 1. APPL is a robustness factor required for Mushroom Bodies development. (A–D) Developmental APPL expression in MB. Immunofluorescence analysis using anti-APP-Cterm (A–D in magenta) and anti-FasII (A9–D9 in green) antibodies. The A-–D- panels show the merge of the FasII and GFP channel for the indicated samples. APPL is expressed at high levels in the developing brain and is enriched along MB axons both during development and in adult stages. (A, B) 48 h APF MB lobes. The images are single confocal stacks; zoom shows 636magnification of the boxed area (scale bar, 50 mm in A, 25 mm in B). (C) Adult MB axons. The images are single confocal stacks (scale bar, 50 mm). (D) Zoom shows 636magnification of the boxed area in panel C. The images are z-projections of two confocal image stacks (step size, 0.6 mm; scale bar, 25 mm). (E–G) Adult MB lobes labeled with FascilinII antibody (FasII). All images are z-projections of confocal image stacks (scale bar, 50 mm). (E) Morphologically normal a/b neurons in control adult brain. (F–G) Structure of a/b neurons in Appld w* null mutant adult brains. In the absence of APPL, MB lobes show an aberrant pattern of growth in 26% of the analyzed sample (n = 101). In particular, in 14% of the cases (F), the a lobe fails to project towards the dorsal side of the brains (as indicated by the arrow), whereas in 12% of the cases (G), the b lobe fails to project towards the midline (as indicated by the arrow). (H) Morphologically normal a/b neurons of a 48APF Canton S brain. (I–J) Morphologically aberrant a/b neurons of an Appld w*48APF brain. The a- (I) and b-lobe (J) loss observed in the adult brain is already present at 48 APF, thus suggesting that is not due to degeneration but rather to failure in axon growth. doi:10.1371/journal.pbio.1001562.g001 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 4 May 2013 | Volume 11 | Issue 5 | e1001562 Figure 2. APPL is cell-autonomously required for b-lobe development. (A–B) Z-projections of confocal image stacks of GFP-labeled clones. Recombination was induced at 0–24 h APF. Immunofluorescence analysis of adult MBs using anti-GFP (green) and anti-FasII (A9, B9 in magenta) antibodies (scale bar, 50 mm). The A0 and B0 panels show the merge of the FasII and GFP channel for the indicated samples. (A) Morphologically normal a/b neurons in single-cell control clones obtained by crossing FRT19A; ry506 flies with FRT19A,tub-Gal80,hsFLP/FM7;UAS-CD8-GFP/CyO;OK107 (n = 41 clones). (B) Appld single-cell clones. Clones were obtained by crossing Appld FRT19A/FM7 with FRT19A, tub-Gal80, hsFLP/FM7; UAS-CD8-GFP/ CyO; OK107. The cell-autonomous loss of APPL leads to b-lobe loss in 10% of the clones analyzed as indicated by the arrow (n = 44 clones). (C–G) Adult MB lobes labeled with FasII antibody. All the images are z-projections of confocal image stacks (scale bar, 50 mm). (C) Morphologically normal a/ b neurons in a control adult brain. (D) Morphologically aberrant ab neurons in Appld w*/Y;L/+;P247Gal4 analyzed as a control for the rescue experiment (n = 47). In 13% of the brains the b axons failed to grow towards the midline. (E) Morphologically normal ab neurons in Appld w*/Y;UASAppl/+;P247Gal4 adult brains. The re-introduction of full-length APPL in MBs during development is sufficient to rescue the b-lobe defect. Only 2% of the samples show a phenotype; p value = 0.03467 calculated with G-test (n = 50). (F) Morphologically aberrant ab neurons in Appld w*/Y;UAS-sAppl/ +;P247Gal4. The re-introduction of a secreted form of APPL fails to rescue the b-lobe defect, where 12% of the analyzed brains showed phenotype. p value = 0.9831 calculated with G-test (n = 50). (G) Morphologically aberrant ab neurons in Appld w*/Y;UAS-ApplDC/+;P247Gal4. The re-introduction of a form of APPL lacking the C-terminal domain fails to rescue the b-lobe defect, where 11% of the analyzed brains showed phenotype (n = 54); p value = 0.8863 calculated with G-test. (H) The graph shows the penetrance of the b-lobe loss in the rescue flies normalized against the penetrance of the phenotype in the Appld w* background. * Indicates a p value,0.05 calculated with G-test. doi:10.1371/journal.pbio.1001562.g002 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 5 May 2013 | Volume 11 | Issue 5 | e1001562 in these axons (Figure 5G). Moreover, in the developing fly retina, APPL and Vang do not colocalize and are found in juxtaposed domains (Figure S5L). Taken together these results suggest that both APPL and Vang are present in developing ab axons where their genetic interaction is required for the correct development of the b axons. On the contrary, the two proteins are expressed in different compartments in the fly retina where APPL function is not required for PCP activity. Finally, to confirm that PCP signaling is indeed positively modulated by APPL and that the activation of the signaling is required for the b-lobe growth, we performed rescue experiments with Dsh. As shown in Figure 5F, overexpression of Dsh in the Appl2/2 background strongly reduces (7%) the b-lobe loss but does not fully rescue the phenotype. This result indicates that the phosphorylation of Dsh by Abelson is the limiting step in the activation of PCP signaling; increasing the amount of Dsh present in the neurons improves the phenotype, but probably the endogenous Abelson is not sufficient to phosphorylate the whole pool of Dsh. To overcome this problem we decided to enhance the activation of PCP signaling by overexpressing Wnt5 in the Appl2/ 2 background. Interestingly, overexpression of Wnt5 significantly rescues the b-lobe loss (Figure 5H), indicating that the PCP signaling activation is required for the development of the b lobe and is reduced in absence of APPL. APP Is Required for Dvl2 Phosphorylation in Response to Wnt5 and Forms Complexes with PCP Core Proteins The previously described results raise two important questions. First, is the interaction between APPL and the Wnt-PCP pathway conserved in mammalian APP proteins? Second, if so, do the genetic interactions observed in Drosophila reflect a biochemical association of APPL/APP with the Wnt-PCP receptors? To address these issues, we first investigated if mouse APP proteins mediate the phosphorylation of mouse Disheveled (Dvl) in response to the Wnt-PCP ligand Wnt5a. To this end, we analyzed Dvl phosphorylation in response to Wnt5a treatment in wild-type versus APP/APLP2 double knock-out mouse embryonic fibroblast (dKO MEFs). In WT MEFs, Dvl2 phosphorylation increases dramatically upon treatment with Wnt5a. In contrast, Dvl2 in dKO MEFs completely fails to respond to Wnt5a treatment (Figure 6A). The effect on the activation of Dvl2 is a direct consequence of APP loss because upon reintroduction of APP cDNA Dvl2 phosphorylation is restored (Figure 6A). Next, we tested if APPL and APP interacts with core Wnt-PCP receptor proteins. In particular, we performed co-immunoprecipitation (Co-IP) analyses of tagged proteins expressed in HEK-293T cells. As shown in Figures 6B and S6B, APPL immunoprecipitates Vang when the two proteins are co-expressed in the same cells. Drosophila APPL immunoprecipitates human Van Gogh 2 (Vangl2) Figure 3. Abelson kinase is a downstream effector of APPL required for MB axons outgrowth. (A–B) Loss of b lobes in Appld w*/Y;Abl4 /+ adult brains. (A) In 40% of the flies that have lost one copy of Abl in the Appl2/2 background, no axons grow toward the midline (indicated by the arrow), whereas in 10% of the cases (B) only few axons project normally (n = 29); p value = 1.29E-02 calculated with G-test. (C) Morphologically normal ab neurons in Appld w*/Y;UAS-Abl/+;P247Gal4 adult brains. Overexpression of Abl in the MB rescues the b-lobe loss in the Appl2/2 background. Only 2% of the analyzed samples show b-lobe loss (n = 46); p value = 0.03192 calculated with G-test. (D) Morphologically aberrant ab neurons in Appld w*/ Y;UAS-Abl-KD/+;P247Gal4 adult brains. Overexpression of a Kinase dead form of Abl fails to rescue the b-lobe loss in the Appl2/2 background. Thirteen percent of the analyzed brains showed b-lobe loss as indicated by the arrow (n = 31). (E) The graph shows the penetrance of b-lobe loss in the Abl rescue flies and in the flies heterozygous for Abl, normalized against the penetrance observed in the Appld w* background. (F) The table lists the number of brains analyzed in the rescue experiments. * Indicates a p value,0.05 calculated with G-test; ** indicates a p value,0.001 calculated with G-test. doi:10.1371/journal.pbio.1001562.g003 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 6 May 2013 | Volume 11 | Issue 5 | e1001562 (Figures 6C and S6C). Importantly, this multiprotein complex can only be detected when the proteins are expressed in the same cells, but not when lysates from separately expressing cells are mixed (Figure S6E). This observation also accounts for the absence of rescue observed when a form of APPL lacking the C terminal is expressed in the Appl2/2 animals (Figure 2H). Similarly, the membrane tethered C-terminus of human APP (APP-C99) immunoprecipitates Vangl2 (Figures 6D and S6D). Moreover, we tested if APPL also immunoprecipitates other PCP receptors like Fz. As shown in Figures 6E and S6F, APPL immunoprecipitates Fz, suggesting that the core PCP proteins and APPL might form a multiprotein complex on the membrane, responsible for the efficient activation of the pathway. Similarly, the membranetethered C-terminus of human APP (APP-C99) immunoprecipitates human Fzd5 (Figure 6F). Discussion AD is a neurodegenerative disorder characterized by progressive loss of neurons in specific regions of the brain that correlates with progressive impairment of higher cognitive functions. A growing body of evidence identifies the APP and its metabolite the Ab peptide as main players in the pathogenesis of AD. In particular, the accumulation of Ab peptides in the brain seems to be the trigger of the pathological cascade that eventually results in neuronal loss and degeneration [33]. Despite efforts to characterize the molecular mechanisms underlying Ab’s toxic function, it is Figure 4. Dsh phosphorylation is required for MB b-lobe development. Structure of ab neurons labeled with FasII antibody. All the images are z-projections of confocal image stacks (scale bar, 50 mm). (A) Morphologically aberrant ab neurons in dsh1 adult brains. Flies homozygous for a PCP specific allele of dsh (dsh1 ) show a phenotype comparable to Appl2/2 mutants but with increased penetrance of b-lobe loss as indicated by the arrow (30%; n = 36). (B) Morphologically normal ab neurons in dsh1 /Y;dsh-DshGFP-wt/+ adult brains. Reintroduction of wild-type Dsh in the dsh1 mutant background completely rescues b-lobe loss with no brains showing defects (n = 31). (C) Morphologically aberrant ab neurons in dsh1 /Y;;dshDshGFP-Y473F/+ adult brains. Reintroduction of a Tyrosine 473 phospho-mutant form of Dsh in a dsh1 mutant background fails to rescue the b-lobe loss as indicated by the arrow (n = 41). (D–E) Expression pattern of Dsh-GFP in ;dsh-DshGFP-wt/TM6b and of Dsh-Y473F-GFP in ;dsh-DshGFP-Y473F/+. Immuno-fluorescence analysis of adult brains using anti-GFP (green) and anti-FasII (D9, E9 in magenta) antibodies (scale bar, 50 mm). All images are zprojections of two confocal sections (0.9 mm steps). The D0 and E0 panels show the merge of the FasII and GFP channel for the indicated samples. doi:10.1371/journal.pbio.1001562.g004 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 7 May 2013 | Volume 11 | Issue 5 | e1001562 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 8 May 2013 | Volume 11 | Issue 5 | e1001562 Figure 5. APPL interacts with PCP signaling during MB development. (A–D) Structure of ab neurons labeled with the FasII antibody. All the images are z-projections of confocal image stacks (scale bar, 50 mm). (A) Morphologically normal ab neurons in a control adult brain. (B) Morphologically aberrant ab neurons in Appld w*/Y;;FzKD /+ adult brains. Loss of one copy of Fz in the Appl2/2 background increases moderately, but not significantly, the penetrance of b-lobe loss (indicated by the arrow) to 21% (n = 28); p value = 0.2166 calculated with G-test. (C) Morphologically aberrant a/b neurons in Appld w*/Y;UAS-Fz-DN/+;P247Gal4/+ adult brains. Overexpression of a dominant negative form of Fz in the Appl2/2 background increases moderately, but not significantly, the penetrance of the phenotype to 18% (n = 28); p value = 0.4226 calculated with G-test. (D) Morphologically aberrant ab neurons in Appld w*/Y;Vangstbm-6 /+ adult brains. Loss of one copy of Vang in the Appl2/2 background increases the penetrance of the b-lobe loss up to 33% (n = 21); p value = 0.02304 calculated with G-test. (E) The graph shows the penetrance of b-lobe loss in genetic interaction experiments, normalized against the penetrance in the Appld w* background. (F, G) APPL and Vang localization during development in brain of flies Act-Stbm-EYFP expressing an EYPF tagged form of Vang under the control of the Actin promoter. Immuno-fluorescence analysis using anti-APP-Cterm (F, G in blue), anti-GFP (F9, G9 in green), and anti-FasII (F0, G0 in red) antibodies. The images are single confocal stacks (scale bar, 25 mm). The F- and G- panels show the merge of the FasII and GFP channel for the indicated samples. (F) At 48 h APF, Vang is broadly expressed along the MB b axons, where also APPL is expressed as indicated by the arrow. (G) At adult stage, the level of Vang detectable in the b axons is strongly reduced compared to the previously analyzed time point, while APPL is enriched in the MB neurons (as indicated by the arrow). (H) The graph shows the penetrance of the b-lobe loss in the Appldw* flies overexpressing Dsh or Wnt5 (p value equal to 0.1287 and 0.0006692, respectively). Activation of Wnt-PCP signaling upon Wnt5 signaling rescues the b-lobe phenotype. ** Indicates a p value,0.001 calculated with G- test. doi:10.1371/journal.pbio.1001562.g005 Figure 6. APP is required for the proper response to Wnt5 and forms complexes with PCP core proteins. (A) Analysis of the responsiveness of wt MEFs and MEFs lacking all the APP isoforms to Wnt5 treatment. MEFs were treated for 2 h with rmWnt5a and subsequently analyzed by Western blot. After the Wnt5 treatment, Dvl2 is phosphorylated and this modification is indicated by a shift of the band detected by Dvl2 Ab. KO MEFs respond less efficiently to Wnt5. Re-introduction of hAPP rescues the responsiveness to Wnt5. The graph shows a quantification of the ratio between phospho-Dvl2 and Dvl2 in the analyzed samples. * Indicates a p value,0.05 calculated with one-way ANOVA plus Tukey’s multiple comparison test. (B) Co-immunoprecipitation (Co-IP) of Appl-FLAG and Vang-Myc. The tagged proteins were co-expressed in HEK293T cells and immunoprecipitated with anti-FLAG antibody. Vang-Myc can be precipitated upon IP of Appl-FLAG. (C) Co-IP of Appl-FLAG and human Vangl2-HA. Human Vangl2-HA can be precipitated upon IP of Appl-FLAG. (D) Co-IP of human APP (C99)-FLAG and human Vangl2-HA. Human Vangl2-HA can be precipitated upon IP of APP (C99)-FLAG, indicating that interaction with PCP proteins is a conserved feature of APP proteins. (E) Co-IP of Appl-FLAG and dFz-GFP. Drosophila Fz can be precipitated upon IP of Appl-FLAG. (F) Co-IP of human APP (C99)-FLAG and human V5-Fzd5. V5-Fzd5 can be precipitated upon IP of APP (C99)-FLAG. doi:10.1371/journal.pbio.1001562.g006 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 9 May 2013 | Volume 11 | Issue 5 | e1001562 still not clear what triggers the accumulation of the peptide and how this is correlated with the pathogenesis of the disease and the dementia. In fact, most of the work done to unveil the pathogenesis of the disease has focused on the analysis of Ab-peptide and the search for its receptors and downstream effectors. Even though the numerous in vitro studies performed in cell culture identified several molecules that interact with Ab peptide, the in vivo biological relevance of these interactions remains to be clarified. The amyloid cascade hypothesis has also dominated the search for AD treatments, but the promising molecular candidates developed to modulate the Ab peptide and reached clinical trials failed [34,35]. Finally, over the last few years many studies indicated that there is no linear correlation between the accumulation of the peptide and the cognitive decline, leading to a revision of the amyloidogenic hypothesis. Taken together, these observations suggest that the accumulation of the peptide is not the only cause of the pathology and that other factors are involved. Interestingly, under physiological conditions APP is mainly found in its uncleaved or a-cleaved form, suggesting that the shift towards amyloidogenic processing not only increases the production of Ab peptide but also depletes the pool of APP that undergoes non-amyloidogenic processing, with hitherto unknown consequences. It is therefore of paramount importance to understand the physiological role of APP and how perturbing this role could contribute to the pathogenesis of the disease. An important contribution to the study of the function of a protein comes from the analysis of the knock-out (KO) animals. In the case of APP, several KO models have been generated and analyzed in detail both from the morphological and behavioral point of view [13,36– 38]. Despite these efforts, the normal physiological function of APP in vivo in the nervous system remains largely elusive and highly controversial. This is due to the lack of consensus over the neuronal phenotypes in null mutant animals and the mechanism of action in vivo. The data collected by different labs confirmed the involvement of APPs in development and function of the nervous system, but these studies do not provide an in-depth analysis of the development of the brain during the pre-natal stages or the molecular mechanism underlying APPs’ putative functions. We therefore took advantage of Drosophila melanogaster to further analyze the consequence of loss of APP Like (APPL) during brain development. In the present study we demonstrate that APPL is involved in brain development of Drosophila melanogaster, particularly in the Mushroom Body (MB) neurons. We show that APPL is required for the development of ab neurons. In Appl2/2 flies, MB neurons fail to project the a lobe in 14% of the cases and the b-lobe in 12% of the cases (Figure 1). Further analysis of the phenotype reveals that APPL is required cell-autonomously for the development of the b lobe and non-cell autonomously for the development of the a lobe. In fact, single cell Appl2/2 clones display only b-lobe loss and no a loss. The re-introduction of a full-length, membrane-tethered form of APPL, but not a soluble form, rescues b-lobe loss (Figure 2). This is of particular interest because it confirms that, similar to mammalian APPs, the physiological role of APPL is mediated both by its full-length form, required in the neurons to achieve the correct b-lobe pattern, and by its soluble form (sAPPL) that regulates the extension of the a lobe. Moreover, the rescue data indicate that, at least in this context, the function of sAPPL is mediated not by homo-dimerization with the full-length form but by some other receptor, hitherto unknown. Further experiments are required to clarify the sAPPL non-cell autonomous effect, but we hypothesize that it might be involved in modulating signaling mediated by the cells that surround the MB axons. Taken together, the analysis of the Appl2/2 animals confirmed the important role of APPs during brain development but reinforced the idea that the phenotypes are present with incomplete penetrance and might be subtle. It would therefore be of interest to analyze the phenotype of the KO mice in greater detail and, in particular, to better characterize the APP’s downstream pathway leading to these defects. Moreover, the results described clearly support a model of APPL as a novel, neuronal-specific positive modulator of the WntPCP pathway (Figure 4). The PCP pathway was initially described because of its role in tissue polarity establishment and, in particular, of its regulation of cell orientation in plane of an epithelium. Among the different processes regulated by PCP signaling, we are interested in axon growth and guidance. It has been described that mice null for Fzd3/Ceslr32/2 genes show severe defects in several major axon tracts like thalamocortical, corticothalamic, and nigrostriatal tracts, defects of the anterior commissure, and similarly to APP KO mice, the variable loss of the corpus callosum [5,39]. The molecular mechanism underlying the function of PCPsignaling in regulating tissue polarity has been broadly studied. The current model suggests that, upon polarized expression of the different core proteins, Dsh is recruited to the membrane via Fz and leads to the activation of a cascade of small GTPases finally resulting in cytoskeleton rearrangements. In the case of regulation of axon growth and guidance, it is less clear how the signaling is regulated and transmitted to the cytoskeleton. A recent publication suggested that during axon growth the transmembrane PCP receptor-like Vang and Fzd are localized at the growth cone area on the tip of the fillopodia, thus suggesting that in this context the asymmetric localization is not needed [28]. Furthermore, Dsh needs to relocalize from the cytoplasm to the membrane to ensure the proper activation of PCP signaling, and this is dependent on its phosphorylation status. Singh and colleagues showed that Abelson is one kinase responsible for this modification, but the receptor upstream of the kinase was not identified [26]. Based on the evidence we generated, we propose that APPL is a novel regulator of Wnt-PCP pathway involved in axon growth and guidance (Figure 7). This is of interest because while the PCP core proteins are ubiquitously expressed, APPL is restricted to the nervous system, suggesting that it could be the first described tissue-specific modulator of the pathway. Mechanistically, we propose that APPL-Abl complex modulates Dsh via dual protein-protein interactions. First, Abl might have an intrinsic affinity for its substrate Dsh [26]. Secondly, this interaction is strengthened or stabilized by the inclusion of APPL in a PCP receptor complex. This dual affinity complex leads to increased PCP signaling efficiency at the developing growth cone. Both biochemical and physiological data show that this function is highly conserved in mammalian APP, suggesting that it may play a similar role in the mammalian brain. The canonical-Wnt signaling pathway has already been connected to AD pathogenesis because of its link to the tau-kinase GSK-3b. Interestingly, no clear link between the WntPCP pathway and this neurodegenerative disorder has been made. Previous reports [27,28] show that, in flies and mice, Jun N-terminal Kinase (JNK) is the final effector of PCP in axon outgrowth and JNK was shown to be required for the effect of APP overexpression in the fly [25,40]. Interestingly, JNK signaling has also been linked to the neuronal loss observed in AD [41]. It is therefore worth investigating whether the physiological function of APP as a neuronal PCP modulator explains the JNK-AD connection. Materials and Methods Fly Stocks Drosophila stocks used include Appld w* , Appld ,FRT19A w* , elavC155 ,hsFLP,w*;UAS-mCD8::GFP.,UAS-lacZ/CyO;tubP-GAL80, APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 10 May 2013 | Volume 11 | Issue 5 | e1001562 FRT2A/TM6,Tb,Hu, hsFlp,UAS-CD8-GFP;;FRT2A,tubGal80/ TM3;OK107, hsFlp, UAS-CD8-GFP; FRT2A, tubGal80/ TM3;OK107, Flp122; sp/CyO;Fz p21,ri,FRT2A/TM2, fz2C1 ri,FRT2A/TM3,Sb, yw,hsflip;Fz1H51 Fz2C1 riFRT2A /TM2, UAS-Fz-DN/CyO;P247Gal4/TM6c, UAS-Abl-KD/CyO;P247Gal. Abl4 kar1 red1 e1 /TM6B, Tb1 , w*;UAS-Abl/CyO;P247, w; fz [KD] /TM3,Sb, Vangstbm-6 , w;201Y,UAS-GFP, FRT19A;ry50 , FRT19A,tub-Gal80,hsFLP/FM7;UAS-CD8-GFP/CyO;OK107, UAS-Appl/CyO;P247Gal4, UAS-sAppl/Cyo;P247Gal4, w1 , dsh:1 , dsh.Dsh-GFP (J7)/TM6. dshV26,dsh.Dsh-GFP; dsh.Dsh-GFP Y473F, and Act-stbm-EYFP/TM3. Accession Numbers/ID Numbers APPL (FBgn0000108), fz (FBgn0001085), fz2 (FBgn0016797), Fzd5 (NP_003459.2), dsh (FBgn0000499), dvl1 (AAB65242.1), dvl2 (AAB65243.1), Vang (FBgn0015838), Vagl2 (NP_065068.1), and Abl (FBgn0000017). Immunochemistry Larval, pupal, or adult brains were dissected in phosphate buffered saline (PBS) and fixed in 3.7% formaldehyde in PBT (PBS+ Triton 6100 0.1%) for 15 min. The samples were subsequently rinsed three times in PBT and blocked in PAXDG for 1 h. Following these steps, the brains were incubated with the primary antibody diluted in PAX-DG overnight at 4uC. This incubation was followed by three washes with PBT and a subsequent incubation with the appropriate fluorescent secondary antibodies for 2 h at RT. After three rinses in PBT, the brains were put in 50% Glycerol diluted in PBS and then mounted in Vectashield (Vector Labs) mounting medium. The following antibodies were used: rabbit anti-GFP (Invitrogen, 1:1,000), mouse anti-FasII (Hybridoma Bank, 1:50), rabbit anti-APP-Cterm (kind gift of Bart de Strooper lab, 1:5,000), and antiPhalloidin TRITC (1:1,000). Microscopy and Image Analysis The mounted brains were imaged either on a LEICA DM 6000 CS microscope coupled to a LEICA CTR 6500 confocal system or on a Nikon A1-R confocal (Nikon) mounted on a Nikon Ti-2000 inverted microscope (Nikon) and equipped with 405, 488, 561, and 639 nm lasers from Melles Griot. The pictures were then processed using ImageJ and Adobe Photoshop. MARCM Procedure Crosses were set up at 25uC and transferred every day. We transferred 0 to 24 pupae in a fresh vial, and they were heath shocked for 459 at 37uC and shifted back at 25uC until eclosion. The morphology of the MB clones was analyzed in flies 0–7 d old. Cell Culture and Treatments WT MEFs, APP/APLP2 double KO MEFs, APP/APLP2 double KO+hAPP, and HEK-293T cells were propagated in DMEM, 10% FCS, 2 mM L-glutamine, 50 units/ml penicillin, 50 units/ml streptomycin. MEFs (200,000 cells per well) were seeded in 24-well plates for biochemical analyses. MEFs were treated 2 d after seeding with rmWnt5a (R&D Systems) for 2 h. Cells were harvested for immunoblotting by direct lysis in 16Laemmli buffer followed by boiling at 95uC for 5 min. Control stimulations were done with 0.1% BSA in PBS. Gel Electrophoresis and Western Blots Protein from total cell lysates/samples was resolved in 10% polyacrylamide gels (SDS-PAGE) under denaturing conditions and then transferred to nitrocellulose membranes. The blots were probed using polyclonal anti-FLAG M2 (F1804, Sigma-Aldrich 1:1,000), monoclonal anti-Myc (M4439,Sigma-Aldrich 1:1,000), anti-Dvl1 (sc-8025, Santa Cruz Biotechnologies, 1:1,000), antiDvl2 (#3224, Cell Signaling Technologies, 1:1,000), anti-V5 (R960-25, Invitrogen, 1:1,000), anti-HA (HA.11, MMS-101R, Covance, 1:2,000), and anti-beta-actin (sc-1615, Santa Cruz Biotechnology, 1:2,000). Bands were visualized using anti-IgG HRP-conjugated secondary antibodies, and the ECL Western Blotting Detection System (GE Healthcare, UK). Co-Immunoprecipitation For the Co-IP of Drosophila proteins, pCDNA3-APPL-FLAG and pCDNA3-Vang-Myc were transiently transfected in HEK293T cells (4.56106 cells per 10 cm dish) using Fugene HD (Roche). After 3 d, cells were collected in Lysis Buffer (150 mM NaCl, 50 mM Tris/HCl pH 7.5, 10% glycerol, 0.4% Nonidet P-40) and cleared with Dynabeads M-270 epoxy (Invitrogen) for 459 at 4uC. After the clearing, lysates (half volume) were incubated with anti-FLAG covalently conjugated to Dynabeads M-270 (pre-saturated with BSA) for 1 h at 4uC. Beads were then washed, and bound proteins were resuspended in 66 Laemmli and subjected to SDS-PAGE followed by Western blot analysis. For the Co-IP of Drosophila and human proteins, HEK293 cells grown at 50% confluency on 10-cm plates were transfected with 6 mg of each plasmid. After 2 d, cells were lysed for 15 min in 1 ml of lysis buffer ([0 mM Tris buffer pH 7.4, 150 mM NaCl, 1 mM EDTA, 0.5% NP40 supplemented with 1 mM DTT and protease inhibitor cocktail (Roche, cat. no. 11836145001)]. Lysates were centrifuged at 13,2006 g for 20 min at 4uC, supernatants were collected, and 0.4 ml of the supernatant was incubated with 1 mg of indicated immunoprecipitating antibody for 1 h at 4uC. Immunoprecipitates were collected on Protein G sepharose beads by overnight rotation, washed four times with lysis buffer, resuspended in 26Laemmli sample buffer, Figure 7. APPL is a novel modulator of Wnt-PCP signaling required for axon guidance. Schematic representation of the proposed model. APPL is a novel regulator of Wnt-PCP pathway. In the presence of Wnt signaling, Fz binds Dsh, which needs to be phosphorylated by Abelson kinase to correctly relocalize to the membrane. We propose that Appl is part of the membrane complex formed by the core PCP proteins Fz1 and Vang. In turn, Appl recruits Abelson kinase to the complex and positively modulates Dsh phosphorylation. The subsequent activation of the signaling is required for MB b-axon growth. doi:10.1371/journal.pbio.1001562.g007 APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 11 May 2013 | Volume 11 | Issue 5 | e1001562 and subjected to SDS-PAGE followed by Western blot analysis. The antibodies used for immunoprecipitation include FLAG M2 (F1804 Sigma-Aldrich), anti-HA (ab9110; Abcam), anti-Myc (M4439, Sigma-Aldrich), and anti-V5 (R960-25, Invitrogen). Supporting Information Figure S1 Appl is required during the development of Mushroom Bodies for a- and b-lobe growth. (A) The table shows the number of brains analyzed to characterize the Appl loss of function phenotype. (PDF) Figure S2 APPL is required cell-autonomously for the b-lobe outgrowth. (A)The table lists the number of brains analyzed in the MARCM experiments. (B, C) The table lists the number of brains analyzed in the rescue experiments. (D) Adult MB lobes labeled with FascilinII II antibody (FasII). The image is a z-projection of confocal image stacks (scale bar, 50 mm). Morphologically normal ab neurons in Appld w*/Y;UAS-sAppl/+;P247Gal4 adult brains. The reintroduction of soluble APPL in MBs during development strongly reduces the loss of the a lobe. (PDF) Figure S3 Abelson kinase is the downstream effector of APPL in MB development. (A) Structure of a/b neurons of a Appld;;Abl4 48APF brain. The image is a z-projection of confocal image stacks (scale bar, 50 mm). The MBs are labeled with anti-FasII antibody. b-lobe loss is detectable already at 48 APF, similarly to what is observed in the Appl2/2 background. (PDF) Figure S4 Dsh phosphorylation is required for MB development. (A) The table lists the number of brains analyzed in the dsh rescue experiments. (PDF) Figure S5 Appl interacts with the PCP signaling during MB development. (A) The table lists the number of brains analyzed in the PCP genetic interaction experiments. (B) Structure of a/b neurons labeled with anti-FasII antibody. The image is a z-projection of confocal image stacks (scale bar, 50 mm). Morphologically normal a/ b neurons in UAS-FzDN/+;P247Gal4/+ adult brains. Expression of a dominant negative form of fz in the MB is not sufficient to induce defects. (C–E) Z-projections of confocal image stacks of GFP-labeled MARCM clones. Immuno-fluorescence analysis of adult MB, using anti-GFP (green) and anti-FasII (magenta) antibodies. (C) fzp21 mutant clones obtained by crossing elavC155,hsFLP,w*;UAS- mCD8::GFP,UAS-lacZ/CyO;tubP-GAL80,FRT2A/TM6,Tb,Hu or hsFlp, UAS-CD8-GFP;; FRT2A, tubGal80/TM3; OK107 with yw Flp122; sp/CyO;fz p21,ri,FRT2A/TM2. fz mutant cells show normal axon projections, similar to their wild-type counterparts. (D) fz2C1 mutant clones obtained by crossing virgins elavC155,hsFLP,w*;UAS-mCD8::GFP.,UAS-lacZ./CyO; tubP-GAL80,FRT2A/TM6,Tb,Hu or hsFlp, UAS-CD8-GFP; FRT2A, tubGal80/TM3; OK107 with males fz2C1 ri, FRT2A/ TM3, Sb. Single-cell clones do not show any difference in their projection pattern compared to wild-type cells. (E) fzH51, fz2C1 double mutant clones obtained by crossing elavC155,hsFLP, w*;UAS-mCD8::GFP.,UAS-lacZ./CyO;tubP-GAL80,FRT2A/ TM6,Tb,Hu with yw,hsflip; ;fz H51fz2 C1ri FRT2A/TM2. Loss of both fz and fz2 does not influence b-lobe growth, thus excluding possible compensatory effects. (F) Appl expression in third instar larvae eye disc. (G–I) Tangential adult eye sections in areas around the equator. The colored bars indicate the orientation of the ommatidia. (H) dsh1 mutant flies show PCP defects and reduction of symmetric ommatidia. (I) Appl2/2 adult flies show ommatidia orientation comparable to wild-type flies (G). (J) APPL and Vang localization during development in brain of flies expressing a EYPF tagged form of Vang under the control of Actin promoter. Immunofluorescence analysis using anti-APP-Cterm (blue), anti-GFP (green), and anti-FasII (red) antibodies. The images are single confocal stacks (scale bar, 15 mm). APPL and Vang are expressed in mutually exclusive compartments in the developing retina. (PDF) Figure S6 APP proteins are found in core PCP complexes. (A) APP expression levels in wild-type MEFs, KO MEFs, or KO MEFs stably transfected with APP. Two clones of rescue MEFs were analyzed. Clone B shows detectable APP levels and was used for the Wnt5 stimulation assay. (B) Co-immunoprecipitation (CoIP) of Appl-FLAG and Vang-Myc. The tagged proteins were coexpressed in HEK293T cells and immunoprecipitated with antiMyc antibody. Appl-FLAG can be precipitated upon IP of Vang. The Co-IP in this direction is weaker than after pull-down of ApplFlag. (C) Co-IP of Appl-FLAG and human Vangl2-HA. ApplFLAG can be precipitated upon IP of human Vangl2-HA. (D) CoIP of APP (C99)-FLAG and Vangl2-HA. The tagged proteins were immunoprecipitated from whole cells with anti-HA antibody. APP (C99)-FLAG can be precipitated upon IP of human Vangl2HA. (E) Control co-immunoprecipitation of overexpressed ApplFLAG and Vangl2-HA. The proteins were separately expressed in different cells plated in two different dishes and pooled during the Co-IP procedure. The IP was performed with anti-FLAG and anti-HA antibody and followed by Western blot analysis. The proteins do not co-immunoprecipitate when expressed in different populations of cells. (F) Co-immunoprecipitation (Co-IP) of ApplFLAG and Fz-GFP. Appl-FLAG can be precipitated upon IP of Fz. (PDF) Acknowledgments We thank Gary Struhl for providing Flp122; sp/CyO;Fz p21,ri,FRT2A/TM2 stock, David Strutt for providing the Act-stbm-EYFP/TM3, Marek Mlodzik for sharing with us the ;dsh.Dsh-GFP (J7)/TM6 and the dshV26,dsh.DshGFP; dsh.Dsh-GFP Y473F stocks, Ulrike Mu¨ller for providing APP/APLP2 KO fibroblasts, Patrick Callaerts for sharing the hsFlp,UAS-CD8GFP;;FRT2A,tubGal80/TM3;OK107 stock, and the lab of Stein Aerts for the support on the statistical analysis of the data. We thank Ariane Ramaekers and Luca Tiberi for critical discussion. Author Contributions The author(s) have made the following declarations about their contributions: Conceived and designed the experiments: AS J-MD VB BAH. Performed the experiments: AS ZO PJ KT ER AC JY JMD. Analyzed the data: AS ZKA J-MD VB BAH. Contributed reagents/ materials/analysis tools: BDS. Wrote the paper: AS BAH. References 1. Bayly R, Axelrod JD Pointing in the right direction: new developments in the field of planar cell polarity. Nat Rev Genet 12: 385–391. 2. Seifert JR, Mlodzik M (2007) Frizzled/PCP signalling: a conserved mechanism regulating cell polarity and directed motility. Nat Rev Genet 8: 126–138. 3. Lyuksyutova AI, Lu CC, Milanesio N, King LA, Guo N, et al. (2003) Anteriorposterior guidance of commissural axons by Wnt-frizzled signaling. Science 302: 1984–1988. 4. Ng J (2012) Wnt/PCP proteins regulate stereotyped axon branch extension in Drosophila. Development 139: 165–177. APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 12 May 2013 | Volume 11 | Issue 5 | e1001562 5. Tissir F, Bar I, Jossin Y, De Backer O, Goffinet AM (2005) Protocadherin Celsr3 is crucial in axonal tract development. Nat Neurosci 8: 451–457. 6. Jacobsen KT, Iverfeldt K (2009) Amyloid precursor protein and its homologues: a family of proteolysis-dependent receptors. Cell Mol Life Sci 66: 2299–2318. 7. Luo LQ, Martin-Morris LE, White K (1990) Identification, secretion, and neural expression of APPL, a Drosophila protein similar to human amyloid protein precursor. J Neurosci 10: 3849–3861. 8. Luo L, Tully T, White K (1992) Human amyloid precursor protein ameliorates behavioral deficit of flies deleted for Appl gene. Neuron 9: 595–605. 9. Bayer TA, Wirths O (2008) Review on the APP/PS1KI mouse model: intraneuronal Abeta accumulation triggers axonopathy, neuron loss and working memory impairment. Genes Brain Behav 7 Suppl 1: 6–11. 10. Iijima K, Liu HP, Chiang AS, Hearn SA, Konsolaki M, et al. (2004) Dissecting the pathological effects of human Abeta40 and Abeta42 in Drosophila: a potential model for Alzheimer’s disease. Proc Natl Acad Sci U S A 101: 6623– 6628. 11. Ling D, Song HJ, Garza D, Neufeld TP, Salvaterra PM (2009) Abeta42-induced neurodegeneration via an age-dependent autophagic-lysosomal injury in Drosophila. PLoS One 4: e4201. 12. Pereira C, Ferreiro E, Cardoso SM, de Oliveira CR (2004) Cell degeneration induced by amyloid-beta peptides: implications for Alzheimer’s disease. J Mol Neurosci 23: 97–104. 13. Muller U, Cristina N, Li ZW, Wolfer DP, Lipp HP, et al. (1994) Behavioral and anatomical deficits in mice homozygous for a modified beta-amyloid precursor protein gene. Cell 79: 755–765. 14. Herms J, Anliker B, Heber S, Ring S, Fuhrmann M, et al. (2004) Cortical dysplasia resembling human type 2 lissencephaly in mice lacking all three APP family members. EMBO J 23: 4106–4115. 15. Young-Pearse TL, Suth S, Luth ES, Sawa A, Selkoe DJ Biochemical and functional interaction of disrupted-in-schizophrenia 1 and amyloid precursor protein regulates neuronal migration during mammalian cortical development. J Neurosci 30: 10431–10440. 16. Nikolaev A, McLaughlin T, O’Leary DD, Tessier-Lavigne M (2009) APP binds DR6 to trigger axon pruning and neuron death via distinct caspases. Nature 457: 981–989. 17. Kamal A, Almenar-Queralt A, LeBlanc JF, Roberts EA, Goldstein LS (2001) Kinesin-mediated axonal transport of a membrane compartment containing beta-secretase and presenilin-1 requires APP. Nature 414: 643–648. 18. Lazarov O, Morfini GA, Lee EB, Farah MH, Szodorai A, et al. (2005) Axonal transport, amyloid precursor protein, kinesin-1, and the processing apparatus: revisited. J Neurosci 25: 2386–2395. 19. Russo C, Venezia V, Repetto E, Nizzari M, Violani E, et al. (2005) The amyloid precursor protein and its network of interacting proteins: physiological and pathological implications. Brain Res Brain Res Rev 48: 257–264. 20. Ring S, Weyer SW, Kilian SB, Waldron E, Pietrzik CU, et al. (2007) The secreted beta-amyloid precursor protein ectodomain APPs alpha is sufficient to rescue the anatomical, behavioral, and electrophysiological abnormalities of APP-deficient mice. J Neurosci 27: 7817–7826. 21. Neve RL, McPhie DL, Chen Y (2000) Alzheimer’s disease: a dysfunction of the amyloid precursor protein(1). Brain Res 886: 54–66. 22. Selkoe DJ, Yamazaki T, Citron M, Podlisny MB, Koo EH, et al. (1996) The role of APP processing and trafficking pathways in the formation of amyloid betaprotein. Ann N Y Acad Sci 777: 57–64. 23. Goguel V, Belair AL, Ayaz D, Lampin-Saint-Amaux A, Scaplehorn N, et al. (2011) Drosophila amyloid precursor protein-like is required for long-term memory. J Neurosci 31: 1032–1037. 24. Lee T, Luo L (1999) Mosaic analysis with a repressible cell marker for studies of gene function in neuronal morphogenesis. Neuron 22: 451–461. 25. Leyssen M, Ayaz D, Hebert SS, Reeve S, De Strooper B, et al. (2005) Amyloid precursor protein promotes post-developmental neurite arborization in the Drosophila brain. EMBO J 24: 2944–2955. 26. Singh J, Yanfeng WA, Grumolato L, Aaronson SA, Mlodzik M (2010) Abelson family kinases regulate Frizzled planar cell polarity signaling via Dsh phosphorylation. Genes Dev 24: 2157–2168. 27. Srahna M, Leyssen M, Choi CM, Fradkin LG, Noordermeer JN, et al. (2006) A signaling network for patterning of neuronal connectivity in the Drosophila brain. PLoS Biol 4: e348. 28. Shafer B, Onishi K, Lo C, Colakoglu G, Zou Y (2011) Vangl2 promotes Wnt/ planar cell polarity-like signaling by antagonizing Dvl1-mediated feedback inhibition in growth cone guidance. Dev Cell 20: 177–191. 29. Zhou L, Bar I, Achouri Y, Campbell K, De Backer O, et al. (2008) Early forebrain wiring: genetic dissection using conditional Celsr3 mutant mice. Science 320: 946–949. 30. Shimizu K, Sato M, Tabata T (2011) The Wnt5/planar cell polarity pathway regulates axonal development of the Drosophila mushroom body neuron. J Neurosci 31: 4944–4954. 31. Grillenzoni N, Flandre A, Lasbleiz C, Dura JM (2007) Respective roles of the DRL receptor and its ligand WNT5 in Drosophila mushroom body development. Development 134: 3089–3097. 32. Axelrod JD, Miller JR, Shulman JM, Moon RT, Perrimon N (1998) Differential recruitment of Dishevelled provides signaling specificity in the planar cell polarity and Wingless signaling pathways. Genes Dev 12: 2610–2622. 33. Tanzi RE, Bertram L (2005) Twenty years of the Alzheimer’s disease amyloid hypothesis: a genetic perspective. Cell 120: 545–555. 34. Hardy J, Selkoe DJ (2002) The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeutics. Science 297: 353–356. 35. Karran E, Mercken M, De Strooper B (2011) The amyloid cascade hypothesis for Alzheimer’s disease: an appraisal for the development of therapeutics. Nat Rev Drug Discov 10: 698–712. 36. Li ZW, Stark G, Gotz J, Rulicke T, Gschwind M, et al. (1996) Generation of mice with a 200-kb amyloid precursor protein gene deletion by Cre recombinase-mediated site-specific recombination in embryonic stem cells. Proc Natl Acad Sci U S A 93: 6158–6162. 37. Zheng H, Jiang M, Trumbauer ME, Sirinathsinghji DJ, Hopkins R, et al. (1995) beta-Amyloid precursor protein-deficient mice show reactive gliosis and decreased locomotor activity. Cell 81: 525–531. 38. Heber S, Herms J, Gajic V, Hainfellner J, Aguzzi A, et al. (2000) Mice with combined gene knock-outs reveal essential and partially redundant functions of amyloid precursor protein family members. J Neurosci 20: 7951–7963. 39. Wang Y, Thekdi N, Smallwood PM, Macke JP, Nathans J (2002) Frizzled-3 is required for the development of major fiber tracts in the rostral CNS. J Neurosci 22: 8563–8573. 40. Taru H, Iijima K, Hase M, Kirino Y, Yagi Y, et al. (2002) Interaction of Alzheimer’s beta -amyloid precursor family proteins with scaffold proteins of the JNK signaling cascade. J Biol Chem 277: 20070–20078. 41. Kihiko ME, Tucker HM, Rydel RE, Estus S (1999) c-Jun contributes to amyloid beta-induced neuronal apoptosis but is not necessary for amyloid beta-induced cjun induction. J Neurochem 73: 2609–2612. APPL Regulates Wnt-Dependent Axon Growth PLOS Biology | www.plosbiology.org 13 May 2013 | Volume 11 | Issue 5 | e1001562 A Genotype n α loss β loss Appl-/- 101 14% 12% Appl-/+ 97 0 0 BA Genotype n α loss β loss Appl -/- 101 14% 12% Appl +/- 97 0 0 Control clones 41 0 0 Appl -/- clones 44 0 10% Genotype n β loss Appl -/- 101 12% Appl -/+ 97 0 Appl -/-, rescue fl -APPL 45 2% Appl -/+, rescue fl -APPL 35 0 Appl -/-, rescue sAPPL 50 12% Appl -/+, rescue sAPPL 23 0 Appl -/-, rescue APPL ΔC 54 11% Appl -/+, rescue APPLΔC 47 0 Appl -/-, rescue control (driver) 47 13% Genotype n α loss β loss Appl -/- 101 14% 12% Appl -/+ 97 0 0 Appl -/- , rescue sAPPL 50 4% 12% Appl -/+, rescue sAPPL 23 0 0 Appl -/- 47 13% 13%rescue control (driver), FasII C D 48APF Appl -/-, Abl -/+ FasII A Genotype n β loss Dsh1 36 30% Dsh1 , Dsh>Dsh-GFP 31 0 Dsh1 , Dsh>Dsh -Y473F-GFP 41 29% fz-DN FasII B fz mutant single cell clones C C’ C’’ FasIIGfp Merge MergeFasIIGfp fz2 mutant single cell clones fz/fz2 mutant single cell clones MergeFasIIGfp D D’ D’’ E E’ E’’ A Genotype n β loss Appl -/- 101 12% Appl +/- 97 0 dsh1 36 30% Appl -/-, fz -/+ 28 21% Appl -/+, fz -/+ 21 0 Appl -/-, fzDN 28 18% fzDN 19 0 Appl -/-, Vang +/ - 21 33% fz clones 40 0 wild type dsh1 Appl -/-F G H I Vang-GFP PhalloidinAPPL Merge J J’ J’’ J’’’ B C D HA-Vangl2 WB: HA + + + +-App (C99)-Flag WB: Flag + + + +- IPHA INPUT HA-Vangl 2 dAppl -Flag + + + +- - + + + +- WB: HA WB: Flag IP HA INPUT + - + + - + - + + - + + Appl -flag Vang-myc WB: flag WB: myc IP myc INPUT HA-Vangl2 + + + +- -dAppl-Flag + + + +- - + + + +- WB: HA WB: Flag INPUTIP HAIP FLAG E - + - + - + - + Wnt 5a 100 ng/ml MEFwt dKO +APP A +APP BA WB: flag WB: GFP - + + - + + + - + + - + APPL - flag dFz - GFP IP GFP INPUT F       Vítězslav Bryja, 2014    Attachments      #16      Kriz  V1 ,  Pospichalova  V1 ,  Masek  J,  Kilander  MB,  Slavik  J,  Tanneberger  K,  Schulte  G,  Machala M, Kozubik A, Behrens J, Bryja V. (2014): β‐arrestin promotes Wnt‐induced  Lrp6 phosphorylation via increased membrane recruitment of Amer1. J Biol Chem. 289  (2): 1128 –1141.  1  equal contribution      Impact factor (2012): 4.651  Times cited (without autocitations, WoS, Feb 21st 2014): 0  Significance: This work explains how β‐arrestin, a Dvl‐binding protein, regulates Wnt/β‐ catenin  signaling.  This  question  has  been  puzzling  for  a  long  time  because  in  contrast to non‐canonical Wnt pathway, where β‐arrestin regulates endocytosis,  the  mechanism  of  action  in  canonical  Wnt  pathway  was  unknown.  We  demonstrate that β‐arrestin via its interaction with the scaffolding protein Amer1  mediates phophatidyl‐inositol phosphate‐controlled contact between Fzd/Dvl and  Lrp6/Axin membrane complexes.  Contibution of the author/author´s team: Full design, coordination and experimental  parts  of  the  project  with  the  exception  of  lipid  analysis  and  some  FRAP  validations.                ␤-Arrestin Promotes Wnt-induced Low Density Lipoprotein Receptor-related Protein 6 (Lrp6) Phosphorylation via Increased Membrane Recruitment of Amer1 Protein*□S Received for publication,July 1, 2013, and in revised form, November 5, 2013 Published, JBC Papers in Press,November 21, 2013, DOI 10.1074/jbc.M113.498444 Víteˇzslav Krˇízˇ‡§1 , Vendula Pospíchalová‡1,2 , Jan Masˇek‡¶ , Michaela Brita Christina Kilanderʈ , Josef Slavík**, Kristina Tanneberger‡‡ , Gunnar Schulte‡ʈ3 , Miroslav Machala**, Alois Kozubík‡§ , Juergen Behrens‡‡4 , and Víteˇzslav Bryja‡§5 From the ‡ Faculty of Science, Institute of Experimental Biology, Masaryk University, 611 37 Brno, Czech Republic, the § Department of Cytokinetics, Institute of Biophysics, Academy of Science of the Czech Republic, 612 65 Brno, Czech Republic, the ‡‡ Nikolaus-Fiebiger-Center, University of Erlangen-Nürnberg, 91054 Erlangen, Germany, the ʈ Department of Physiology and Pharmacology, Karolinska Institutet, 171 77 Stockholm, Sweden, the **Department of Toxicology, Pharmacology, and Immunotherapy, Veterinary Research Institute, 621 00 Brno, Czech Republic, and the ¶ Institute of Molecular Genetics, Academy of Science of the Czech Republic, 142 20 Prague, Czech Republic Background: ␤-Arrestins are required for Wnt/␤-catenin signaling, but the mechanism of their action is unclear. Results: ␤-Arrestins bind PtdInsP2-interacting protein Amer1, bind PtdInsP2-producing kinases, and control Wnt3a-induced PtdInsP2 production and Amer1 membrane dynamics. Conclusion: ␤-Arrestins bridge Wnt3a-induced Dvl-associated PtdInsP2 production to the phosphorylation of Lrp6 via control of Amer1 dynamics. Significance: The first mechanistic explanation how ␤-arrestin regulates Wnt/␤-catenin signaling is provided. ␤-Arrestin is a scaffold protein that regulates signal transduction by seven transmembrane-spanning receptors. Among other functions it is also critically required for Wnt/␤-catenin signal transduction. In the present study we provide for the first time a mechanistic basis for the ␤-arrestin function in Wnt/␤catenin signaling. We demonstrate that ␤-arrestin is required for efficient Wnt3a-induced Lrp6 phosphorylation, a key event in downstream signaling. ␤-Arrestin regulates Lrp6 phosphorylation via a novel interaction with phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2)-binding protein Amer1/WTX/ Fam123b. Amer1 has been shown very recently to bridge Wntinduced and Dishevelled-associated PtdIns(4,5)P2 production to the phosphorylation of Lrp6. Using fluorescence recovery after photobleaching we show here that ␤-arrestin is required for the Wnt3a-induced Amer1 membrane dynamics and downstream signaling. Finally, we show that ␤-arrestin interacts with PtdIns kinases PI4KII␣ and PIP5KI␤. Importantly, cells lacking ␤-arrestin showed higher steady-state levels of the relevant PtdInsP and were unable to increase levels of these PtdInsP in response to Wnt3a. In summary, our data show that ␤-arrestins regulate Wnt3a-induced Lrp6 phosphorylation by the regulation of the membrane dynamics of Amer1. We propose that ␤-arrestins via their scaffolding function facilitate Amer1 interaction with PtdIns(4,5)P2, which is produced locally upon Wnt3a stimulation by ␤-arrestin- and Dishevelled-associated kinases. Wnt/␤-catenin signaling plays a key role in homeostasis and embryonic development. Deregulation of Wnt/␤-catenin pathway has been shown to cause pathophysiological conditions such as developmental abnormalities, tumorigenesis, or osteoporosis (1, 2). The Wnt/␤-catenin cascade is activated by Wnts, which act as agonists for the class Frizzled (Fzd)6 receptors (3). When extracellular Wnt is present, it links its receptor Fzd and its co-receptor low density lipoprotein receptor-related protein 5 or 6 (Lrp5/6). Fzds bind Dishevelled (Dvl), which is required for the Wnt-induced phosphorylation of the intracellular domain (ICD) of Lrp5/6 (4, 5). Dvl and two Dvl-associated kinases, phosphatidylinositol 4-kinase type II␣ (PI4KII␣) and * This work was supported by Czech Science Foundation Grants 204/09/ 0498,204/09/J030,andGA13-32990S;aEuropeanMolecularBiologyOrganization (EMBO) installation grant; and Ministry of Education Youth and Sport of the Czech Republic Grant MSM0021622430. The collaboration between Masaryk University and Karolinska Institutet is supported by Project KI-MU Grant CZ.1.07/2.3.00/20.0180. Author’s Choice—Final version full access. □S This article contains supplemental Figs. 1 and 2. 1 Both authors contributed equally to this work. 2 Supported by a project co-financed by the European Social Fund and the state budget of Czech Republic “Postdoc I,” Grant CZ.1.07/2.3.00/30.0009. 3 Supported by Knut and Alice Wallenberg Foundation Grant KAW2008.0149, Swedish Research Council Grants K2008-68P-20810-01-4 and K2008-68K- 20805-01-4, and The Swedish Foundation for International Cooperation in Research and Higher Education (STINT). 4 Supported by Deutsche Forschungsgemeinschaft Grant Be1550/6-1. 5 To whom correspondence should be addressed: Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlárˇská 2, 611 37 Brno, Czech Republic. Tel.: 420-549-49-3291; Fax: 420-541-21-1214; E-mail: bryja@sci.muni.cz. 6 The abbreviations used are: Fzd, Frizzled; Apc, adenomatous polyposis coli; ␤-arr, ␤-arrestin; CK1, casein kinase 1; CM, conditioned medium; DKO, double knock-out; Dvl, Dishevelled; EGFP, enhanced green fluorescent protein; FRAP, fluorescence recovery after photobleaching; GSK3␤, glycogen synthase kinase 3␤; ICD, intracellular domain; Lrp, low density lipoprotein receptor-related protein; MEF, mouse embryonic fibroblast; PI4KII␣, phosphatidylinositol 4-kinase type II␣; PIP5KI␤, phosphatidylinositol-4-phosphate 5-kinase type I␤; PtdIns, phosphatidylinositol; PtdIns(4,5)P2, phosphatidylinositol 4,5-bisphosphate; TCL, total cell lysate. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 289, NO. 2, pp. 1128–1141, January 10, 2014 Author’s Choice © 2014 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A. 1128 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom phosphatidylinositol-4-phosphate 5-kinase type I␤ (PIP5KI␤), which produce phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) in two sequential steps from phosphatidylinositol (PtdIns), were found to be crucial for Lrp5/6 phosphorylation (6). PtdIns(4,5)P2 is recognized by Amer1/WTX, which links PtdIns(4,5)P2 production with the machinery phosphorylating Lrp5/6 (7). Phosphorylated Lrp5/6 subsequently recruits Axin1 (8), which is a key component of the ␤-catenin destruction complex formed by the scaffolding proteins Axin1 and Apc (adenomatous polyposis coli) and kinases GSK3␤ (glycogen synthase kinase 3␤) and CK1␣ (casein kinase 1␣). As a consequence, phosphorylation and subsequent proteosomal degradation of ␤-catenin are blocked, ␤-catenin is stabilized in the cytoplasm, and following translocation into the nucleus it drives transcription of Wnt-responsive genes in cooperation with the Tcf/Lef transcription factors (9–11). Previous studies by us and others have demonstrated the key role of ␤-arrestin (␤-arr) scaffolding protein in Wnt signaling (for review, see Ref. 12). ␤-Arrestins were shown to be required for Wnt/␤-catenin signaling (13, 14) as well as for branches of Wnt signaling, which do not activate ␤-catenin (collectively referred to as noncanonical Wnt pathways) (15–18). Whereas in noncanonical Wnt signaling ␤-arrestins regulate signal propagation via their well defined role in the clathrin-mediated receptor endocytosis (17, 18), the mechanism how ␤-arrestins control Wnt/␤-catenin signaling, which does not depend on the clathrin-mediated endocytosis (19), is unclear. In the present study we aimed to elucidate the role of ␤-arrestins in the process of the Wnt/␤-catenin pathway activation. We demonstrate that ␤-arrestin is a novel binding partner of Amer1/WTX/Fam123b (further referred to as Amer1). We show that ␤-arrestin binds PtdIns(4,5)P2-producing kinases PI4KII␣/PIP5KI␤ and that it is required for PtdIns(4,5)P2-controlled membrane dynamics of Amer1 upon Wnt3a stimulation. This function of ␤-arrestin governs Wnt-induced Lrp6 phosphorylation. We propose that ␤-arrestins acts as a scaffold, which brings Amer1 physically close to the site of Wnt3a-induced PtdIns(4,5)P2 production. EXPERIMENTAL PROCEDURES Cell Culture and Transfection—HEK293T cells were cultured in DMEM supplemented with 10% FBS, 2 mM L-glutamine, 50 units/ml penicillin and 50 units/ml streptomycin. Cells were transfected with polyethyleneimine (PEI) as described earlier (20). Briefly, working stocks of 25-kDa PEI were prepared from commercially available PEI (40.872-7; Sigma-Aldrich) by dilution 1:500 in PBS, pH was adjusted to 7.0, and sterile stocks of PEI were stored for months in 4 °C. DNA and stock PEI (1 ␮g:2.5 ␮l) were mixed in serum- and antibiotic-free DMEM and incubated for 30 min. The transfection mixture was then added dropwise to cells. Media were changed 3 h after transfection. Cells were harvested 24–48 h after transfection. For immunofluorescence experiments cells were transfected using calcium phosphate method or using Lipofectamine 2000 (Invitrogen). siRNA transfection was performed on 24-well plates. Each transfection reaction contained 1 ␮l of Lipofectamine RNAiMAX (Invitrogen), 30 ␮M siRNA, and 48 ␮l of serum-free medium. Following the incubation for 20–30 min at room temperature the mixture was added to the suspension of trypsinized cells (0.5 ml/well). The media were changed 6 h after transfection. Cells were harvested 48 h after transfection. Both wild type and ␤-arrestin1/2 DKO MEFs (21) were cultured in DMEM supplemented with 10% FBS, 2 mM L-glutamine, 50 units/ml penicillin, and 50 units/ml streptomycin. The NB4 cell line was cultured in RPMI 1640 medium supplemented with 10% heat-inactivated FBS, 2 mM L-glutamine, 50 units/ml penicillin, and 50 units/ml streptomycin. Plasmids, Antibodies, and siRNA—Plasmids encoding FLAGAmer1, EGFP-Amer1, EGFP-Amer1 deletion mutants, EGFPLrp6-ICD-Amer1 (7, 22), HA-␤-arrestin2, FLAG-␤-arrestin2, FLAG-␤-arrestin2 deletion mutants (14), FLAG-hDvl3 (23), FLAG-hDvll1 (24), MYC-mDvl2 (25), EGFP-mDvl2 (16), HA-PI4KII␣ and HA-PIP5KI␤ (6), SuperTOPFLASH and SuperFOPFLASH Tcf/Lef (26) reporters were described earlier. Renilla luciferase (pRL-TK) was purchased from Promega. pEGFP-Amer1(2–838)-Lrp6-ICD was cloned by inserting Lrp6-ICD (cut from the pEGFP-Amer1-Lrp6-ICD by NotI) into a NotI site located between EGFP and Amer1(2–838) in the pEGFP-Amer1(2–832) construct. Lef1-VP16 was kindly provided by V. Korˇínek (IMG AS CR, Prague) and myristoylmCherry by Jyrki Kukkonen. The following antibodies were used: mouse anti-FLAG antibody (F1804; Sigma-Aldrich), rat anti-GFP antibody (3H9; Chromotek), mouse HA.11 (MMS-101R; Covance), rabbit anti-HA (ab9110; Abcam), anti-Dvl3 (sc-8027; Santa Cruz Biotechnology), anti-p(S1490)-Lrp6(2568;CellSignaling),anti-␤-catenin(610153; BD Biosciences), anti-␤-actin (sc-1615; Santa Cruz Biotechnology), anti-Myc antibody (sc-40; Santa Cruz Biotechnology), anti␣-catenin (sc-7894; Santa Cruz Biotechnology) mouse antiAmer1 (27), rabbit anti-Amer1 (AP17838PU-N; Tocris), rabbit anti-␤-arrestin (3857; Cell Signaling), anti-␤-arrestin (A1CT and A2CT), a kind gift from R. J. Lefkowitz), rabbit anti-PI4KII (a kind gift from P. De Camilli), and rabbit anti-IgG control antibody (3900; Cell Signaling). siRNA sequences targeting ␤-arrestin1/2 (15) and PI4KII (6) were described earlier. Fluorescence Recovery after Photobleaching (FRAP)—Thirtyfive-mm glass-bottom dishes (MatTek) were precoated with 0.1% collagen. MEFs were transfected in suspension with 1.6 ␮g of EGFP-Amer1 or EGFP-Amer1(2–838) plasmid using DreamFect Gold (OZ Bioscience) according to the protocol recommended by the supplier and plated on the dish. FRAP analysis was carried out 24 h after transfection as described earlier (7) on a Zeiss LSM710 scanning microscope. Immunoprecipitation and Western Blotting—Immunoprecipitation of overexpressed proteins in HEK293T cells was performed at 4 °C as described previously (28). Briefly, confluent 10-cm dishes were lysed and scraped 24 h after transfection in 1 ml of Nonidet P-40 lysis buffer (50 mM Tris, pH 7.4, 150 mM NaCl, 1 mM EDTA, 0.5% Nonidet P-40) supplied with 0.1 mM DTT and 1ϫ Complete protease inhibitor mix (11836145001; Roche Applied Science). Lysates were spun down (16,000 ϫ g, 10 min, 4 °C). Similarly, protein extracts were prepared from T75 flasks of NB4 cells for immunoprecipitation of endogenous proteins. The total protein concentration was determined using DCTM Protein Assay (500-0112; Bio-Rad). Sixty ␮g were used ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1129 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom as total cell lysate (TCL), 1 mg of the lysate was mixed with 1 ␮g of the immunoprecipitating antibody (anti-FLAG (F1804; Sigma), anti-GFP (20R-GR011; Fitzgerald), anti-HA (ab9110; Abcam), rabbit anti-␤-arrestin (3857; Cell Signaling), anti-␤arrestin (A1CT and A2CT), or IgG control antibody (3900; Cell Signaling)) and kept rotating on a carousel at 4 °C. Fifteen ␮l of solid protein G-Sepharose beads (17-0618-05; GE Healthcare) were added to each reaction 30 min after the addition of antibodies. Test tubes were placed on the carousel overnight at 4 °C. The next day, beads were collected by spinning down at 100 ϫ g for 1 min at 4 °C and washed six times with the lysis buffer. Immunoprecipitation and TCL samples were mixed with denaturing reducing Laemmli buffer, boiled, and if necessary also sonicated before loading on the SDS-PAGE. The proteins were separated according their molecular mass on 8–15% SDS-PAGE and transferred to an Immobilon-P membrane (Millipore). When required Western blots were quantified by densitometry analysis using ImageJ software. Immunofluorescence Microscopy—Cell were transfected a day after plating on 0.1% collagen-precoated coverslips. Twenty four h after the transfection, cells were washed with PBS and fixed with 4% formaldehyde in PBS. Cells were washed in PBS and blocked in PBTA (1ϫ PBS, 5% BSA, 0.25% Triton X-100, 0.01% NaN3) and incubated with the primary antibody overnight (4 °C). Cells were then washed three times with PBST (1ϫ PBS, 0.1% Triton X-100) and incubated for 1 h with the secondary antibody conjugated with Alexa Fluor dye 594 or 647 (Invitrogen). Cells were washed three times with PBST, incubated with DAPI (1 ␮g/ml in PBST) for 10 min, and washed once in PBST. PBST was replaced with PBS, and glass coverslips were mounted with glycerol gelatin mix (079K6006; Sigma) and stored in the dark at 4 °C until scanning. Microscopy was performed on a Zeiss LSM710 laser scanning microscope or sp5 confocal microscope (Leica). Quantification of co-localization is shown as a graph of the overlap of fluorescence intensity peaks of individual channels along profiles indicated in the merged micrographs (Zen software - Zeiss in Fig. 2 and LAS AF software (Leica) in Fig. 6). Dual Luciferase Assay—Dual luciferase assay was carried out in HEK293T cells. Cells were transfected on 24-well plates. Transfected cells were analyzed 24 (plasmid DNA) or 48–72 (siRNA) h after transfection according to a slightly modified protocol recommended by the supplier (E1960; Promega). Briefly, culture medium was removed, and cells were washed with PBS. Each well was lysed for 15 min at room temperature in 50 ␮l of lysis buffer. To measure firefly luciferase activity 20 ␮l from each well was pipetted into a microtiter plate and mixed with 25 ␮l of luciferase substrate. Luminescence was immediately measured with a Microtiter Plate Luminometer. The signal was normalized to Renilla luciferase activity, which was measured after addition of 25 ␮l of Stop and Glo mix. Graphs represent averages Ϯ S.E. of fold changes in relative luciferase units (ratio luciferase/Renilla) over the control condition. Membrane Fractionation—Membrane fractions were prepared from MEFs using a ProteoJET Membrane Protein Extraction kit (K0321; Fermentas) down-scaled for 6-well plate format. Briefly, cells were stimulated with Wnt3a (50 ng/ml), or alternatively the cells were treated only with an equal volume of 0.1% BSA in PBS used for dilution of Wnt3a. After 1 h of treatment, medium was removed and replaced with 1 ml of Cell Wash Solution. Each well was treated with 700 ␮l of ice-cold permeabilization buffer supplemented with 1 mM DTT, 10 mM NaF, and 1ϫ Complete protease inhibitor mixture. Plates were slowly shaken on ice for 10 min. Cells were scraped and spun down at 16,000 ϫ g, 15 min, 4 °C. Supernatant was used for membrane fractionation. Pellet was mixed with 60 ␮l of Membrane Protein Extraction Buffer and vortexed at 1200 rpm, 30 min, 4 °C. Tube content was spun down at 16,000 ϫ g, 15 min, 4 °C. The membrane fraction was present in the supernatant. Membrane and cytoplasmic fractions of NB4 cells were prepared using Subcellular Protein Fractionation kit for Cultured Cells (78840; Thermo Scientific) according to the manufacturer’s recommendations. In short, NB4 cells were stimulated with Wnt3a conditioned (or control) media produced in L-cells for 2 h, spun down at 200 ϫ g, 5 min, 4 °C, washed in ice-cold PBS and Wash cells by suspending the cell pellet with ice-cold PBS. One ml of ice-cold cytoplasm extraction buffer containing Complete protease inhibitor mixture was added to the cell pellet, and the tube was incubated at 4 °C for 10 min with gentle mixing. The supernatant cytoplasmic extract was gathered after centrifugation at 500 ϫ g for 5 min. Next, ice-cold membrane extraction buffer containing Complete protease inhibitor mixture was added to the pellet. The tube was vortexed for 5 s on the highest setting and incubated at 4 °C for 10 min with gentle mixing. The supernatant (membrane extract) was collected by centrifugation at 3000 ϫ g for 5 min. The total amount of protein was determined in each fraction by DC Protein Assay, and the extracts were subjected to immunoprecipitation and/or analyzed using Western blotting. Phospholipid Species Analysis by Tandem Mass Spectrometry (MS)—Wild type and ␤-arrestin1/2 DKO MEFs seeded onto 150-mm plates were treated with Wnt3a conditioned (or control) media produced in L-cells and harvested into 3-ml icecold methanol 2 h later. Lipid extraction was performed by modified Bligh/Dyer extraction protocol. Briefly, the cell pellet was transferred by 3.0 ml of methanol into glass tubes (20 ϫ 110 mm). The tubes were previously strongly sulfuric acid degreased and ethanol washed. After adding of 1.5 ml of chloroform and 50 ␮l of HCl (30%), the content was dispersed using probe sonicator at room temperature. The extraction process was continued by a 1-h incubation in an ice water bath. The above mentioned monophase extraction system was broken by the addition of 1.5 ml of chloroform and 2.5 ml of HCl (0.1 M). The samples were vigorously shaken for 1 min and subsequently centrifuged at 5000 ϫ g for 60 min until two phases become clearly visible. The upper phase was carefully removed, and the lower organic phase was dried by nitrogen. Solvent mixture A (chloroform/methanol/water/glacial acetic acid/ ammonium acetate solution 60:30:1:1:1 v/v) in a volume of 300 ␮l was used for reconstitution and the next mass spectrometric analysis using liquid chromatograph (Agilent 1200, Santa Clara, CA)-coupled mass spectrometer TripleQuand 6410 with electrospray ionization (Agilent). Tandem MS conditions were as follows: (i) direct infusion to mobile phase; (ii) sample volume 100 ␮l, mobile phase flow rate 20 ␮l/min, mobile phase:solvent mixture A; (iii) electrospray ␤-Arrestin Controls Lrp6 Phosphorylation 1130 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom ionization in positive mode, drying nitrogen flow 6 liters/min, fragmentor voltage and collision energy voltage optimized for each lipid group, neutral loss scan modes (m/z 277 for PtdIns lipid species, m/z 357 for PtdInsP lipid species, m/z 437 for PtdInsP2 lipid species). The amounts of selected PtdInsP lipid species, based on corresponding signal peak area, were normalized to the total PtdInsP amount in the sample. RESULTS Amer1 Is a Novel Interaction Partner of ␤-Arrestin2—It has been shown earlier that the scaffold protein ␤-arrestin is a necessary component of the Wnt/␤-catenin signaling pathway and a direct interactor of Dvl (13, 14). However, it is not clear how ␤-arrestin regulates Wnt signal transduction. To clarify this issue we have searched for high affinity physical interactions between ␤-arrestin and the conserved components of the Wnt signaling pathway. We have taken an advantage of the currently published study, which employed tandem affinity purification and mass spectrometry to map binding partners of 20 key components of the Wnt signaling pathway (29). Among the tested proteins only Wilms tumorassociated protein Amer1 co-precipitated with endogenous ␤-arrestin2. We have decided to confirm this interaction with the co-immunoprecipitation assay. As we show in Fig. 1A, overexpressed Amer1 and ␤-arrestin2 efficiently co-precipitated with each other. Moreover, Amer1 and ␤-arrestin2 interaction was detected also on endogenous level in the NB4 cell line of acute promyelocytic leukemia origin (Fig. 1B). This cell line was chosen using the Oncomine database as a suitable candidate having high expression levels of both Amer1 and ␤-arrestin2. First, we checked that this cell line is Wnt3a-responsive and found that Wnt3a conditioned medium (CM) stimulates Lrp6 phosphorylation at Ser-1490, a hallmark of Wnt/␤-catenin signaling (Fig. 1B, left panel). Furthermore, we detected that upon Wnt3a stimulation Amer1 gets recruited to the cytoplasmic membrane and/or associates more tightly with the cytoplasmic membrane (Fig. 1B, left panel) of NB4 cells. This protein dynamics is similar to what has been described in other cell lines (7). Endogenous immunoprecipitation in NB4 cells showed that Amer1 co-immunoprecipitated with two of three anti ␤-arrestin antibodies (Fig. 1B, right panel). To define the regions of individual proteins responsible for the interaction we have performed co-immunoprecipitation assays with the EGFP-tagged deletion mutants of Amer1 and FLAG-tagged deletion mutants of ␤-arrestin2. These experiments revealed that it is the very C-terminal part of Amer1 (amino acids 942–1135) that is required for the interaction with ␤-arrestin (Fig. 1C and supplemental Fig. 1, A and B). On the side of ␤-arrestin, we can demonstrate that full-length Amer1 binds the central pocket of ␤-arrestin2, which is defined by amino acids 163–300 (Fig. 1D and supplemental Fig. 1C). Finally, the interaction of these two short regions of Amer1 and ␤-arrestin2 can be demonstrated by an efficient co-immunoprecipitation of Amer1 (amino acids 942–1135) and ␤-arrestin (amino acids 163–300) (Fig. 1E). Amer1/␤-Arrestin2 Co-localization in the Plasma Membrane Is Disrupted by Dvl—Our findings show that Amer1 and ␤-arrestin2 efficiently interact with each other. In the next step we asked where in the cell this interaction takes place. We overexpressed FLAG-tagged Amer1 and EGFP-tagged ␤-arrestin2 in HEK293T cells and analyzed their subcellular localization by immunofluorescence. This analysis showed that Amer1 and ␤-arrestin2 co-localize in the proximity of the cell membrane (Fig. 2A). Importantly, it has been shown earlier that Dvl, another key signaling molecule in the Wnt pathway, binds, similarly to Amer1, the central part of ␤-arrestin2 (14). This observation opened the possibility that binding of Amer1 and Dvl to ␤-arrestin2 can be mutually exclusive, and ␤-arrestin2 is either present in complex with Amer1 or with Dvl. To test this possibility, we overexpressed Amer1, Dvl2, and ␤-arrestin2 and analyzed the effect of Dvl2 on the co-localization of Amer1 and ␤-arrestin2. As we show in Fig. 2B (in contrast to Fig. 2A), Dvl2, which is localized in the typical dots composed of Dvl2 multimers (30, 31), efficiently prevented membrane localization of ␤-arrestin2 and recruited ␤-arrestin2 to Dvl2 dots. However, the membrane localization of Amer1 was not affected. To support this observation biochemically, we have analyzed the mutual binding of Amer1 and ␤-arrestin2 in the presence of Dvl1, Dvl2, and Dvl3 by co-immunoprecipitation. As shown in Fig. 2C and supplemental Fig. 2, inclusion of Dvl3 completely and inclusion of Dvl1 and Dvl2 partially abolished the binding of Amer1 to ␤-arrestin2. These observations suggest that binding of Amer1 and Dvl to the central region of ␤-arrestin2 is competitive and that Dvl has the capacity to interfere with the interaction of Amer1 to ␤-arrestin2. ␤-Arrestin Regulates Lrp6 Phosphorylation and the Activity of Lrp6-ICD-Amer1 Fusion Protein—It has been shown recently that Amer1 has a positive role in the Wnt3a-induced phosphorylation of Lrp6. Amer1 links Dvl-associated PtdIns(4,5)P2 production to the formation of Axin-GSK3␤-CK1␥ complexes required for the phosphorylation and activation of the Lrp6 receptor (7). Our next goal was to clarify whether or not ␤-arrestin has a role in this Amer1-mediated process. Importantly, the analysis of ␤-arr1/2 DKO MEFs demonstrates that endogenous ␤-arrestin is required for the efficient Lrp6 phosphorylation at Ser-1490 triggered by Wnt3a (Fig. 3A). These observations demonstrate that ␤-arrestin is required for the Lrp6 phosphorylation. In the next step we therefore tested whether or not ␤-arrestins regulate Lrp6 phosphorylation via regulation of the Amer1 function. It is known that membrane targeting of Lrp6 ICD is absolutely crucial to trigger the downstream signaling by Lrp6-ICD. The cytoplasmic ICD of Lrp6 alone shows no activity (Fig. 3B, first bar) in the Tcf/Lef-dependent transcription luciferase reporter assay (TOPFLASH, (32)), which serves as a convenient readout of the activity of Wnt/␤-catenin signaling in cultured cells. (In contrast when Lrp6-ICD is membrane-targeted due to the presence of the transmembrane domain (Lrp6-ICD/TM) it becomes constitutively active (Fig. 3B, second bar). Amer1 can mediate membrane targeting via its PtdIns(4,5)P2 binding domains, which are responsible for both Amer1 membrane recruitment and Amer1 activity in the Wnt/␤-catenin pathway (7). Not surprisingly, the fusion protein of Lrp6-ICD and ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1131 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom Amer1 (Lrp6-ICD-Amer1) (7) acted as a strong activator of TOPFLASH (but not its negative control counterpart FOPFLASH) (Fig. 3B, third bar). The activity of Lrp6-ICD-Amer1 was dependent on PtdIns(4,5)P2 because PtdIns(4,5)P2 depletion by ionomycin, which induced release of intracellular Ca2ϩ stores and subsequent cleavage of PtdIns(4,5)P2 by Ca2ϩ activated phospholipase C, (i) resulted in the translocation of EGFP-Amer1, but not an unrelated mCherry protein bound to membrane via a myristoyl anchor, from the membrane to the cytoplasm (Fig. 3C) and (ii) at the same time dose-dependently reduced the TOPFLASH reporter activity triggered by the Lrp6-ICD-Amer1 fusion protein (Fig. 3D), whereas the TOPFLASH activity induced by a constitutively active form of Lef1 (Lef1-VP16) remained unchained (Fig. 3E). Together, these experiments suggest that the activity of Lrp6ICD-Amer1 is based on PtdIns(4,5)P2-dependent membrane recruitment of Amer1. This allowed us to use the Lrp6-ICDAmer1 fusion as the readout to analyze the role of ␤-arrestin in PtdIns(4,5)P2-mediated Amer1 membrane recruitment in the context of the Wnt/␤-catenin signaling. Using this system we can FIGURE 1. Amer1 interacts with ␤-arrestin2. A, HEK293T cells were transfected with FLAG-Amer1 and HA-␤-arrestin2 individually or together as indicated. Cell lysates were immunoprecipitated (IP) with anti-FLAG or anti-HA antibody. Immunoblotting (WB) was done with mouse M2 anti-FLAG antibody and with mouse anti-HA antibody. TCL was used as a loading control. B, NB4 cells were stimulated for 2 h with Wnt3a (or control) CM, and cytoplasmic and membrane fractions were harvested. Left panel, NB4 cells are Wnt3a-responsive as shown by the enhanced Lrp6 phosphorylation at Ser-1490. Both Amer1 and ␤-arrestin2 are isolated preferentially in the cytoplasmic fraction. However, upon Wnt3a stimulation, Amer1 is recruited to the membrane and/or associates more tightly with the membrane and is therefore isolated also in the membrane fraction of NB4 cells. Right panel, Amer1 co-immunoprecipitated with A1CT and A2CT anti ␤-arrestin antibodies, but not with anti-IgG control antibody and anti-␤-arrestin2 cs-3857 antibody, which poorly precipitates endogenous ␤-arrestin2. C, domain mapping of the Amer1/␤-arrestin2 interaction using full-length ␤-arrestin2 and deletion mutants of Amer1 shows schematic view of Amer1 mutants. The interaction with ␤-arrestin2 is indicated with ϩ; M1/M2 indicate regions interacting with membrane via PtdIns(4,5)P2; A1/A2/A3 are regions required for the interaction with Apc. Full blots are shown in supplemental Fig. 1, A and B. D, domain mapping of the Amer1/␤-arrestin2 interaction uses full-length Amer1 and deletion mutants of ␤-arrestin2. The mutual interaction is indicated with ϩ. Full blots are shown in supplemental Fig. 1C. E, the interaction between the C-terminal region of Amer1 (amino acids 942–1135) and the central part of ␤-arrestin2 is demonstrated by co-immunoprecipitation. ␤-Arrestin Controls Lrp6 Phosphorylation 1132 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom demonstrate that expression of ␤-arrestin2 promoted (Fig. 4A), whereas knockdown of ␤-arrestin1/2 decreased (Fig. 4B) TOPFLASH activation by Lrp6-ICD-Amer1. These observations are compatible with the possibility that ␤-arrestin modulates the PtdIns(4,5)P2-dependent membrane recruitment of Amer1. The positive effect of ␤-arrestin2 on the Lrp6-ICD-Amer1-induced TOPFLASH activity is not due to interference with the downstream signaling events because ␤-arrestin2 overexpression failed to further promote TOPFLASH activation by the constitutively active Lrp6-ICD/TM (Fig. 4C). ␤-Arrestin Is Required for Wnt3a-induced Amer1 Membrane Dynamics—We have shown previously that Wnt3a stimulation leads to the increase in the immobilized, likely PtdIns(4,5)P2bound, Amer1 fraction in the plasma membrane (7). We have FIGURE 2. Co-localization of Amer1 and ␤-arrestin2 in the proximity of cell membrane is disrupted by Dvl2. A and B, top, confocal microscopy images of HEK293Tcellsco-transfectedwiththeindicatedplasmids,fixed,andstainedwiththerelevantantibodies.Bottom,overlapoffluorescenceintensitypeaksalong profiles indicated in the merged micrographs by a white line. White arrowheads indicate cell membrane position. A, FLAG-Amer1 (red) and EGFP-␤-arrestin2 (green) co-localizing (yellow) at the plasma membrane. B, MYC-Dvl2 (gray) abolishing the co-localization between Amer1 (red)/␤-arrestin2 (green) and recruiting ␤-arrestin2 to Dvl dots. C, HEK293T cells transfected with HA-␤-arrestin2, FLAG-Amer1, and EGFP-Dvl2. The interaction of individual proteins was analyzed using co-immunoprecipitation. The efficient interaction between ␤-arrestin2 and Amer1 is partially abolished by Dvl2 protein. ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1133 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom applied FRAP to test whether Amer1 membrane stabilization requires ␤-arrestin. Wild type (WT) and ␤-arr1/2 DKO MEFs were transfected with Amer1-EGFP-tagged protein and subjected to FRAP essentially as described previously (7). As we show in Fig. 5, Amer1-EGFP is predominantly membranous in both WT (Fig. 5A) and ␤-arr1/2 DKO (Fig. 5B) fibroblasts. FIGURE 3. ␤-Arrestin2 mediates Lrp6 phosphorylation by the regulation of Amer1 membrane recruitment. A, wild type (WT) or ␤-arrestin-deficient (␤-arr1/2 DKO)MEFswerestimulatedwithrecombinantmouseWnt3a(ordiluentcontrol).TheanalysisofphosphorylatedLrp6(atSer-1490)inthemembranefractionsshows that the activation of Lrp6 is attenuated in ␤-arr1/2 DKO MEFs. ␣-Catenin was used as the loading control. B, schematic represents various Lrp6 ICD-containing constructs and their activity in the Tcf/Lef reporter assay (TOPFLASH activity). TM, transmembrane domain; LRP6-ICD-Amer1, fusion of Amer1 and Lrp6 ICD). The Tcf/Lef-responsive(orWnt-responsive)constructpTOPFLASHencodesfireflyluciferasereportergeneunderthecontrolofaminimalpromoterandmultiplecopiesof the optimal Tcf-binding motif CCTTTGATC. The negative control (Wnt-nonresponsive) construct pFOPFLASH harbors multiple copies of the mutant motif CCTTTGGCC. Each transfection is accompanied with constitutively expressing Renilla luciferase construct, which serves as an internal control for normalizing transfection efficienciesandmonitoringcellviability.GraphsrepresentaveragesϮS.E.(errorbars)of-foldchangesinrelativeluciferaseunits(luciferasecounts/Renillacounts)over thecontrolcondition.C,timelapseimagesshowHEK293TcellstransfectedwithEGFP-Amer1ormyristolatedmCherryandtreatedwithionomycin.Top,EGFP-Amer1 is predominantly membranous (t ϭ Ϫ20 s) but within 40 s after application of ionomycin (added at t ϭ 0 s), which depletes intracellular PIP2 via PKC-dependent pathway, it translocates to the cytoplasm. Bottom, ionomycin does not alter distribution of a mCherry protein bound to membrane via its myristoyl anchor. In both experiments, GFP and mCherry photobleaching was software-controlled. D and E, HEK293T cells were transfected with the indicated plasmids and treated with ionomycin.Tcf/Lef-dependenttranscriptionalactivitywasdeterminedbyTOPFLASHreportersystem.ThereporteractivityofLrp6-ICD-Amer1isdependentonPIP2 as demonstrated by treatment with increasing doses of ionomycin (D), which depletes PIP2 via a PKC-dependent pathway. However, ionomycin does not affect Lef1VP16 (constitutively active form of Lef1)-driven TOPFLASH reporter expression (E). ␤-Arrestin Controls Lrp6 Phosphorylation 1134 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom When we performed FRAP (exampled by two small squares in membrane regions and one control region outside the cell in Fig. 5, A and B) we have not observed any difference between WT and DKO fibroblast in the mobile pool of Amer1 (71.9% in WT and 72.1% in ␤-arr1/2 DKO MEFs) in the unstimulated conditions (Fig. 5, A and B, right panel). However, following Wnt3a stimulation the mobile pool of Amer1 in WT fibroblasts decreased to 58.2%, which is similar to the results, which we observed earlier in HEK293T cells (7). Interestingly, after Wnt3a stimulation ␤-arr1/2 DKO MEFs completely failed to FIGURE 4. ␤-Arrestin2 promotes Tcf/Lef transcriptional activity of Lrp6-ICD-Amer1 fusion protein. A–C, top, HEK293T cells were transfected with the indicated combinations of plasmids and siRNA. Tcf/Lef-dependent transcriptional activity was determined by TOPFLASH/FOPFLASH reporter system. Bottom, Western blots (WB) of expression levels of the indicated proteins or effective knockdown or overexpression of ␤-arrestin are shown in each condition. Lrp6 fusion proteins are N-terminally EGFP-tagged, and ␤-arrestin2 is FLAG-tagged. A, ␤-arrestin2 promotes the TOPFLASH activity of Lrp6-ICD-Amer1. B, depletion of ␤-arrestin1/2 by siRNA reduces the transcriptional activity of Lrp6-ICD-Amer1. C, ␤-arrestin2 does not affect TOPFLASH activity of the constitutively active (ca) Lrp6 (Lrp6-ICD/TM). pcDNA3 was used as a control plasmid. Data are from at least five independent replicates. ***, p Ͻ 0.001 (one-way analysis of variance, Tukey’s post test). Error bars, S.E. ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1135 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom regulate the mobile pool of Amer1, which only slightly increased to 75% (Fig. 5B, right panel). Moreover, the Amer1(2–838) deletion variant, which lacks the C terminus of Amer1 responsible for the interaction with ␤-arrestin, showed higher mobility demonstrated by a shorter recovery halftime (approximately 5 s compared with 10 s in the full-length Amer1) and a higher mobile fraction (approximately 80%) (Fig. 5C). Most importantly, Wnt3a stimulation was unable to FIGURE 5. Analysis of Amer1 membrane dynamics in the presence and absence of ␤-arrestin in MEF cells reveals requirement of the interaction between Amer1 and ␤-arrestin for Wnt3a-induced Amer1 dynamics. A–C, WT and ␤-arr1/2 DKO MEFs were transfected with EGFP-Amer1 (A and B) or EGFP-Amer1(2–838) (C), and membrane dynamics of the constructs were analyzed by the FRAP method. Examples of cells subjected to FRAP are given on the left, squares indicate bleached regions. On the right, statistical analysis of FRAP experiment using WT and ␤-arr1/2 DKO MEFs stimulated with PBS or Wnt3a is shown. The graphs show mean values Ϯ S.E. (error bars), and the best fitting curve model, which was used for calculation of mobile pool of EGFP-Amer1 (percentage of fluorescence recovered) and of the recovery halftime (t1/2). N, number of analyzed cells. Wnt3a is unable to regulate Amer1 dynamics in the absence of ␤-arrestin and in the case of Amer1(2–838) construct, which lacks the ␤-arrestin interaction domain. D, levels of PtdIns-P in WT and DKO MEFs stimulated with Wnt3a CM were measured by mass spectrometry. The level of selected PtdInsP was increased by Wnt3a and by the absence of ␤-arr. Note the lack of Wnt3a-induced dynamics in ␤-arr1/2 DKO MEFs. The individual lipid species from the PtdInsP pool (Total) are shown separately. n ϭ 3. ***, p Ͻ 0.05 (one-way analysis of variance, Tukey’s post test). ␤-Arrestin Controls Lrp6 Phosphorylation 1136 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom change the membrane dynamics of Amer1(2–838) (Fig. 5C, blue line), which suggests that Wnt3a-triggered Amer1 membrane recruitment and/or Amer1 mobility within the membrane requires physical interaction with ␤-arrestin. To further analyze the role of ␤-arrestin in the PtdIns phosphate metabolism, we directly measured the levels of PtdIns, PtdInsP, and PtdInsP2 in WT and ␤-arr1/2 DKO MEFs stimulated with control or Wnt-3a CM. Mass spectrometry-based analysis (for details, see “Experimental Procedures”) identified numerous PtdIns-based lipid species. Some of these species (containing the following combination of fatty acids: palmitic/ oleic, stearic/linoleic, stearic/oleic, stearic/arachidonic, and stearic/h-linoleic) were detected repeatedly as PtdIns, PtdInsP, and PtdInsP2, which suggests that they are subject of phosphorylation by PI and PIP kinases. Interestingly, in these lipid species we have observed (i) Wnt3a-induced increase of PtdInsP and (ii) increase in the steady-state levels of PtdInsP in ␤-arr1/2 DKO MEFs (Fig. 5D). These data suggest that the defects in the dynamic Wnt3a-induced phosphorylation of PtdIns are compensated by an increase in the level of PtdInsP. Of note, lack of Wnt3a-induced dynamics accompanied by an increased steady-state amount, which is reminiscent with this phenotype, has been observed in ␤-arr1/2 DKO MEFs for the levels of TOPFLASH activity and of phosphorylated Dvl (14). In summary, the analysis of PtdInsPs is compatible with the possibility that the ␤-arr1/2 DKO MEFs are unable to respond dynamically to Wnt-3a treatment by formation of PtdInsPs and compensate for this deficit by increased steady-state levels of relevant PtdInsPs. PI4KII␣ and PIP5KI␤ Kinases Bind ␤-Arr2 and Promote Amer1⅐␤-Arr2 Complex Formation—It has been shown that the local PtdIns(4,5)P2 synthesis triggered by Wnt3a is mediated by sequential action of two Dvl-associated kinases, PI4KII␣ and PIP5KI␤ (6, 33). Moreover, a closely related PIP5KI␣ was found to interact with ␤-arrestin2 (34). We thus hypothesized that ␤-arrestins, which serve as scaffold proteins, may facilitate the signal transduction by recruiting Amer1 close to the source of PtdIns(4,5)P2 synthesis by their interaction with PtdIns(4,5)P2FIGURE 6. PI4KII␣ and PIP5KI␤ kinases bind to ␤-arrestin2 and co-localize with ␤-arrestin2 and Amer1 in the membrane. A and B, HEK293T cells were transfected with FLAG-␤-arrestin2 and HA-PI4KII␣ or HA-PIP5KI␤ individually or in combination. Cell lysates were immunoprecipitated (IP) with anti-FLAG antibody and anti-HA antibodies, and the composition of immunoprecipitates was analyzed by Western blotting. C, endogenous Amer1 and PI4KII␣ were co-immunoprecipitated with anti ␤-arrestin antibodies in the cytoplasmic fraction of NB4 cells. D and E, HEK293T cells were transfected with HA-PI4KII␣ (D) or HA-PIP5KI␤ (E), and EGFP-␤-arrestin2 and the subcellular localization of individual proteins were analyzed by immunocytofluorescence (left panel). EGFP-␤arrestin2 and endogenous Amer1 (stained with rabbit anti-Amer1 antibody) co-localize with PI4KII␣/PIP5KI␤ in the membrane compartment. Right, panels show the overlap of fluorescence intensity peaks of individual channels along profiles indicated in the merged micrographs by a white line. Black arrowheads indicate cell membrane position. ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1137 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom producing kinases. As we demonstrate in Fig. 6, A and B, both PI4KII␣ and PIP5KI␤ strongly interacted with ␤-arrestin2 physically in the co-immunoprecipitation assay. On the endogenous level, we were able to co-immunoprecipitate not only Amer1 but also PI4kII␣ (Fig. 6C) with anti-␤-arrestin antibodies in the cytoplasmic fraction of NB4 cells. Furthermore, both kinases also co-localized with ␤-arrestin2 and endogenous Amer1 with the most prominent co-localization observed at the membrane (Fig. 6, D and E). Interestingly, when we blocked PtdIns(4,5)P2 production by PI4KII␣ knockdown the interaction between Amer1 and ␤-arrestin2 significantly decreased (Fig. 7A). On the contrary, overexpression of either PI4KII␣ or PIP5KI␤, which increases the levels of PtdIns(4,5)P2 (35), strongly promoted the interaction between Amer1 and ␤-arrestin (Fig. 7B). Both PI4KII␣ and PIP5KI␤ were found in the ␤-arrestin2 pulldown together with Amer1. In contrast, overexpression of Dvl3 almost completely inhibited the interaction of ␤-arrestin2 with PI4KII␣ and PIP5KI␤ (Fig. 7C), which is very reminiscent of the Dvl effect on the ␤-arrestin2⅐Amer1 complex (see Fig. 2). Of note, this analysis also demonstrated that ␤-arrestin2 has higher affinity for PI4KII␣ (Fig. 7C, compare third and fourth lanes), whereas Dvl3 binds efficiently mainly to PIP5KI␤ (Fig. 7B, compare fifth and sixth lanes). Taken together, these protein-protein interaction data suggest (i) that ␤-arrestin and Amer1 can physically interact with the PtdIns(4,5)P2-producing kinases PI4KII␣ and PIP5KI␤, (ii) that PtdIns(4,5)P2 production triggers and is required for the efficient interaction of Amer1 and ␤-arrestin, and (iii) that the complex of Amer1⅐␤-arr2/PI-kinase can be disrupted by Dvl3. DISCUSSION In the present study we show for the first time that ␤-arrestins regulate Wnt3a-induced Lrp6 phosphorylation by the regulation of membrane recruitment and of the dynamics of Amer1. We propose that ␤-arrestin2 functionally links Fzdassociated (Fzd-Dvl-PI4KII-PIP5KI) and Lrp5/6-associated (Amer1-Axin-GSK3␤-CK1␥) complexes, which are both required for the efficient downstream signaling induced by Wnt3a. Phosphorylation of Lrp5/6 represents a key event required for downstream signaling leading to the stabilization of FIGURE 7. PI4KII␣ and PIP5KI␤ kinases control the interaction between ␤-arrestin2 and Amer1. A–C, HEK293T cells were transfected with the indicated plasmids and siRNAs. Interaction of individual proteins with ␤-arrestin2 was analyzed by anti-tag immunoprecipitation. Depletion of PI4KII␣ by siRNA prevents (A) whereas overexpression of PI4KII␣ promotes (B) efficient interaction of ␤-arrestin2 and Amer1. Numbers above blots represent the quantification of the signal by densitometry relative to the respective control condition which was set as 1. C, the interaction of ␤-arrestin2 with PI4KII␣/PIP5KI␤ kinases is completely abolished by the overexpression of Dvl3. ␤-Arrestin Controls Lrp6 Phosphorylation 1138 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom ␤-catenin and subsequent Tcf/Lef-driven transcription (8). The current model of Lrp5/6 phosphorylation pinpointed the key role of PtdIns(4,5)P2 as the required signal mediators, which transduce signal between Fzd/Dvl and Lrp6. Two Dvlassociated kinases, PI4KII␣ and PIP5KI␤, which together in two sequential steps produce PtdIns(4,5)P2, were found to be crucial for Lrp5/6 phosphorylation (6). We have shown recently that the scaffold protein Amer1 (27), also known as WTX (36), which was first described as the negative regulator of the Wnt signaling (36), facilitates Wnt3a-induced Lrp6 phosphorylation (7). Amer1 is a PtdIns(4,5)P2-binding protein, which associates with PtdIns(4,5)P2 produced locally in the membrane. As a consequence of PtdIns(4,5)P2 production by Dvl-associated kinases Amer1 recruits Lrp6 kinases in the proximity of ICD of Lrp6 (7). Phosphorylation of Lrp6 receptors then takes place in the five-times itinerated PPP(S/T)PXS motives in the Lrp6 ICD. Several kinases including GSK3␤, CK1␥, MAPKs, cyclin-dependent kinase, or G protein-coupled receptor kinases phosFIGURE 8. Model: role of ␤-arrestin in Lrp6 phosphorylation. A, the data presented in Figs. 1–7 support a model, where ␤-arrestin acts as a scaffold, which brings Amer1 close to the site of PIP2 production. A Wnt ligand activates pathway via Frizzled (FZD)/Dvl, which subsequently leads to the activation of PI4KII␣/PIP5KI␤ kinases. We propose that the activation of Dvl and the initial production of PIP2 allow translocation of ␤-arrestin and PI4KII␣/PIP5KI␤ toward Amer1-based Lrp6 phosphorylation complex composed of Amer1, Axin, and Lrp6-phosphorylating kinases CK1 and GSK3␤. Local production of PIP2 then stabilizes the Lrp6 phosphorylation complex and feeds the phosphorylation process. B, the FoldIndex prediction shows that Amer1 is largely an intrinsically disordered protein (in red) where the domains required for binding of key proteins/metabolites required for Lrp6 phosphorylation do not overlap. ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1139 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom phorylating Lrp6 have been identified (37–42); however, the relative contribution of individual kinases is still a matter of debate. Based on our data we propose a model of ␤-arrestin and Amer1 function in the Wnt3a-induced phosphorylation of Lrp6 (schematized in Fig. 8A). The model is based on our finding that ␤-arrestin can be either present in complex with Dvl or with Amer1. The interaction of Amer1 and ␤-arrestin takes place near the membrane and most importantly requires PtdIns-P2 and possibly other membrane lipids or membrane itself. The interaction of ␤-arrestin with Dvl seems to have higher affinity than the interaction of ␤-arrestin with Amer1 or PI4KII/PIP5KI kinases. As a consequence, overexpression of Dvl is able to efficiently disrupt the binding of ␤-arrestin with Amer1 and PI4KII/PIP5KI (see Figs. 2 and 7). It is known that the Wnt signaling cascade is induced by binding of Wnt to Fzd. This interaction subsequently triggers by an yet unidentified mechanism involving Dvl the activation of PtdIns(4,5)P2-producing kinases PI4KII and PIP5KI (6). The production of PtdIns(4,5)P2 is required and further promotes the interaction of Amer1 and ␤-arrestin. We speculate that following the activation, Dvl undergoes a conformational change (induced either by posttranslational modifications or by recruitment of other proteins), which breaks the ␤-arrestin/Dvl interaction and allows the formation of the ␤-arrestin⅐Amer1 and ␤-arrestin⅐PI4KII⅐PIP5KI complexes. ␤-Arrestin thus acts as a switch, which translocates PtdIns(4,5)P2-producing kinases from Dvl toward the Lrp6-phosphorylating complex. This allows efficient phosphorylation of Lrp6 fed by the local production of PtdIns(4,5)P2 associated with ␤-arrestin. Indeed, direct measurements of PtdInsP and PtdInsP2 showed (i) a Wnt3a-induced increase in these PtdInsPs and (ii) an increase in the steady-state levels of PtdInsP in ␤-arrestin1/2 DKO MEFs. These data suggest that the defects in the dynamic Wntinduced phosphorylation of PtdIns are compensated by increase in the level of PtdInsP. Of note, similar phenotype (lack of Wnt-induced dynamics accompanied by increased steadystate activation) has been observed in ␤-arrestin1/2 DKO MEFs for the levels of TOPFLASH activity and of phosphorylated Dvl (14). The known properties of Amer1 make the scenario schematized in Fig. 8A sterically possible. The individual regions of Amer1, which interact with PtdIns(4,5)P2 (7), APC (27), Axin (22), and ␤-arrestin (this study) do not overlap (Fig. 8B). The N-terminal part of Amer1 recognizes PtdIns(4,5)P2 and central regions interact with Lrp6 and Axin, whereas the very C-terminal region is responsible for the interaction with ␤-arrestin. Moreover, Amer1 is, based on the computer predictions, an intrinsically disordered protein with the lack of clearly defined secondary structure (Fig. 8B). This feature allows Amer1 to act as scaffold and to exist in numerous conformations depending on the individual binding partners. Amer1 has a dual role in Wnt/␤-catenin signaling. It was first identified as a negative regulator of Wnt/␤-catenin signaling, which interacts with the components of the destruction complex (Apc, Axin, GSK3␤, CK1) (7, 27, 36). Amer1 was identified only recently also as a positive regulator of the Wnt/␤-catenin signaling, which acts at the level of Lrp6 phosphorylation (7). It is of interest that its positive role seems to be limited to Amer1 and does not apply to related Amer2 (43), which lacks the C-terminal sequence (44) required for the interaction with ␤-arrestin. In summary, in the present study we provide the so far missing molecular mechanism utilized by ␤-arrestin to positively regulate Wnt/␤-catenin signaling. According to our data ␤-arrestin acts in the Wnt/␤-catenin pathway via Amer1, which is a protein conserved only in vertebrates. This raises the possibility that the function of ␤-arrestin in the Wnt/␤-catenin pathway evolved in parallel and will be limited to vertebrates. This is in agreement with the lack of Wnt/␤-catenin-related phenotypes in the Drosophila ␤-arrestin homologue kurtz and the Caenorhabditis elegans homologue arr-1 mutants. Phosphorylation of Lrp6 via the ␤-arrestin/Amer1 pathway thus represents a mechanism for the efficient and tightly controlled activation of the Wnt/␤-catenin pathway that evolved in vertebrates. Acknowledgments—We thank P. De Camilli for PI4KII antibody and PI4KII/PIP5K plasmids; R. J. Lefkowitz for A1CT/A2CT antibodies, ␤-arrestin, and Dvl2 plasmids; M. Maurice, Randall Moon, V. Korˇínek, J. Kukkonen, and S. Yanagawa for plasmids; and K. Soucˇek for the NB4 cell line. REFERENCES 1. Clevers, H. (2006) Wnt/␤-catenin signaling in development and disease. Cell 127, 469–480 2. Logan, C. Y., and Nusse, R. (2004) The Wnt signaling pathway in development and disease. Annu. Rev. Cell Dev. Biol. 20, 781–810 3. Schulte, G., and Bryja, V. (2007) The Frizzled family of unconventional G protein-coupled receptors. Trends Pharmacol. Sci. 28, 518–525 4. Bilic, J., Huang, Y. L., Davidson, G., Zimmermann, T., Cruciat, C. M., Bienz, M., and Niehrs, C. (2007) Wnt induces LRP6 signalosomes and promotes Dishevelled-dependent LRP6 phosphorylation. Science 316, 1619–1622 5. Zeng, X., Huang, H., Tamai, K., Zhang, X., Harada, Y., Yokota, C., Almeida, K., Wang, J., Doble, B., Woodgett, J., Wynshaw-Boris, A., Hsieh, J. C., and He, X. (2008) Initiation of Wnt signaling: control of Wnt coreceptor Lrp6 phosphorylation/activation via Frizzled, Dishevelled, and Axin functions. Development 135, 367–375 6. Pan, W., Choi, S. C., Wang, H., Qin, Y., Volpicelli-Daley, L., Swan, L., Lucast, L., Khoo, C., Zhang, X., Li, L., Abrams, C. S., Sokol, S. Y., and Wu, D. (2008) Wnt3a-mediated formation of phosphatidylinositol 4,5-bisphosphate regulates LRP6 phosphorylation. Science 321, 1350–1353 7. Tanneberger, K., Pfister, A. S., Brauburger, K., Schneikert, J., Hadjihannas, M. V., Kriz, V., Schulte, G., Bryja, V., and Behrens, J. (2011) Amer1/WTX couples Wnt-induced formation of PtdIns(4,5)P2 to LRP6 phosphorylation. EMBO J. 30, 1433–1443 8. Tamai, K., Zeng, X., Liu, C., Zhang, X., Harada, Y., Chang, Z., and He, X. (2004) A mechanism for Wnt coreceptor activation. Mol. Cell 13, 149–156 9. Behrens, J., Jerchow, B. A., Würtele, M., Grimm, J., Asbrand, C., Wirtz, R., Kühl, M., Wedlich, D., and Birchmeier, W. (1998) Functional interaction of an Axin homolog, conductin, with ␤-catenin, APC, and GSK3␤. Science 280, 596–599 10. Kimelman, D., and Xu, W. (2006) ␤-Catenin destruction complex: insights and questions from a structural perspective. Oncogene 25, 7482–7491 11. MacDonald, B. T., Tamai, K., and He, X. (2009) Wnt/␤-catenin signaling: components, mechanisms, and diseases. Dev. Cell 17, 9–26 12. Schulte, G., Schambony, A., and Bryja, V. (2010) ␤-Arrestins: scaffolds and ␤-Arrestin Controls Lrp6 Phosphorylation 1140 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289•NUMBER 2•JANUARY 10, 2014 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom signalling elements essential for WNT/Frizzled signalling pathways? Br. J. Pharmacol. 159, 1051–1058 13. Chen, W., Hu, L. A., Semenov, M. V., Yanagawa, S., Kikuchi, A., Lefkowitz, R. J., and Miller, W. E. (2001) ␤-Arrestin1 modulates lymphoid enhancer factor transcriptional activity through interaction with phosphorylated Dishevelled proteins. Proc. Natl. Acad. Sci. U.S.A. 98, 14889–14894 14. Bryja, V., Gradl, D., Schambony, A., Arenas, E., and Schulte, G. (2007) ␤-Arrestin is a necessary component of Wnt/␤-catenin signaling in vitro and in vivo. Proc. Natl. Acad. Sci. U.S.A. 104, 6690–6695 15. Bryja, V., Schambony, A., Cajánek, L., Dominguez, I., Arenas, E., and Schulte, G. (2008) ␤-Arrestin and casein kinase 1/2 define distinct branches of noncanonical WNT signalling pathways. EMBO Rep. 9, 1244–1250 16. Chen, W., ten Berge, D., Brown, J., Ahn, S., Hu, L. A., Miller, W. E., Caron, M. G., Barak, L. S., Nusse, R., and Lefkowitz, R. J. (2003) Dishevelled 2 recruits ␤-arrestin 2 to mediate Wnt5A-stimulated endocytosis of Frizzled 4. Science 301, 1391–1394 17. Kim, G. H., Her, J. H., and Han, J. K. (2008) Ryk cooperates with Frizzled 7 to promote Wnt11-mediated endocytosis and is essential for Xenopus laevis convergent extension movements. J. Cell Biol. 182, 1073–1082 18. Kim, G. H., and Han, J. K. (2007) Essential role for ␤-arrestin 2 in the regulation of Xenopus convergent extension movements. EMBO J. 26, 2513–2526 19. Kikuchi, A., Yamamoto, H., and Sato, A. (2009) Selective activation mechanisms of Wnt signaling pathways. Trends Cell Biol. 19, 119–129 20. Bernatik, O., Ganji, R. S., Dijksterhuis, J. P., Konik, P., Cervenka, I., Polonio, T., Krejci, P., Schulte, G., and Bryja, V. (2011) Sequential activation and inactivation of Dishevelled in the Wnt/␤-catenin pathway by casein kinases. J. Biol. Chem. 286, 10396–10410 21. Kohout, T. A., Lin, F. S., Perry, S. J., Conner, D. A., and Lefkowitz, R. J. (2001) ␤-Arrestin 1 and 2 differentially regulate heptahelical receptor signaling and trafficking. Proc. Natl. Acad. Sci. U.S.A. 98, 1601–1606 22. Tanneberger, K., Pfister, A. S., Kriz, V., Bryja, V., Schambony, A., and Behrens, J. (2011) Structural and functional characterization of the Wnt inhibitor APC membrane recruitment 1 (Amer1). J. Biol. Chem. 286, 19204–19214 23. Angers, S., Thorpe, C. J., Biechele, T. L., Goldenberg, S. J., Zheng, N., MacCoss, M. J., and Moon, R. T. (2006) The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wnt-␤-catenin pathway by targeting Dishevelled for degradation. Nat. Cell Biol. 8, 348–357 24. Tauriello, D. V., Haegebarth, A., Kuper, I., Edelmann, M. J., Henraat, M., Canninga-van Dijk, M. R., Kessler, B. M., Clevers, H., and Maurice, M. M. (2010) Loss of the tumor suppressor CYLD enhances Wnt/␤-catenin signaling through K63-linked ubiquitination of Dvl. Mol. Cell 37, 607–619 25. Lee, J. S., Ishimoto, A., and Yanagawa, S. (1999) Characterization of mouse Dishevelled (Dvl) proteins in Wnt/Wingless signaling pathway. J. Biol. Chem. 274, 21464–21470 26. Veeman, M. T., Slusarski, D. C., Kaykas, A., Louie, S. H., and Moon, R. T. (2003) Zebrafish prickle, a modulator of noncanonical Wnt/Fz signaling, regulates gastrulation movements. Curr. Biol. 13, 680–685 27. Grohmann, A., Tanneberger, K., Alzner, A., Schneikert, J., and Behrens, J. (2007) Amer1 regulates the distribution of the tumor suppressor APC between microtubules and the plasma membrane. J. Cell Sci. 120, 3738–3747 28. Bryja, V., Pacherník, J., Faldíková, L., Krejcí, P., Pogue, R., Nevrivá, I., Dvorák, P., and Hampl, A. (2004) The role of p27Kip1 in maintaining the levels of D-type cyclins in vivo. Biochim. Biophys. Acta 1691, 105–116 29. Major, M. B., Roberts, B. S., Berndt, J. D., Marine, S., Anastas, J., Chung, N., Ferrer, M., Yi, X., Stoick-Cooper, C. L., von Haller, P. D., Kategaya, L., Chien, A., Angers, S., MacCoss, M., Cleary, M. A., Arthur, W. T., and Moon, R. T. (2008) New regulators of Wnt/␤-catenin signaling revealed by integrative molecular screening. Sci. Signal. 1, ra12 30. Smalley, M. J., Signoret, N., Robertson, D., Tilley, A., Hann, A., Ewan, K., Ding, Y., Paterson, H., and Dale, T. C. (2005) Dishevelled (Dvl-2) activates canonical Wnt signalling in the absence of cytoplasmic puncta. J. Cell Sci. 118, 5279–5289 31. Schwarz-Romond, T., Merrifield, C., Nichols, B. J., and Bienz, M. (2005) The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J. Cell Sci. 118, 5269–5277 32. Korinek, V., Barker, N., Morin, P. J., van Wichen, D., de Weger, R., Kinzler, K. W., Vogelstein, B., and Clevers, H. (1997) Constitutive transcriptional activation by a ␤-catenin-TCF complex in APCϪ/Ϫ colon carcinoma. Science 275, 1784–1787 33. Qin, Y., Li, L., Pan, W., and Wu, D. (2009) Regulation of phosphatidylinositol kinases and metabolism by Wnt3a and Dvl. J. Biol. Chem. 284, 22544–22548 34. Nelson, C. D., Kovacs, J. J., Nobles, K. N., Whalen, E. J., and Lefkowitz, R. J. (2008) ␤-Arrestin scaffolding of phosphatidylinositol 4-phosphate 5-kinase I␣ promotes agonist-stimulated sequestration of the ␤2-adrenergic receptor. J. Biol. Chem. 283, 21093–21101 35. Balla, A., Tuymetova, G., Barshishat, M., Geiszt, M., and Balla, T. (2002) Characterization of type II phosphatidylinositol 4-kinase isoforms reveals association of the enzymes with endosomal vesicular compartments. J. Biol. Chem. 277, 20041–20050 36. Major, M. B., Camp, N. D., Berndt, J. D., Yi, X., Goldenberg, S. J., Hubbert, C., Biechele, T. L., Gingras, A. C., Zheng, N., Maccoss, M. J., Angers, S., and Moon, R. T. (2007) Wilms tumor suppressor WTX negatively regulates WNT/␤-catenin signaling. Science 316, 1043–1046 37. Chen, M., Philipp, M., Wang, J., Premont, R. T., Garrison, T. R., Caron, M. G., Lefkowitz, R. J., and Chen, W. (2009) G protein-coupled receptor kinases phosphorylate LRP6 in the Wnt pathway. J. Biol. Chem. 284, 35040–35048 38. Cˇervenka, I., Wolf, J., Masˇek, J., Krejci, P., Wilcox, W. R., Kozubík, A., Schulte, G., Gutkind, J. S., and Bryja, V. (2011) Mitogen-activated protein kinases promote WNT/␤-catenin signaling via phosphorylation of LRP6. Mol. Cell. Biol. 31, 179–189 39. Krejci, P., Aklian, A., Kaucka, M., Sevcikova, E., Prochazkova, J., Masek, J. K., Mikolka, P., Pospisilova, T., Spoustova, T., Weis, M., Paznekas, W. A., Wolf, J. H., Gutkind, J. S., Wilcox, W. R., Kozubik, A., Jabs, E. W., Bryja, V., Salazar, L., Vesela, I., and Balek, L. (2012) Receptor tyrosine kinases activate canonical WNT/␤-catenin signaling via MAP kinase/LRP6 pathway and direct ␤-catenin phosphorylation. PLoS ONE 7, e35826 40. Davidson, G., Wu, W., Shen, J., Bilic, J., Fenger, U., Stannek, P., Glinka, A., and Niehrs, C. (2005) Casein kinase 1␥ couples Wnt receptor activation to cytoplasmic signal transduction. Nature 438, 867–872 41. Davidson, G., Shen, J., Huang, Y. L., Su, Y., Karaulanov, E., Bartscherer, K., Hassler, C., Stannek, P., Boutros, M., and Niehrs, C. (2009) Cell cycle control of Wnt receptor activation. Dev. Cell 17, 788–799 42. Zeng, X., Tamai, K., Doble, B., Li, S., Huang, H., Habas, R., Okamura, H., Woodgett, J., and He, X. (2005) A dual-kinase mechanism for Wnt coreceptor phosphorylation and activation. Nature 438, 873–877 43. Pfister, A. S., Tanneberger, K., Schambony, A., and Behrens, J. (2012) Amer2 protein is a novel negative regulator of Wnt/␤-catenin signaling involved in neuroectodermal patterning. J. Biol. Chem. 287, 1734–1741 44. Boutet, A., Comai, G., and Schedl, A. (2010) The WTX/Amer1 gene family: evolution, signature and function. BMC Evol. Biol. 10, 280 ␤-Arrestin Controls Lrp6 Phosphorylation JANUARY 10, 2014•VOLUME 289•NUMBER 2 JOURNAL OF BIOLOGICAL CHEMISTRY 1141 atMASARYKOVAUNIVERZITAonJanuary10,2014http://www.jbc.org/Downloadedfrom Vítězslav Bryja, 2014    Attachments      #17      K. Seitz1 , V. Dürsch1 , J. Harnoš, V. Bryja, M. Gentzel, A. Schambony (2014): β‐Arrestin  interacts with the beta/gamma subunits of trimeric G proteins and Dishevelled in the  Wnt‐/Ca2+ pathway in Xenopus gastrulation. Plos ONE 9(1):e87132.  1  equal contribution      Impact factor (2012): 3.730  Times cited (without autocitations, WoS, Feb 21st 2014): 0  Significance:  Identification  of  β‐arrestin  as  a  component  of  the  non‐canonical  Wnt/Ca2+ pathway.  β‐arrestin  mediates activation of protein kinase C together  with Dishevelled and beta/gamma subunits of trimeric G protein.  Contibution  of  the  author/author´s  team:  Immunoprecipitation  analysis  of  the  interaction among Dvl1/2/3‐β‐arrestin‐beta/gamma‐G proteins.                b-Arrestin Interacts with the Beta/Gamma Subunits of Trimeric G-Proteins and Dishevelled in the Wnt/Ca2+ Pathway in Xenopus Gastrulation Katharina Seitz1. , Verena Du¨ rsch1. , Jakub Harnosˇ2 , Vitezslav Bryja2,3 , Marc Gentzel4" , Alexandra Schambony1,*" 1 Biology Department, Developmental Biology, Friedrich-Alexander University Erlangen-Nuremberg, Erlangen, Germany, 2 Institute of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republic, 3 Institute of Biophysics, Academy of Sciences of the Czech Republic, Brno, Czech Republic, 4 Max Planck Institute of Molecular Cell Biology and Genetics, Dresden, Germany Abstract b-Catenin independent, non-canonical Wnt signaling pathways play a major role in the regulation of morphogenetic movements in vertebrates. The term non-canonical Wnt signaling comprises multiple, intracellularly divergent, Wntactivated and b-Catenin independent signaling cascades including the Wnt/Planar Cell Polarity and the Wnt/Ca2+ cascades. Wnt/Planar Cell Polarity and Wnt/Ca2+ pathways share common effector proteins, including the Wnt ligand, Frizzled receptors and Dishevelled, with each other and with additional branches of Wnt signaling. Along with the aforementioned proteins, b-Arrestin has been identified as an essential effector protein in the Wnt/b-Catenin and the Wnt/Planar Cell Polarity pathway. Our results demonstrate that b-Arrestin is required in the Wnt/Ca2+ signaling cascade upstream of Protein Kinase C (PKC) and Ca2+ /Calmodulin-dependent Protein Kinase II (CamKII). We have further characterized the role of b-Arrestin in this branch of non-canonical Wnt signaling by knock-down and rescue experiments in Xenopus embryo explants and analyzed protein-protein interactions in 293T cells. Functional interaction of b-Arrestin, the b subunit of trimeric G-proteins and Dishevelled is required to induce PKC activation and membrane translocation. In Xenopus gastrulation, b-Arrestin function in Wnt/Ca2+ signaling is essential for convergent extension movements. We further show that b-Arrestin physically interacts with the b subunit of trimeric G-proteins and Dishevelled, and that the interaction between b-Arrestin and Dishevelled is promoted by the beta/gamma subunits of trimeric G-proteins, indicating the formation of a multiprotein signaling complex. Citation: Seitz K, Du¨rsch V, Harnosˇ J, Bryja V, Gentzel M, et al. (2014) b-Arrestin Interacts with the Beta/Gamma Subunits of Trimeric G-Proteins and Dishevelled in the Wnt/Ca2+ Pathway in Xenopus Gastrulation. PLoS ONE 9(1): e87132. doi:10.1371/journal.pone.0087132 Editor: Ali H. Brivanlou, The Rockefeller University, United States of America Received August 7, 2013; Accepted December 18, 2013; Published January 29, 2014 Copyright: ß 2014 Seitz et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Funding: This work was supported by Deutsche Forschungsgemeinschaft (SCHA965/6-1 to A.S., www.dfg.de), the Max-Planck-Society (to M.G., www.mpg.de) and by Czech Science Foundation (204/09/0498, 204/09/J030 to V.B., www.gacr.cz) and EMBO Installation Grant, the Ministry of Education Youth and Sport of the Czech Republic (MSM0021622430 to V.B., www.embo.org). In addition, the authors acknowledge support by Deutsche Forschungsgemeinschaft (www.dfg.de) and Friedrich-Alexander-Universita¨t Erlangen-Nu¨rnberg (www.fau.de) within the funding programme Open Access Publishing. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. Competing Interests: The authors have declared that no competing interests exist. * E-mail: alexandra.schambony@fau.de . These authors contributed equally to this work. " These authors also contributed equally to this work. Introduction Wnt signaling plays a crucial role in pattern formation, tissue specification and cellular organization during embryogenesis. Wnt signaling pathways are generally subdivided into the canonical Wnt pathway, which leads to stabilization and nuclear translocation of b-Catenin, and the more divergent non-canonical, bCatenin independent pathways. Both b-Catenin dependent and independent Wnt pathways are activated by heteromeric receptor complexes of Frizzled family seven-pass transmembrane receptors and LRP5/6 or PTK7/Ryk/Ror2 co-receptors, respectively. The b-Catenin independent Wnt pathways encompass a number of biochemically and functionally distinct signaling cascades, including the Wnt/PCP and the Wnt/Ca2+ pathways. Wnt/PCP signaling leads to a Dishevelled (Dvl) mediated activation of Rho family small GTPases (RhoA, Rac1) and subsequent activation and phosphorylation of c-Jun N-terminal Kinase (JNK [1]). The Wnt/Ca2+ pathway is characterized by Gai/o-triggered, Pertussis Toxin (PTX)-sensitive calcium release and activation of Ca2+ regulated effector proteins, including Protein Kinase C alpha (PKCa), Ca2+ /Calmodulin-dependent Protein Kinase II (CamKII) and Nuclear Factor of Activated T-Cells (NFAT [2]). In multiple developmental processes, including gastrulation and cardiac development, Wnt/PCP and Wnt/Ca2+ signaling are required simultaneously and are often activated by the same Wnt ligand and Frizzled receptor. Wnt-11 and Frizzled 7 activated PKC signaling is required for tissue separation of mesoderm and ectoderm during gastrulation [3]. In the mesoderm, Wnt-11, Frizzled 7 (Fzd7), Dvl2 and b-Arrestin2 (Arrb2) activate Wnt/PCP signaling during convergent extension movements [4–7]. Similarly, Wnt-11 mediates both Wnt/Ca2+ and Wnt/PCP signaling in cardiac development [8,9]. Recent reviews have depicted Wnt signaling cascades as a signaling network rather than as distinct PLOS ONE | www.plosone.org 1 January 2014 | Volume 9 | Issue 1 | e87132 signaling cascades [10], however, the biochemical interaction and integration of these different branches of non-canonical Wnt signaling is yet unclear. b-Arrestins were initially described as proteins involved in desensitizing and endocytosis of G-protein coupled receptors (GPCRs [11]). More recent studies revealed that b-Arrestins play a role in multiple signal transduction pathways, including canonical and non-canonical Wnt signaling (for review [11,12]). In contrast to classical GPCR desensitizing, b-Arrestin2 acts as a positive regulator in Wnt signaling. We have previously shown that bArrestin2 interacts with Dvl and is required for Wnt/b-Catenin signal transduction [13]. Others and we have also described a role for b-Arrestin2 in the activation of the PCP pathway [7,14]. Here we show that b-Arrestin2 is required in the Wnt/Ca2+ signaling cascade; it interacts functionally and physically with the b subunit of trimeric G-proteins and Dishevelled, which are both known effectors in the Wnt/Ca2+ cascade [15,16]. We show further that Wnt/Ca2+ signaling is essential for proper convergent extension movements in Xenopus embryos and that b-Arrestin2 functionally links Wnt/Ca2+ and Wnt/PCP signaling in convergent extension movements. Materials and Methods Xenopus laevis Embryos Xenopus embryos were generated and cultured according to general protocols and staged according to the normal table of Nieuwkoop and Faber [17]. All procedures were performed according to the German animal use and care law (Tierschutzgesetz) and approved by the German state administration Bavaria (Regierung von Mittelfranken). Cell Culture and Transfection 293T human embryonic kidney cells (Leibniz Institute Collections of Microorganisms and Cell Culture, DSMZ, Germany) were cultured in DMEM supplemented with 10% fetal calf serum (Life Technologies, CA, USA) at 37uC in a humidified atmosphere of 10% CO2. Plasmid transfections were performed using either TransPassD2 (New England Biolabs, MA, USA) or Nanofectin (GE Healthcare, Freiburg, Germany) according to the manufacturers’ protocols. Plasmids and Morpholinos The following plasmids and Morpholinos have been described previously: pcDNA-flag-arrb2 [13], pCS2+ fzd7, pCS2+ rhoA V12, pCS2+ rac1 V14 [18], pCS2+ cdc42 [19]; Arrb2 Morpholino 1 (Arrb2 MO1, [13]), Fzd7 Morpholino [3], Dvl2 MO [7], Dvl1 MO, Dvl3 MO [20]. The Arrb2 Morpholino 2 (Arrb2 MO2 59-CGCACGGTTCCAAACGCACAGTAGG- 39) and a Control Morpholino (Control MO) were purchased from Gene Tools LLC (OR, USA). The additional 59UTR sequence of the arrb2 mRNA has been submitted to Genebank (accession number KF831094). The expression plasmids pCS2+ pkca-gfp, pCS2+ dn pkca, pCS2+ camkII K42M, pCS2+ camkII T286D were generously provided by Michael Ku¨hl (University Ulm, Germany); pcDNA HA-gb1 (HAgnb1), pcDNA HA-gc2 (HA-gng2) and pRK5 b-ARKct were provided by G. Schulte (Karolinska Institute, Stockholm, Sweden). The fragment encoding b-ARKct was subcloned into pCS2+. The open reading frames encoding Xenopus Dvl1, Dvl2 and Dvl3 were amplified by PCR and cloned into pCS2+ myc (R. Rupp, LMU Munich, Germany). Preparation of Cell Lysates, Immunoprecipitation, and Western Blotting Cells were washed once with PBS and lysed in NP-40 buffer (20 mM Tris-HCl (pH 7.4), 150 mM NaCl, 2 mM EDTA, 1% NP-40) supplemented with complete Protease Inhibitor and PhosStop Phosphatase Inhibitor Cocktails (Roche, Mannheim, Germany) at 4uC. For embryo lysates, embryos were collected at the desired stage and lysed in the same lysis buffer. Animal Cap lysates were also prepared using the same lysis buffer. Lysates were cleared at 16,0006 g for 10 min. For co-immunoprecipitation, lysates were incubated for 4 h at 4uC with the appropriate antibody and protein G-magnetic beads (Life Technologies, CA, USA). Immunoprecipitates were collected, washed four times with lysis buffer and eluted with SDS sample buffer. For Western blotting, proteins were visualized colorimetrically with NBT/ BCIP. Antibodies Commercial antibodies were obtained from Abcam, UK (mouse anti-Arrb2, rabbit anti-GFP; rabbit anti-HA, goat anti-myc), Cell Signaling Technology Inc., USA (rabbit anti-Arrb2, rabbit antimyc, rabbit anti-Flag), ProteinTech Inc., USA (rabbit anti-Arrb2) and Santa Cruz Biotechnology Inc., USA (mouse anti-Gb, mouse anti-b-Catenin, rabbit anti-Dvl2). The anti-tubulin b hybridoma developed by Michael Klymkowski and the anti-actin antibody developed by Jim Jung-Chin Lin were obtained from the Developmental Studies Hybridoma Bank developed under the auspices of the NICHD and maintained by The University of Iowa, Department of Biology, Iowa City, IA 52242. Secondary antibodies were anti-mouse-Alkaline Phosphatase and anti-rabbitAlkaline Phosphatase (Cell Signaling Technology, Inc. USA), antimouse Cy3 (Jackson ImmunoResearch, PA, USA), anti-rabbit Alexa 488 and anti-mouse Alexa 647 (Life Technologies, CA, USA). Injection and Analysis of Xenopus laevis Embryos RNA for microinjection was prepared using the mMessage mMachine Kit (Life Technologies, CA, USA). Injection amounts were 200 pg for ca camkII (camkII T286D), dn camkII (camkII K42M), and for myc-arrb2, myc-arrb1, HA-gb1 (HA-gnb1), and HA-gc2 (HAgng2); injection amounts were 500 pg for pkca, dn pkca, b-ARKct, and 1 ng for pkca-gfp and fzd7. For ca rhoA 5 pg and ca rac1 10 pg plasmid DNA were injected. Knock-down was achieved by injection of the following antisense Morpholino oligonucleotides: Dvl2 MO, Dvl1 MO, and Dvl3 MO (0.8 pmol each), Arrb2 MO1 (0.8 pmol), Arrb2 MO2 (0.8 pmol), Fzd7 MO (4 pmol). Pertussis Toxin (PTX) was added to the culture medium at 100 ng/ml. Embryos were injected at the two-cell stage for Animal Cap explants or at the four-cell stage in both dorsal blastomeres for Keller explants and cultured until they reached stage 10 or 10.5, respectively. Keller open face explants were prepared and cultured as described in [18]. Explants were scored as ‘‘fully elongated’’ if they showed .75% elongation, as ‘‘partially elongated’’ if elongation was between 25% and 75% and ‘‘not elongated’’ if explants showed less than 25% elongation when compared to fully elongated control explants. Immunofluorescence staining of Animal Cap explants was performed as described previously [21]. Photographs were taken on a Zeiss Apotome imaging system (Zeiss, Oberkochen, Germany). b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 2 January 2014 | Volume 9 | Issue 1 | e87132 Results b-Arrestin2 is required for PKCa activation in Wnt/Ca2+ signaling Wnt- or Frizzled-induced PKCa activation and translocation to the plasma membrane is an indicator for the activation of Wnt/ Ca2+ signaling. We therefore investigated the role of b-Arrestin2 in Wnt/Ca2+ signaling by monitoring PKCa translocation in Xenopus Animal Cap explants. As expected from earlier studies [3], overexpression of Xenopus Frizzled 7 (Fzd7) induced a robust translocation of PKCa-GFP to the plasma membrane (Figure 1A, B). Interestingly, overexpression of b-Arrestin2 (Arrb2, Figure 1C) was also sufficient to induce partial membrane association of PKCa-GFP, likely by enhancing endogenous signaling. For further analysis knock-down of Arrb2 was achieved using two different translation-blocking antisense Morpholinos (Arrb2 MO1 [13] and Arrb2 MO2). Both Morpholinos repressed Frizzled 7 induced PKCa-GFP association with the plasma membrane to comparable extent (Figure 1D and Figure 1E). In contrast, a Control MO had no effect on Fzd7-mediated PKCa-GFP membrane localization (Figure 1F). The inhibition of Fzd7induced PKCa-GFP translocation by Arrb2 MO1 or Arrb2 MO2 was rescued by co-injection of a Morpholino insensitive mycarrb2 RNA. Expression of myc-tagged Arrb2 restored PKCa-GFP localization to the plasma membrane in the presence of Arrb2 MO1 (Figure 1G, G9, G") or Arrb2 MO2 (Figure 1H, H9, H"), which confirmed the specificity of both Arrb2 Morpholino oligonucleotides used in this study. To determine the efficiency of protein depletion by the Morpholino oligonucleotides used herein, lysates of stage 10 Animal Cap explants were tested with an anti-Arrb2 antibody to detect the endogenous protein. Both Arrb2 MO1 and Arrb2 MO2 efficiently reduced endogenous Arrb2 protein levels in Animal Cap explants (Figure 1I), which confirmed the capability and efficiency of these Morpholino oligonucleotides to deplete endogenous Arrb2 in Animal Cap tissue. The novel Arrb2 MO2 has been designed to bind in the 59UTR of arrb2 mRNA, thereby targeting a non-overlapping sequence compared to Arrb2 MO1 (Figure S1), which targets the translation start. Although arrb2 and arrb1 mRNA sequences are only 48% identical in the overall binding region of Arrb2 MO1, they share 12 out of 15 bases in the sequence surrounding the ATG codon (Figure S1). Therefore, the specificity of Arrb2 MO1 to Arrb2 as compared to Arrb1 was further confirmed using MO-sensitive GFP fusion constructs and by co-injecting PKCa-GFP expressing Animal Caps with Fzd7, Arrb2 Morpholinos and arrb1 RNA. We observed that Arrb2 MO1 was not able to suppress translation of GFP constructs fused to the arrb1 59 UTR, while it efficiently blocked translation when the arrb2 59UTR was present (Figure S2A). In contrast to myc-Arrb2 (Figure 1G, H), myc-Arrb1 only partially restored Fzd7-induced PKCa-GFP translocation in Animal Caps injected with Arrb2 MO1 or Arrb2 MO2 (Figure S2B-S2E). These results further confirmed the specificity of the Arrb2 Morpholino oligonucleotides used here. Moreover, we found that mRNA levels of arrb1 were on average more than 500 fold lower than those of arrb2 in gastrula stage embryos (Figure S3). Overall, we conclude that Arrb2 is required downstream of Frizzled 7 for membrane translocation and activation of PKC in the Wnt/Ca2+ pathway. b-Arrestin2 functionally interacts with the b subunit of trimeric G-proteins Frizzled-induced PKCa translocation requires activation of trimeric G-proteins and signal transduction mediated by the band c-subunits (Gb and Gc) [16], which was recapitulated in our experiments by treatment of Animal Cap explants with Pertussis Toxin (PTX) for 1 hour (Figure 2A, B, C). Overexpression of HAtagged Gb1 and Gc2 (HA-Gnb1 and HA-Gng2, respectively) resulted in PKCa activation and translocation in the HA-positive cells (Figure 2D, D9, D"). Interestingly, expression of the b and c subunits of trimeric G-proteins was sufficient to revert the effect of b-Arrestin2 knock-down on PKC localization. Co-injection of Arrb2 MO1 blocked Fzd7-induced association of PKCa-GFP with the plasma membrane (Figure 2E and Figure 1C). PKCa translocation was restored by co-expression of HA-Gb and HAGc (Figure 2F, F9, F") indicating that Gb/Gc signaling activated PKCa independent or downstream of Arrb2. In the reverse experiment, endogenous Gb activity was blocked by the Gbsequestering b-ARKct that is derived from the C-terminus of bAdrenergic receptor kinase 2 (Adrbk2, [22]). b-ARKct efficiently inhibited Fzd7 induced PKCa-GFP translocation (Figure 2G). Coexpression of Arrb2 was not sufficient to rescue PKCa-GFP translocation in Animal Caps overexpressing b-ARKct (Figure 2H), although Arrb2 localization at the plasma membrane was not affected (Figure 2H9, H"). These results suggested that Arrb2 depends on Gb signaling to induce PKC activation and translocation. b-Arrestin2 functionally interacts with Dishevelled in Wnt/Ca2+ signaling In our previous studies, we have found that the interaction between b-Arrestin2 and Dvl is required for Wnt signal transduction in the Wnt/b-Catenin [13] and in the Wnt/PCP pathways [7]. In addition, it has been shown that overexpression of DvlDDIX, a Dvl2 mutant that activates b-Catenin independent Wnt pathways, is capable of activating PKC and CamKII [15]. In Xenopus as in other vertebrates, three Dvl isoforms have been identified, and an initial study suggested that at least Xenopus Dvl1 and Dvl2 are functionally redundant while Dvl3 showed a different expression pattern and differential functionality in tadpole stage embryos [20]. To analyze the functional interaction of Dvl and Arrb2 in Xenopus gastrulation, we first confirmed that Dishevelled was required for Fzd7-induced PKCa-GFP translocation to the plasma membrane in Animal Cap explants. To avoid redundancy and to obtain sufficient depletion of Dvl proteins, which are already present maternally [20], we simultaneously knocked-down Dvl1, Dvl2 and Dvl3. As expected, Fzd7-induced PKCa translocation was impaired in triple Dvl morphant embryos (Figure 3A, B). Interestingly, overexpression of Arrb2 rescued PKCa membrane translocation (Figure 3C), indicating that Arrb2 functionally interacted with Dvl in Wnt/Ca2+ signaling. Convergent extension (CE) movements of the dorsal mesoderm are tightly regulated mass cell movements during vertebrate gastrulation, which require the activation of b-Catenin independent Wnt signaling pathways [23,24]. Consistently, triple Dvl knock-down impaired elongation of Keller open face explants, which recapitulate CE movements of the dorsal mesoderm [25]. The majority of explants from triple Dvl morphant embryos showed either no elongation or only partial elongation (Figure 3D), and this elongation phenotype was rescued by co-injection of either arrb2 or pkca mRNA, confirming again the role of Dvl in Wnt/Ca2+ signaling and the functional interaction of Dvl and Arrb2. To further characterize the role of b-Arrestin in Wnt/Ca2+ signaling and its function in the regulation of CE movements, we performed additional rescue experiments in Keller open face explants. b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 3 January 2014 | Volume 9 | Issue 1 | e87132 Explants from Frizzled 7 morphant embryos showed only mild elongation defects. However, in the vast majority of elongating explants, strong constriction defects were observed (Figure 4A). The Fzd7 morphant phenotype was fully rescued by co-injection of a mRNA encoding arrb2 as well as by pkca mRNA (Figure 4A), confirming the requirement of Arrb2 in Fzd7-mediated Ca2+ dependent signaling and the role of this branch of Wnt signaling in CE movements [26,27]. In contrast to Fzd7 knock-down, treatment of Keller explants with PTX completely abrogated CE movements (Figure 4B). Scoring for constriction defects was not applicable in non-elongating explants, and we did not observe significant constriction defects under the conditions that rescue the PTX phenotype (not shown). PTX treatment was not effective in explants overexpressing PKCa or caCamKII, which further emphasized the role of Ca2+ -dependent signaling in CE movements. However, overexpression of Arrb2 was not sufficient to restore CE movement in PTX-treated explants, which further supported the conclusion that b-Arrestin2 signaling to PKC was dependent on trimeric G-protein activity. Previously, we have reported that b-Arrestin2 is required for convergent extension movements in Xenopus embryos as an effector in the Wnt/PCP pathway [7]. As shown in this previous Figure 1. Arrb2 is required for membrane translocation of PKCa. Xenopus embryos were injected with 500 pg pkca-gfp RNA and co-injected as indicated above the images. Animal Caps were prepared at stage 10 and immunostained as indicated. Nuclei were stained with Hoechst 33258 (blue). Images show representative results of at least three independent experiments with a minimum of six Animal Caps per experiment. Scale bars: 50 mm. (A) PKCa-GFP localized predominantly to the cytoplasm. (B) Co-injection of 1ng fzd7 RNA induced PKCa-GFP translocation to the plasma membrane. (C) Overexpression of Arrb2 partially induced PKCa-GFP membrane translocation. Co-injection of fzd7 mRNA with 0.8 pmol Arrb2 MO1 (D) or 0.8 pmol Arrb2 MO2 (E) blocked PKCa-GFP localization to the plasma membrane, while a Control MO had no effect (F). (G) The inhibitory effect of Arrb2 MO1 on PKCa-GFP membrane translocation was rescued by co-expression of myc-Arrb2 (anti-myc (red): G9, merge: G"). (H) Comparably, inhibition of PKCa-GFP membrane translocation by Arrb2 MO2 was restored by co-expression of myc-Arrb2 (anti-myc (red): H9, merge: H"). (I) Western Blot of Animal Cap lysates prepared from stage 10 embryos, which were either uninjected or injected with Control MO, Arrb2 MO1 or Arrb2 MO2 as indicated. Both Morpholinos efficiently downregulated proteins levels of endogenous Arrb2; an immunoblot for b-Actin is shown as loading control. doi:10.1371/journal.pone.0087132.g001 b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 4 January 2014 | Volume 9 | Issue 1 | e87132 study, knock-down of Arrb2 strongly inhibited explant elongation when compared to control explants (Figure 4C). Interestingly, Arrb2 loss-of-function (LOF) was again fully rescued by coinjection of pkca mRNA or ca camkII mRNA and surprisingly, a partial, but not significant rescue was observed by co-injection of dn camkII mRNA. A Morpholino-insensitive arrb2 mRNA also rescued the CE phenotype in Arrb2 morphant explants, which served as specificity control for the Morpholino-induced CE phenotype (Figure 4C). Overall, these results demonstrated that b-Arrestin2 is required downstream of Fzd7 and upstream of PKC in Wnt/Ca2+ signaling. Moreover, we confirmed a functional interaction between the beta and gamma subunits of trimeric G-proteins and Dishevelled in signal transduction from Fzd7 to PKC in the regulation of CE movements during Xenopus gastrulation. Arrb2 forms a complex with Dvl and Gbc Frizzled receptors belong to the GPCR superfamily and have been shown to interact with Dishevelled [21], trimeric G-proteins [28,29] and likely indirectly with b-Arrestin2 [30]. In addition, the b subunit of trimeric G-proteins has been found to interact with bArrestin1 [31] and Dvl [32]; and in our previous studies, we have found that the interaction between b-Arrestin2 and Dvl is required for Wnt signal transduction in the Wnt/b-Catenin [13] and in the Wnt/PCP pathways [7]. Moreover, our present results clearly demonstrated a functional interaction between b-Arrestin2, Gbc and Dishevelled in early Xenopus embryos. These observations prompted us to further investigate the physical interaction among these proteins. Co-immunoprecipitation experiments showed that all three Dvl isoforms formed a complex with Gb1 and b-Arrestin2, and all Figure 2. Arrb2 depends on Gbc to induce membrane translocation of PKCa. Xenopus embryos were injected with 500 pg pkca-gfp RNA and co-injected as indicated above the images. Animal Caps were prepared at stage 10 and immunostained as indicated. Nuclei were stained with Hoechst 33258 (blue). Images show representative results from at least two independent experiments with a minimum of six Animal Caps per experiment. Scale bars: 50 mm. (A) PKCa-GFP control, PKCa-GFP localized predominantly to the cytoplasm. (B) Co-injection of 1ng fzd7 RNA induced PKCa-GFP translocation to the plasma membrane. (C) Treatment with PTX for 1 hour blocked Fzd7-induced PKCa-GFP translocation. (D) Overexpression of HA-Gb and HA-Gc subunits (indicated as HA-Gbc) induced PKCa-GFP translocation (HA-Gbc stained with anti-HA (red): D9, merge: D"). The inhibitory effect of Arrb2 MO1 (E) on PKCa-GFP membrane translocation was rescued by (F) co-injection of HA-Gb and HA-Gc mRNA (anti-HA (red): F9, merge: F"). (G) Overexpression of the Gb-sequestering b-ARKct also blocked Fzd7 induced PKCa-GFP translocation. (H) Co-expression of mycArrb2 in Fzd7 and b-ARKct injected Animal Caps was not sufficient to rescue PKCa-GFP membrane translocation (anti-myc (red): H9, merge: H"). doi:10.1371/journal.pone.0087132.g002 b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 5 January 2014 | Volume 9 | Issue 1 | e87132 enhanced binding of Gb1 to b-Arrestin2, indicating that any Dvl is capable to form a trimeric complex with b-Arrestin2 and Gb1 (Figure 5A). The endogenous existence of an Arrb2-Dvl-Gb complex was confirmed by co-precipitation of Gb and Dvl2 with Arrb2 from unstimulated and Wnt stimulated HEK293T cells (Figure 5B), although the overall low amounts of co-precipitated protein indicated that only a fraction of the respective proteins present in the cell are assembled in this complex. Subsequently, we investigated if the interaction of Gb1 with bArrestin2 or Dvl2 was affected by overexpression of the Gbbinding C-terminal fragment of b-Adrenergic receptor kinase (bARKct). Overexpression of b-ARKct strongly interfered with the binding between Arrb2 and Gb1 (Figure 5C). However, we still observed binding of Dvl2 to Arrb2 in the presence of b-ARKct, although reduced when compared to the binding in the absence of b-ARKct (Figure 5C). Co-expression of Dvl2 partially restored binding of Gb1 to Arrb2, probably by enhancing the interaction between Arrb2 and residual free Gb1 subunits (Figure 5C). We observed a similarly strong interference of b-ARKct with the interaction between Gb and Dvl2 (Figure 5D), which was again Figure 3. Arrb2 functionally interacts with Dvl in Wnt/Ca2+ signaling. Xenopus embryos were injected with 500 pg pkca-gfp RNA and coinjected as indicated above the images. Animal Caps were prepared at stage 10 and immunostained as indicated. Nuclei were stained with Hoechst 33258 (blue). Images show representative results from at least two independent experiments with a minimum of six Animal Caps per experiment. Scale bars: 50 mm. Fzd7 induced PKCa-GFP translocation (A) was impaired by a triple knock-down of Dvl1, Dvl2 and Dvl3 (B). (C) Co-expression of Arrb2 partially rescued PKCa-GFP translocation in the triple Dvl knock-down. (D) Triple Dvl knock-down inhibited elongation of Keller open face explants. Co-injection of PCKa or Arrb2 mRNA rescued the CE phenotype of triple Dvl morphant explants. The average percentage of explants showing full (75-100%, light grey), partial (25-50%, medium grey) or no elongation (,25%, dark grey) from at least three independent experiments are shown. Asterisks indicate statistically significant deviations in the percentage of fully elongated explants (* p.0.95, t-test). doi:10.1371/journal.pone.0087132.g003 b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 6 January 2014 | Volume 9 | Issue 1 | e87132 partially restored by co-expression of Arrb2. In addition, the amount of Arrb2 that co-precipitated with Dvl2 was clearly reduced in the presence of b-ARKct, indicating that the interaction of Dvl2 and b-Arrestin2 did not strictly depend on Gb signaling, but was significantly enhanced by Gb. Discussion Arrb2 is required in Wnt/Ca2+ signaling In addition to its previously described role as a mediator of signal transduction in the Wnt/b-Catenin and the Wnt/PCP pathway, we have demonstrated that b-Arrestin2 is also an Figure 4. Arrb2 is required for CE movements downstream of Frizzled 7 and upstream of PKCa. Xenopus embryos were injected at 4-cell stage in the marginal zone of both dorsal blastomeres as indicated. CE movements in the dorsal mesoderm were monitored by elongation of Keller open face explants. The average percentage of explants showing full (75-100%, light grey), partial (25-50%, medium grey) or no elongation (,25%, dark grey) from at least three independent experiments (exp.) are shown. Asterisks indicate statistically significant deviations in the percentage of fully elongated explants (** p.0.99,* p.0.95, t-test). (A) Frizzled 7 knock-down had little effect on elongation (bottom graph) but impaired constriction. Only fully elongated explants (represented by light grey columns in bottom graph) were additionally scored for constriction. The percentage of elongated explants that showed normal constriction are shown in the upper graph as average values plus SEM. Co-injection of arrb2 or pkca mRNA fully rescued constriction and ca camkII mRNA partially rescued constriction. (B) PTX treatment impaired elongation of Keller open face explants. Co-injection of pkca or ca camkII mRNA fully rescued explant elongation, while dn camkII or arrb2 mRNA did not. (C) Knock-down of Arrb2 interfered with explant elongation and was rescued by co-injection of pkca, ca camkII or a MO-insensitive arrb2 RNA. doi:10.1371/journal.pone.0087132.g004 b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 7 January 2014 | Volume 9 | Issue 1 | e87132 Figure 5. Arrb2 physically interacts with Gb and Dvl. Epitope-tagged proteins were overexpressed, immunoprecipitated and detected by Western Blotting as indicated. (A) Flag-Arrb2 was co-expressed with a combination of HA-Gb1 and HA-Gc2 to allow the formation of Gbc heterodimers (Gbc). Co-expression of Dvl1, Dvl2 or Dvl3 enhanced the interaction between Arrb2 and Gb1 in co-immunoprecipitation experiments from HEK 293T cells. (B) Endogenous Gb and Dvl2 were detected in immunoprecipitates of endogenous Arrb2 from unstimulated and Wnt-stimulated HEK 293T cells. (C) Binding of Dvl2 to Arrb2 was also observed in the absence of exogenous Gb. Myc-Dvl2 co-precipitated equally well with Flag-Arrb2 b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 8 January 2014 | Volume 9 | Issue 1 | e87132 essential effector in the Wnt/Ca2+ pathway that is required for the activation and membrane translocation of PKCa downstream of Frizzled 7. These results are in contrast to an earlier report [14], which stated that Arrb2 had no function in Wnt/Ca2+ signaling. Herein our results clearly show that Arrb2 knock-down achieved by two different Morpholinos targeting non-overlapping sequences in the arrb2 mRNA blocked Fzd7-induced PKCa translocation to the plasma membrane and that overexpressed Arrb2 was capable to partially induce PKC activation. We also confirmed that the antisense Morpholinos used in this study efficiently reduced endogenous Arrb2 protein levels. Overexpression of Arrb2 rescued Arrb2 MO1 or Arrb2 MO2 and we have also observed a partial rescue by overexpression of Arrb1, which indicated partial functional redundancy. This result is consistent with a recent report showing partial redundancy of Arrb1 and Arrb2 in CE movements [33]. However, while Arrb2 is expressed at constantly high levels during gastrulation, Arrb1 mRNA levels decrease during this phase of development (Figure S3), suggesting that in gastrulating Xenopus embryos predominantly Arrb2 is required. In addition to regulating PKCa membrane translocation, we have shown in this study that the convergent extension phenotype induced by Arrb2 knock-down was rescued by overexpression of PKCa or caCamKII and that Arrb2, PKCa or caCamKII rescued CE movements in Fzd7 morphant explants, further emphasizing the functional requirement of Arrb2 in Wnt/Ca2+ signaling. Functional and physical interaction of xArrb2 and Gb with Dvl In addition to Frizzled receptors, overexpression of Gb and Gc [16] induces PKCa translocation, and it has been shown that Dvl is also able to activate Wnt/Ca2+ signaling [15]. Here, we have investigated the functional interaction of these effector proteins in the Wnt/Ca2+ pathway. Overexpression of Gb and Gc was sufficient to rescue PKCa activation in embryos injected with fzd7 mRNA and Arrb2 MO1. In contrast, PKCa translocation blocked by inhibition of Gb signaling was not rescued by overexpression of Arrb2, indicating that Gb / Gc signaling is either activated downstream of Frizzled and b-Arrestin or that b-Arrestin activity is Gb-/ Gc-dependent. Consistently, Arrb2 only weakly rescued CE movements in explants treated with PTX, but efficiently rescued the CE phenotype induced by Frizzled 7 knock-down. Therefore, we conclude that b-Arrestin2 interacts with trimeric G-proteins to activate PKCa, which is required for CE movements in Xenopus gastrulation. In our own earlier studies, we have demonstrated that bArrestin physically interacts with Dvl [13]. Here we show that bArrestin2 was able to rescue PKCa membrane translocation and CE movements in triple Dvl morphant embryos, which is consistent with the previous studies and further confirmed the functional interaction of b-Arrestin2 and Dishevelled in Xenopus gastrulation. Interaction between Arrb1 and Gb as well as the interaction of Dvl2 with Gb have also been reported [31,32]. In light of the functional interactions of b-Arrestin 2, Dvl and Gbc in Wnt/Ca2+ signaling, we hypothesized that these proteins formed a heterotetrameric complex. We have confirmed that Arrb2 binds to Gb1. This result was expected because Gb also binds to Arrb1 [31] and functional redundancy between Arrb1 and Arrb2 was indeed reported in a recent study by Kim and co-workers [33]. In addition to Arrb2, we observed that all Xenopus Dvl isoforms, Dvl1, Dvl2 and Dvl3 physically interacted with Gb1. The interaction of each Dvl with Gb1 was enhanced in the presence of Arrb2; moreover, the interaction between Arrb2 and Gb1 was increased by any Dvl isoform, although Dvl2 seemed to have the strongest effect on Arrb2-Gb1 binding. Notably, Arrb2 rescued the Frizzled 7 knock-down and the triple Dvl phenotype in Xenopus gastrulation movements, but was not sufficient to restore CE movements in PTX-treated Keller explants or PKCa translocation to the plasma membrane in Animal Cap explants overexpressing the Gb-sequestering Cterminus of b-ARK. Our results further showed that co-expression of Dvl strongly enhanced the interaction of b-Arrestin2 with Gb while sequestering Gb interfered with binding between bArrestin2 and Dvl2. Altogether, we conclude that activation of trimeric G-proteins downstream of Frizzled leads to the formation of a protein complex consisting of b-Arrestin2, the b and c subunits of trimeric G-proteins and Dishevelled. In the Wnt/Ca2+ pathway, this complex likely triggers the activation of PKC and other Ca2+ -dependent effector proteins. Strikingly, b-Arrestin2 and Dishevelled also interact and are essential effectors in the Wnt/b-Catenin pathway and in Wnt/ PCP signaling to the small GTPases RhoA and Rac1 [5,7,14,34,35]. Considering that Wnt/Ca2+ and Wnt/PCP signaling are often required simultaneously [8,9] and are activated by the same Wnt ligand and Frizzled receptor [3–7] our findings support a close functional interaction of Wnt/Ca2+ and Wnt/PCP signaling in the regulation of CE movements in Xenopus embryos. Moreover, Frizzled receptors, which belong to the superfamily of G-protein coupled receptors, have recently been shown to interact with different types of trimeric G-proteins [29], and there is accumulating evidence that trimeric G-proteins play a role in apparently all Wnt/Frizzled signaling cascades [28,29,36]. We have shown here that the beta and gamma subunits of trimeric Gproteins, which dissociate from the alpha subunit upon activation, promote the interaction of b-Arrestin2 and Dishevelled, two proteins that have been identified as essential effectors in Wnt/bCatenin [13], Wnt/PCP [7,14] and in Wnt/Ca2+ signaling (this study). Therefore it can be hypothesized that complex formation involving b-Arrestin2 and Dishevelled and Gbc might be a general mechanism of trimeric G-protein mediated activation of Wnt signaling cascades. Supporting Information Figure S1 Alignment of arrb2 and arrb1 sequences. Alignment of the novel sequence stretch of the arrb2 59UTR including the first 132 nucleotides of arrb2 coding sequence to the corresponding sequences of arrb2 mRNA (NCBI NM_001092112) and arrb1 mRNA (NCBI NM_001094402) generated with ClustalW2. Asterisks indicate nucleotides conserved in all three sequences; morpholino binding regions are highlighted in light grey, the start codon is shown as bold and underlined. (PDF) Figure S2 (A) Specificity of the Arrb2 MO1 antisense Morpholino oligonucleotide. The 59 UTR and coding sequence encoding amino acids 1–180 of arrb2 and the corresponding sequences of the two arrb1 pseudoalleles identified in gastrula stage embryos and termed when Gb1 and Gc2 were overexpressed (Gbc) as in the presence of the Gb-sequestering b-ARKct in HEK 293T cells. By contrast, binding of Gb1 to Arrb2 was impaired by b-ARKct and partially restored by co-expression of Dvl2. (D) When myc-Dvl2 was precipitated, the amount of Flag-Arrb2 and that of Gb1 that co-precipitated with Dvl2 was significantly reduced by the co-expression of b-ARKct. doi:10.1371/journal.pone.0087132.g005 b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 9 January 2014 | Volume 9 | Issue 1 | e87132 arrb1.a and arrb1.b were were cloned in frame with GFP. All plasmids were co-injected with either Control MO or Arrb2 MO1 in 2-cell stage embryos and analyzed for expression of the GFP fusion proteins. An antibody against b-Tubulin served as loading control. (B–E) Arrb1 only weakly influences PKCa-GFP membrane translocation. Xenopus embryos were injected with 500 pg pkca-gfp RNA and co-injected as indicated. PKC-GFP localization was analyzed in Animal Caps at stage 10 immunostained as indicated; nuclei stained with Hoechst 33258 (blue). Images show representative results from at least two independent experiments with a minimum of six Animal Caps per experiment. Scale bars: 50 mm. Overexpression of Arrb1 only weakly changed PKCa-GFP localization (Figure B, C). Consistently, co-injection of myc-arrb1 RNA only partially restored PKCa-GFP membrane association in Animal Cap explants co-injected with Fzd7 and Arrb2 MO1 (D, D9: anti-myc, and D": merge). A comparable result was obtained when myc-arrb1 RNA was co-injected with Fzd7 and Arrb2 MO2 (E, E9: anti-myc, E": merge). (PDF) Figure S3 Detection of arrb1 and arrb2 transcripts in early Xenopus embryos. Total RNA was extracted from Xenopus embryos of the indicated developmental stages, reverse transcribed and arrb1, arrb2 and ornithin decarboxylase (odc) transcripts were amplified from the resulting cDNA. The upper panel shows the PCR fragments separated by agarose gelelectrophoresis from one representative experiment. The lower panel shows the corresponding real-time RT-PCR experiment using a different set of primer pairs (Illumina Eco Real Time PCR system; primer sequences are listed in Table S1). Transcription levels of arrb1 and arrb2 are plotted relative to odc. (PDF) Table S1 RT-PCR Primer sequences. (PDF) Acknowledgments The authors are grateful for helpful discussions and support by Andrej Shevchenko. We further thank Manuel Pauling who helped with the PKC translocation experiments and Hanh Nguyen for critical reading of the manuscript. The expression plasmids encoding PKCalpha-GFP, dnPKC, CamKII K42M, CamKII D286T, HA-Gbeta, HA-Ggamma and betaARKct were kind gifts of M. Ku¨hl and G. Schulte. Author Contributions Conceived and designed the experiments: KS VD VB MG AS. Performed the experiments: KS VD JH AS. Analyzed the data: KS VD JH VB MG AS. Wrote the paper: MG AS. References 1. Klein TJ, Mlodzik M (2005) Planar cell polarization: an emerging model points in the right direction. Annu Rev Cell Dev Biol 21: 155–176. doi:10.1146/ annurev.cellbio.21.012704.132806. 2. Rao TP, Ku¨hl M (2010) An updated overview on Wnt signaling pathways: a prelude for more. Circ Res 106: 1798–1806. doi:10.1161/CIRCRE- SAHA.110.219840. 3. Winklbauer R, Medina A, Swain RK, Steinbeisser H (2001) Frizzled-7 signalling controls tissue separation during Xenopus gastrulation. Nature 413: 856–860. doi:10.1038/35101621. 4. Djiane A, Riou J, Umbhauer M, Boucaut J, Shi D (2000) Role of frizzled 7 in the regulation of convergent extension movements during gastrulation in Xenopus laevis. Development 127: 3091–3100. 5. Kim G-H, Her J-H, Han J-K (2008) Ryk cooperates with Frizzled 7 to promote Wnt11-mediated endocytosis and is essential for Xenopus laevis convergent extension movements. J Cell Biol 182: 1073–1082. doi:10.1083/jcb.200710188. 6. Medina A, Swain RK, Kuerner K-M, Steinbeisser H (2004) Xenopus paraxial protocadherin has signaling functions and is involved in tissue separation. EMBO J 23: 3249–3258. doi:10.1038/sj.emboj.7600329. 7. Bryja V, Schambony A, Caja´nek L, Dominguez I, Arenas E, et al. (2008) Betaarrestin and casein kinase 1/2 define distinct branches of non-canonical WNT signalling pathways. EMBO Rep 9: 1244–1250. doi:10.1038/embor.2008.193. 8. Pandur P, La¨sche M, Eisenberg LM, Ku¨hl M (2002) Wnt-11 activation of a noncanonical Wnt signalling pathway is required for cardiogenesis. Nature 418: 636–641. doi:10.1038/nature00921. 9. Gessert S, Ku¨hl M (2010) The multiple phases and faces of wnt signaling during cardiac differentiation and development. Circ Res 107: 186–199. doi:10.1161/ CIRCRESAHA.110.221531. 10. Kestler HA, Ku¨hl M (2008) From individual Wnt pathways towards a Wnt signalling network. Philos Trans R Soc Lond, B, Biol Sci 363: 1333–1347. doi:10.1098/rstb.2007.2251. 11. Kovacs JJ, Hara MR, Davenport CL, Kim J, Lefkowitz RJ (2009) Arrestin development: emerging roles for beta-arrestins in developmental signaling pathways. Dev Cell 17: 443–458. doi:10.1016/j.devcel.2009.09.011. 12. Schulte G, Schambony A, Bryja V (2010) beta-Arrestins - scaffolds and signalling elements essential for WNT/Frizzled signalling pathways? Br J Pharmacol 159: 1051–1058. doi:10.1111/j.1476-5381.2009.00466.x. 13. Bryja V, Gradl D, Schambony A, Arenas E, Schulte G (2007) Beta-arrestin is a necessary component of Wnt/beta-catenin signaling in vitro and in vivo. Proc Natl Acad Sci USA 104: 6690–6695. doi:10.1073/pnas.0611356104. 14. Kim G-H, Han J-K (2007) Essential role for aˆ-arrestin 2 in the regulation of Xenopus convergent extension movements. EMBO J 26: 2513–2526. doi:10.1038/sj.emboj.7601688. 15. Sheldahl LC, Slusarski DC, Pandur P, Miller JR, Ku¨hl M, et al. (2003) Dishevelled activates Ca2+ flux, PKC, and CamKII in vertebrate embryos. J Cell Biol 161: 769–777. doi:10.1083/jcb.200211094. 16. Sheldahl LC, Park M, Malbon CC, Moon RT (1999) Protein kinase C is differentially stimulated by Wnt and Frizzled homologs in a G-proteindependent manner. Curr Biol 9: 695–698. 17. Nieuwkoop PD, Faber J (1975) External and internal stage criteria in the development of Xenopus laevis. Normal Tables of Xenopus laevis: 162–188. 18. Unterseher F, Hefele JA, Giehl K, De Robertis EM, Wedlich D, et al. (2004) Paraxial protocadherin coordinates cell polarity during convergent extension via Rho A and JNK. EMBO J 23: 3259–3269. doi:10.1038/sj.emboj.7600332. 19. Schambony A, Wedlich D (2007) Wnt-5A/Ror2 regulate expression of XPAPC through an alternative noncanonical signaling pathway. Dev Cell 12: 779–792. doi:10.1016/j.devcel.2007.02.016. 20. Gray RS, Bayly RD, Green SA, Agarwala S, Lowe CJ, et al. (2009) Diversification of the expression patterns and developmental functions of the dishevelled gene family during chordate evolution. Dev Dyn 238: 2044–2057. doi:10.1002/dvdy.22028. 21. Tauriello DVF, Jordens I, Kirchner K, Slootstra JW, Kruitwagen T, et al. (2012) Wnt/b-catenin signaling requires interaction of the Dishevelled DEP domain and C terminus with a discontinuous motif in Frizzled. Proc Natl Acad Sci USA 109: E812–E820. doi:10.1073/pnas.1114802109. 22. Koch WJ, Inglese J, Stone WC, Lefkowitz RJ (1993) The binding site for the beta gamma subunits of heterotrimeric G proteins on the beta-adrenergic receptor kinase. J Biol Chem 268: 8256–8260. 23. Tada M, Concha ML, Heisenberg CP (2002) Non-canonical Wnt signalling and regulation of gastrulation movements. Semin Cell Dev Biol 13: 251–260. 24. Keller R (2005) Cell migration during gastrulation. Curr Opin Cell Biol 17: 533–541. doi:10.1016/j.ceb.2005.08.006. 25. Shih J, Keller R (1992) Cell motility driving mediolateral intercalation in explants of Xenopus laevis. Development 116: 901–914. 26. Ku¨hl M, Geis K, Sheldahl LC, Pukrop T, Moon RT, et al. (2001) Antagonistic regulation of convergent extension movements in Xenopus by Wnt/beta-catenin and Wnt/Ca2+ signaling. Mech Dev 106: 61–76. 27. Penzo-Mende`z A, Umbhauer M, Djiane A, Boucaut J-C, Riou J-F (2003) Activation of Gbetagamma signaling downstream of Wnt-11/Xfz7 regulates Cdc42 activity during Xenopus gastrulation. Dev Biol 257: 302–314. 28. Koval A, Katanaev VL (2011) Wnt3a stimulation elicits G-protein-coupled receptor properties of mammalian Frizzled proteins. Biochem J 433: 435–440. doi:10.1042/BJ20101878. 29. Nichols AS, Floyd DH, Bruinsma SP, Narzinski K, Baranski TJ (2013) Frizzled receptors signal through G proteins. Cell Signal 25: 1468–1475. doi:10.1016/ j.cellsig.2013.03.009. 30. Chen W, Berge ten D, Brown J, Ahn S, Hu LA, et al. (2003) Dishevelled 2 recruits beta-arrestin 2 to mediate Wnt5A-stimulated endocytosis of Frizzled 4. Science 301: 1391–1394. doi:10.1126/science.1082808. 31. Yang M, He RL, Benovic JL, Ye RD (2009) beta-Arrestin1 interacts with the Gprotein subunits beta1gamma2 and promotes beta1gamma2-dependent Akt signalling for NF-kappaB activation. Biochem J 417: 287–296. doi:10.1042/ BJ20081561. 32. Angers S, Thorpe CJ, Biechele TL, Goldenberg SJ, Zheng N, et al. (2006) The KLHL12–Cullin-3 ubiquitin ligase negatively regulates the Wnt–b-catenin pathway by targeting Dishevelled for degradation. Nat Cell Biol 8: 348–357. doi:10.1038/ncb1381. b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 10 January 2014 | Volume 9 | Issue 1 | e87132 33. Kim G-H, Park EC, Lee H, Na H-J, Choi S-C, et al. (2013) aˆ-Arrestin 1 mediates non-canonical Wnt pathway to regulate convergent extension movements. Biochem Biophys Res Commun. doi:10.1016/j.bbrc.2013.04.088. 34. Habas R, Dawid IB, He X (2003) Coactivation of Rac and Rho by Wnt/ Frizzled signaling is required for vertebrate gastrulation. Genes Dev 17: 295– 309. doi:10.1101/gad.1022203. 35. Habas R, Kato Y, He X (2001) Wnt/Frizzled activation of Rho regulates vertebrate gastrulation and requires a novel Formin homology protein Daam1. Cell 107: 843–854. 36. Katanaev VL, Ponzielli R, Se´me´riva M, Tomlinson A (2005) Trimeric G protein-dependent frizzled signaling in Drosophila. Cell 120: 111–122. doi:10.1016/j.cell.2004.11.014. b-Arrestin Interactions in the Wnt/Ca2+ -Pathway PLOS ONE | www.plosone.org 11 January 2014 | Volume 9 | Issue 1 | e87132 Figure S1 arrb2 5’UTR+CDS ----------------------------------------GGCTTCCTACTGTGCGTTTGGAACCGTGCGCGACTCCGCGGAGGTGGGAT 50 NM_001092112 arrb2 ------------------------------------------------------------------------------------------ NM_001094402 arrb1 AGCAGCAGCAGCACATTAATCCCGTCCCCTCCTCTCCCTAATCTCCCCGGGATCTGCACAAAAGCACTCGCCTCTCTCGACCCCTGGCAC 90 arrb2 5’UTR+CDS TGGAGCGCGTGAACCCAGAGCAAGCGGAGGAGCTG--------GGAAGATGGGGGAGAGGGCGGGGACCCGGGTTTTCAAGAAATCCAGC 132 NM_001092112 arrb2 -----------------------GAGGAGGAGCTG--------GGAAGATGGGGGAGAGGGCGGGGACCCGGGTTTTCAAGAAATCCAGC 59 NM_001094402 arrb1 TCCTGCACCTCCTGCCTGTCCTCCTGCACCTCCTGCCCCTCCTGCATGATGGGGGACAAAG---GAACCAGAGTATTTAAGAAGGCGAGT 177 * * *** * * ********* * * * *** * ** ** ***** * ** arrb2 5’UTR+CDS CCTAACTGCAAGCTCACCGTGTACCTTGGAAAGCGAGATTTTGTCGATCACCTGGATCGGGTTGATCCTGTGGATGGCGTCGTCCTTGTG 222 NM_001092112 arrb2 CCTAACTGCAAGCTCACCGTGTACCTTGGAAAGCGAGATTTTGTCGATCACCTGGATCGGGTTGATCCTGTGGATGGCGTCGTCCTTGTG 149 NM_001094402 arrb1 CCAAATGGAAAGCTGACTGTTTACTTGGGCAAGAGGGACTTTGTTGATCACGTAGACGTGGTGGATCCTGTGGATGGGGTGGTGTTGGTG 267 ** ** * ***** ** ** *** * ** *** * ** ***** ****** * ** *** ************** ** ** * *** A xFz7+Arrb2 MO1 + myc-Arrb1 D PKC-GFP D’ myc-Arrb1 D’’ merge Arrb1 C PKC-GFPPKC-GFPB Figure S2 xFz7+Arrb2 MO2 + myc-Arrb1 E PKC-GFP E’ myc-Arrb1 E’’ merge Arrb2 (UTR-180)-GFP Arrb1.a (UTR-181)-GFP Arrb1.b (UTR-181)-GFP Control MO Arrb2 MO 1 + + + + ++ ++ ++ ++ 50 kDa 50 kDa unspecific band Arrb-(1-180/181)-GFP WB anti-GFP WB anti-Tubulin 4 (8 cell) 8 10 10.5 11 12 -RT odc arrb2 arrb1 500 bp 500 bp 500 bp NF stage arrb2 arrb1 relativetranscription 1 0.01 0.1 0.001 0.0001 4 (8 cell) 8 10 10.5 11 12 NF stage Figure S3 Table S1 forward Primer reverse Primer arrb1_1 tggaaagctgactgtttacttgg atctcaaaggtgaatggataggc arrb2_1 ctgacaggtctctgcacctagaa caaaggtcgtgagatgttctctg odc_1 gatgggctggatcgtatcgt tggcagcagtacagacagca arrb1_2 aggaattcagtgcgtttggtaat caaatatcagcatactggcgaac arrb2_2 ctatgaaattcgtgccttctgtg attgcctctccgtggtaatagag odc_2 cattgcagagcctgggagata tccactttgctcattcaccataac       Vítězslav Bryja, 2014    Attachments      #18      R. de Groot1 ; R. Sri Ganji1 ; O. Bernatik; B. Lloyd‐Lewis; K. Seipel; K. Šedová; Z. Zdráhal;  V.M. Dhople; T. Dale; H. Korswagen*, V. Bryja*. Huwe1‐mediated ubiquitylation of Dvl  defines a novel negative feedback loop in the Wnt signaling pathway. Sci. Signal. (in  press,  appears  in  March  18,  2014  issue).  Selected  article  highlighted  by  an  online  podcast interviewing H. Korswagen and V. Bryja.  1  equal contribution, *corresponding authors      Impact factor (2012): 7.648  Times cited (without autocitations, WoS, Feb 21st 2014): 0  Significance: This study identifies a novel negative feedback loop in the Wnt signaling  pathway.  It  establishes  that  following  activation  and  phosphorylation  by  casein  kinase 1 epsilon Dvl is ubiquitylated by E3 ligase Huwe1. The ubiquitylation takes  place in the  DIX domain, sterically prevents Dvl multimerization  and inhibits its  function.  Contibution  of  the  author/author´s  team:  We  have  discovered  the  interaction  of  Huwe1  and  Dishevelled,  and  performed  most  experiments  in  the  mammalian  experimental system.                D E V E L O P M E N T A L B I O L O G Y Huwe1-Mediated Ubiquitylation of Dishevelled Defines a Negative Feedback Loop in the Wnt Signaling Pathway Reinoud E. A. de Groot,1 * Ranjani S. Ganji,2 * Ondrej Bernatik,2,3 Bethan Lloyd-Lewis,4† Katja Seipel,4‡ Kateřina Šedová,5,6 Zbyněk Zdráhal,5,6 Vishnu M. Dhople,7 Trevor Dale,4 Hendrik C. Korswagen,1§ Vitezslav BryjaAQ1 2,3§ Wnt signaling plays a central role in development, adult tissue homeostasis, and cancer. Several steps in the canonical Wnt/b-catenin signaling cascade are regulated by ubiquitylation, a protein modification that influences the stability, subcellular localization, or interactions of target proteins. To identify regulators of the Wnt/b-catenin pathway, we performed an RNA interference screen in Caenorhabditis elegans and identified the HECT domain–containing ubiquitin ligase EEL-1 as an inhibitor of Wnt signaling. In human embryonic kidney 293T cells, knockdown of the EEL-1 homolog Huwe1 enhanced the activity of a Wnt reporter in cells stimulated with Wnt3a or in cells that overexpressed casein kinase 1 (CK1) or a constitutively active mutant of the Wnt co-receptor low-density lipoprotein receptor–related protein 6 (LRP6). However, knockdown of Huwe1 had no effect on reporter gene expression in cells expressing constitutively active b-catenin, suggesting that Huwe1 inhibited Wnt signaling upstream of b-catenin and downstream of CK1 and LRP6. Huwe1 bound to and ubiquitylated the cytoplasmic Wnt pathway component Dishevelled (Dvl) in a Wnt3a- and CK1e-dependent manner. Mass spectrometric analysis showed that Huwe1 promoted K63-linked, but not K48-linked, polyubiquitination of Dvl. Instead of targeting Dvl for degradation, ubiquitylation of the DIX domain of Dvl by Huwe1 inhibited Dvl multimerization, which is necessary for its function. Our findings indicate that Huwe1 is part of an evolutionarily conserved negative feedback loop in the Wnt/bcatenin pathway. INTRODUCTION In the canonical Wnt signaling pathway, the stability of the Wnt pathway effector b-catenin is regulated by a destruction complex, composed of the adenomatous polyposis coli protein (APC), Axin, casein kinase 1 (CK1), and glycogen synthase kinase 3b (GSK3b), which phosphorylates b-catenin and targets it for b-transducin repeat–containing protein (b-TrCP)–dependent ubiquitylation and proteasomal degradation. Binding of Wnt to the receptors Frizzled (Fz) and low-density lipoprotein receptor–related protein 6 (LRP6) leads to inhibition of destruction complex function and accumulation of b-catenin. b-Catenin can translocate to the nucleus and interact with members of the T cell factor/lymphoid enhancer factor (TCF/LEF) family of transcription factors to induce target gene expression (1). The binding of Wnt to Fz and LRP6 induces clustering of the two receptors with the cytoplasmic protein Dishevelled (Dvl) into signalosomes at the plasma membrane (2). Dvl plays a key role during these upstream events in Wnt pathway activation. Upon Wnt stimulation, Dvl is phosphorylated by CK1e (3, 4) and Axin is recruited into the Dvl complex. Subsequent phosphorylation of LRP6 by CK1 and GSK3ß induces binding of Axin to LRP6 and inhibition of destruction complex function. An important property of Dvl is that it can multimerize through an N-terminal DIX domain (5, 6). The formation of such dynamic aggregates is thought to increase the overall avidity of Dvl for its interaction partners, and mutations in the DIX domain that interfere with multimerization show strongly reduced signaling activity (5). In addition to phosphorylation, Dvl is also regulated by ubiquitylation (7–9). Ubiquitylation is a versatile posttranslational modification in which specific E3 ubiquitin ligases mediate the addition of ubiquitin molecules— either as single ubiquitin proteins or as ubiquitin polymers—to substrate proteins. Moreover, ubiquitin chains can be linked through different lysines within the ubiquitin sequence, which further increases the functional diversity of this modification. Depending on the type of ubiquitylation, the modified protein is targeted for proteasomal degradation, endocytosed, and routed into the lysosomal degradation pathway, or is affected in its ability to interact with other proteins (10). In the case of Dvl, K48-linked polyubiquitylation by kelch-like family member 12 (KLHL12) controls Dvl stability (7), but additional ubiquitin modifications may control other aspects of Dvl function (8, 11). Here, we identified the E3 ubiquitin ligase Huwe1 as an evolutionarily conserved inhibitor of the Wnt/b-catenin signaling pathway. We demonstrate that Huwe1 binds and ubiquitylates Dvl as part of a negative feedback loop in the Wnt signaling pathway. 1 Hubrecht Institute, Royal Netherlands Academy of Arts and Sciences and University Medical Center Utrecht, Uppsalalaan 8, 3584CT Utrecht, the Netherlands. 2 Institute of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech RepublicAQ2 . 3 Institute of Biophysics, Academy of Sciences of the Czech Republic, Brno, Czech Republic. 4 Cardiff School of Biosciences, Biomedical Sciences Building, Museum Avenue, Cardiff CF10 3AX, UK. 5 Central European Institute of Technology, Masaryk University, Brno, Czech Republic. 6 National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Brno, Czech Republic. 7 Department of Functional Genomics, Interfaculty Institute for Genetics and Functional Genomics, Ernst Moritz Arndt University of Greifswald, Friedrich-Ludwig-Jahn-Str. 15, 17487 Greifswald, Germany. *These authors contributed equally to this work. †Present address: Department of Pathology, University of Cambridge, Tennis Court Road, Cambridge CB2 1QP, UK. ‡Present address: Department of Clinical Research Medical Oncology, University of Bern, Bern, Switzerland. §Corresponding author. E-mail: r.korswagen@hubrecht.eu (H.C.K.); bryja@sci. muni.cz (V.B.) R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 1 MS no: RA2004985/S/CELL BIOL RESULTS Huwe1/EEL-1 is a negative regulator of Wnt/b-catenin signaling in Caenorhabditis elegans and mammalian cells During larval development in Caenorhabditis elegans, EGL-20 (a C. elegans Wnt homo- log)activatesaBAR-1(aC.elegansb-catenin homolog)–dependent Wnt pathway in the QL neuroblasts. Activation of this Wnt pathwayinducesexpressionof thehomeoboxgene mab-5 in QL and migration of the QL descendant (QL.d) cells toward the posterior (12). In the absence of EGL-20 signaling, mab-5isnotexpressedand,asaconsequence, the QL.d migrate in the opposite direction, toward the anterior (F1 Fig. 1A). Using the final position of the QL.d as a sensitive measure of b-catenin–dependent Wnt signaling, we performed an RNA interference (RNAi) screen inC.elegansthattargeted22predictedE2ubiquitin ligases, 173 RING domain E3 ubiquitin ligases, 9 HECT domain E3 ubiquitin ligases, and 34 deubiquitylating enzymes (DUBs) (tableS1andfig.S1A).Toscreenforbothpositive and negative regulators of Wnt signaling, we used a vps-29(tm1320) mutant background, in which EGL-20 secretion is reduced, resulting in a partially penetrant defect in mab-5 expressionandQL.dmigration(13).Interfering with positive regulators of EGL-20 signaling enhances this phenotype, whereas knockdown of negative regulators restores posterior QL.d migration in this mutant background (14). We found that knockdown of the HECT domain–containingubiquitinligaseeel-1suppressed the QL.d migration phenotype of vps-29(tm1320) mutants (Fig. 1B). Furthermore, using a quantitative single-molecule mRNA fluorescence in situ hybridization (FISH) approach (15), we found that loss of eel-1 significantly increased the expression of the EGL-20 target gene mab-5 in Q neuro- blasts(Fig.1C),indicatingthatEEL-1functions as a negative regulator of EGL-20 signaling. To investigate the function of EEL-1 in another Wnt-dependent process, we examined vulva formation, in which Wnt/b-catenin signaling plays a permissive role in preventing fusion of the vulva precursor cells (VPCs) with the surrounding hypodermal syncytium (16). In hypomorphic mutants of the Wnt secretion factor mig-14 (the C. elegans Wntless homolog), a decrease in Wnt signaling leads to a partially penetrant defect in vulva induction (17). RNAi against eel-1 significantly rescued this defect (Fig. 1D), supporting the notion that EEL-1 functions as a general regulator of Wnt/b-catenin signaling in C. elegans. Fig. 1. eel-1isanegativeregulatorofWnt signaling that acts upstream of b-catenin in the Wnt pathway in C. elegans. (A) Dorsal view schematic of the migration of Q neuroblast descendants in wildtype L1 larvae and in animals with impaired or increased EGL-20 signaling. Schematic of the Wnt pathway regulating mab-5 expression in Q neuroblasts. Slashes denote mammalian homologs. (B) Migration of QL.d cells in vps- 29(tm1320) with either systemic or Q neuroblast lineage–specific eel-1 knockdown. Data are means ± SD from three independent experiments with more than 30 worms each; *P = 0.0063, **P = 0.0037, Student’s t test. (C) Quantitative single-molecule mRNA FISH showing the expression of the Wnt target gene mab-5 in QL and QR neuroblasts from wild-type and eel-1(ok1575) null mutants. Data are amalgamated from two experiments; *P = 4.9 × 10−5 , **P = 9.6 × 10−4 , Student’s t test. Representative QL and QR neuroblasts are from wild-type animals: nuclei (blue), mab-5 mRNA (red), and Q cells (outlined). Scale bar, 5 µm. (D) Effect of eel-1 RNAi on the vulva defect, indicated by an arrow. Data are means ± SD from four independent experiments with more than 10 animals each; *P = 0.015. (E and F) Effect of eel-1 RNAi on the response to overexpression of EGL-20 (E) or DN-BAR-1 (F). Data are means±SDfromfourexperimentswithmorethan30animalseach;*P=0.015,Student’sttest.n.s.,notsignificant AQ9. R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 2 MS no: RA2004985/S/CELL BIOL To address whether EEL-1 is required in Wnt-responsiveAQ3 cells, we used a tissue-specific RNAi approach to knock down eel-1 in the Q neuroblast lineage. Similar to systemic RNAi, knockdown of eel-1 in the Q neuroblasts significantly rescued QL.d migration in the vps-29 mutant background (Fig. 1B). In contrast, knockdown of eel-1 in EGL-20–producing cells had no effect (fig. S1B). These results are consistent with a cell-autonomous function of EEL-1 in Wnt-responsive cells. To determine the position of EEL-1 in the Wnt/b-catenin pathway that regulates mab-5 expression in the Q neuroblasts, we performed epistasis analysis with loss-of-function mutations in different Wnt pathway components. RNAi against eel-1 suppressed the QL.d migration phenotype of vps-29 and mig-14 mutants, but had no effect in bar-1 or pop-1/TCF mutants (table S2). Together with the cell-autonomous function of EEL-1 in the Q neuroblast lineage, these results place EEL-1 between the Wnt pathway ligand EGL-20 and bar-1. This conclusion is further supported by experiments in which we overexpressed EGL-20 or constitutively active BAR-1 (DN-BAR-1) (17) to induce ectopic expression of mab-5 in QR and posterior migration of the QR.d. Thus, eel-1 RNAi strongly enhanced the EGL-20–induced posterior migration of the QR.d (Fig. 1E), whereas no effect was observed when DN-BAR-1 was overexpressed (Fig. 1F), consistent with a function of EEL-1 downstream of EGL-20 but upstream of BAR-1. eel-1 RNAi did not restore QL.d migration in egl-20 null mutants (table S2), indicating that loss of EEL-1 is not sufficient to activate the EGL-20 pathway in the absence of Wnt ligand. In parallel to the screen in C. elegans, we knocked down Huwe1, the mammalian ortholog of eel-1, in human embryonic kidney (HEK) 293T cells stably transfected with a Dvl–estrogen receptor fusion protein and a TOP-luciferase TCF reporter (7DF3 cells) (18). We found that knockdown of Huwe1 (F2 Fig. 2B) enhanced the estradiol-dependent activation of the TCF reporter (fig. S2A). This effect was confirmed with nonoverlapping small interfering RNAs (siRNAs) in 7DF3 and U2OS cells (fig. S2, B and C). Next, we activated the Wnt pathway in HEK293T cells with recombinant Wnt3a by expression of constitutively active N-terminally truncated LRP6 (LRP6DN) (19), coexpression of Dvl3 and CK1e (3), or expression of constitutively active b-catenin (20). Knockdown of Huwe1 (Fig. 2B) increased Wnt reporter activity when the pathway was activated by Wnt3a, LRP6DN, or Dvl3 and CK1e, but not when the pathway was activated by expression of constitutively active b-catenin (Fig. 2, C to F). When we overexpressed a Huwe1 fragment containing the HECT domain [hemagglutinin (HA)– tagged DN-Huwe1, encompassing the C-terminal half of the protein from amino acids 2473 to 4374] (21), Wnt reporter activity induced by Dvl1 or LRP6DN was significantly inhibited, whereas reporter activity induced by b-catenin was not changed (Fig. 2, G to I). As an independent readout for Wnt pathway activity, we assessed the mRNA expression of the Wnt target gene Axin2 in cells that expressed constitutively active mutants of b-catenin. Huwe1 overexpression didnotaffect Axin2expressioninthesecells (Fig.2J).Together with the C. elegans data discussed above, these results suggest that Huwe1 functions as an evolutionarily conserved negative regulator of Wnt signaling that acts upstream of b-catenin in the Wnt pathway. Huwe1 binds Dvl We obtained further mechanistic insight into the role of Huwe1 in Wnt signaling when we immunoprecipitated endogenous Dvl2 and Dvl3 from mouse embryonic fibroblasts (MEFs) and analyzed the associated proteins Fig. 2. Huwe1 is a negative regulator of Wnt signaling that acts upstream of b-catenin in the Wnt pathway in mammalian cells. (A) Schematic representation of EEL-1 and the human ortholog Huwe1. (B) Representative Western blot of Huwe1 knockdown efficiency in HEK293T cells. (C to E) TopFlash reporter activity in Huwe1 knockdown HEK293T cells that (C) were treated with recombinant Wnt3a, (D) expressed LRP6DN, or (E) expressed Dvl3 and CK1e. Data are means ± SD from three independent experiments; **P = 0.009, *P = 0.04, ***P = 0.0062 against controls, Student’s t test. (F) TopFlash reporter activity in Huwe1 knockdown cells that expressed constitutively active b-catenin mutants. Data are means ± SEM from seven experiments. (G to I) Effect of HA-DN-Huwe1 overexpression on TopFlash activity in HEK293T cells expressing (G) Dvl1 (*P = 0.0022), (H) LRP6DN (*P = 0.015), or (I) b-catenin. Data are means ± SD from at least three independent experiments. (J) Effect of Huwe1 overexpression on Axin2 mRNA abundance in cells expressing constitutively active b-catenin mutants. Data are means ± SEM from at least three independent experiments. R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 3 MS no: RA2004985/S/CELL BIOL by mass spectrometry (MS) (table S3). We discovered that Huwe1 formed a complex with endogenous Dvl2 and Dvl3. To confirm this result, we performed coimmunoprecipitation experiments with antibodies directed against endogenous Huwe1, Dvl2, and Dvl3 and found that Huwe1 can interact with Dvl2 as well as Dvl3 (F3 Fig. 3A). Treatment of cells with Wnt3a conditioned medium (Wnt3a CM) promoted the interaction of endogenous Huwe1 with endogenous Dvl2 and Dvl3 (Fig. 3A). Similar results were obtained with epitope-tagged versions of Dvl1 and DN-Huwe1 (fig. S3D). This result indicates that Wnt signaling promotes the interaction between Dvl and Huwe1, but that this interaction does not require the N terminus of Huwe1. Dvl proteins form dynamic protein assemblies that are visible as distinct punctae within the cytoplasm (6). Consistent with the interaction between Huwe1 and Dvl, immunofluorescence experiments using epitope-tagged or fluorescently tagged versions of Huwe1 and Dvl2 or Dvl3 showed that Huwe1, which was diffusely localized throughout the cytoplasm when expressed alone, was recruited to Dvl punctae and colocalized with Dvl2 or Dvl3 when Dvl was coexpressed (Fig. 3B and fig. S3E). An E3 ligase–deficient Huwe1(CA) mutant also colocalized with Dvl in cytoplasmic punctae (Fig. 3B), indicating that the interaction between Huwe1 and Dvl is independent of Huwe1 ligase activity. To further characterize the interaction between Dvl and Huwe1, we used truncated forms of Dvl1 and Dvl3 to map the domain of Dvl that is required for binding to HAtagged DN-Huwe1. In both cases, we found that a proline-rich region between the PDZ and DEP domains of Dvl appeared to be necessary and sufficient for the interaction between Huwe1 and Dvl1 and Dvl3 (Fig. 3, C and D, and fig. S3, A to C). One of the earliest events in Wnt pathway activation is the CK1e-mediated phosphorylation of Dvl (22). To address the role of CK1e in the regulation of the Huwe1Dvl interaction, we performed coimmunoprecipitation experiments with FLAG-tagged Dvl3 and HA-tagged DN-Huwe1 in cells overexpressing CK1e. CK1e promoted the Huwe1-Dvl3 interaction, whereas expression of a dominant-negative CK1e or treatment with the chemical CK1e inhibitor D4476 (CK1 inhibitor I) or PF-670462 (CK1 inhibitor II) appeared to reduce this interaction (Fig. 3E). Together, these results indicate that Wnt signaling and phosphorylation Fig. 3. Huwe1 interacts with Dvl. (A) Western blot detecting interaction between endogenous Huwe1, Dvl3,andDvl2inMEFstreatedwith control or Wnt3aCM.TCL,total cell lysate. (B) Localization of FLAGHuwe1 and FLAG-Huwe1(CA) (top row) and colocalization of FLAG-Huwe1 or FLAG-Huwe1 (CA) with Dvl3-eYFP in cytoplasmic Dvl punctae. Data are from three experiments. Scale bars, 20 µm. (C and D) Schematic representation of coimmunoprecipitation (co-IP) of HA-DN-Huwe1 with truncated versions of Dvl1 and Dvl3 presented in fig. S2, A to C. + and − indicate the binding efficiency of the proteins with HA-DN-Huwe1. (E) Coimmunoprecipitation of HA-DN-Huwe1 with FLAG-Dvl3 in HEK293T cells expressing wild-type (wt) or dominant-negative (DN) CK1e or treated with the chemical CK1e inhibitor D4476 (I) or PF-670462 (II). The efficiency of CK1-mediated phosphorylation was monitored using a phospho-Ser643 Dvl3 antibody. Blot is representative of three experiments. R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 4 MS no: RA2004985/S/CELL BIOL of Dvl by CK1e promote the interaction between Dvl and Huwe1. Huwe1 promotes the ubiquitylation of Dvl Huwe1 is a HECT domain E3 ubiquitin ligase that ubiquitylates various substrates and is involved in regulating processes ranging from apoptosis and neuronal differentiation to DNA repair (21, 23–27). To investigate whether Huwe1 ubiquitylates Dvl, we coexpressed FLAG-tagged Dvl1 and His-tagged ubiquitin with DN-Huwe1 or a catalytically inactive DN-Huwe1(CA) mutant (26). Western blotting for FLAG-Dvl1 among His-ubiquitin–modified proteins showed that DN-Huwe1 promoted the ubiquitylation of Dvl1, whereas DN-Huwe1(CA) didnot (F4 Fig.4A). Expressionoffull-length Huwe1 (but not catalytically inactive full-length Huwe1) appeared to induce the ubiquitylation of Dvl3 as well (fig. S4A). Furthermore, deletion of the proline-rich region of Dvl1 appeared to prevent its Huwe1-dependent ubiquitylation (fig. S4B), indicating that binding of Huwe1 to Dvl is required for its ubiquitylation. Toinvestigatewhether Wnt signaling regulates the Huwe1mediated ubiquitylation of Dvl, we treated cells with Wnt3a CM before the ubiquitylation assay. Wnt3a CM induced a clear increase in Huwe1-mediated ubiquitylation of FLAGDvl1 (Fig. 4B), indicating that the Huwe1-dependent ubiquitylation of Dvl is stimulated by Wnt signaling. Because CK1e activity promoted the interaction between Dvl and Huwe1 (Fig. 3E), we investigated the effect of CK1e in the Huwe1-mediated ubiquitylation of Dvl. We found that inhibition of CK1e reduced FLAG-Dvl1 ubiquitylation in cells expressing DN-Huwe1 (Fig. 4C), indicating that CK1e activity promotes the Huwe1-dependent ubiquitylation of Dvl. To gain further insight into the functional consequences of Huwe1-mediated ubiquitylation of Dvl, we used an MSbased approach to identify the lysines of Dvl1 that are ubiquitylated by Huwe1. We expressed His-tagged ubiquitin and FLAG-tagged Dvl1, or His-ubiquitin, FLAG-Dvl1, and DN-Huwe1, in HEK293T cells. We lysed the cells under denaturing conditions, isolated Dvl by immunoprecipitation, and prepared the samples for MS analysis. The detection of ubiquitylated residues in Dvl1 was based on the characteristic ubiquitin-derived diglycine mass increment of ubiquitylated peptides (114.1 daltons per modified lysine residue) after tryptic digestion. In addition to five lysine residues that were ubiquitylated in the control condition, we found five additional lysine residues that were ubiquitylated when DN-Huwe1 was coexpressed (Lys34 , Lys46 , Lys60 , Lys69 , and Lys412 ) (table S4). Most of these lysines were located within the DIX domain of Dvl (Fig. 4D), suggesting that this region is preferentially ubiquitylated by Huwe1. In agreement with this hypothesis, we found that a Dvl1 mutant (DIX7KR, in which the seven lysines in the DIX domain were mutated into arginines) was ubiquitylated with a much lower efficiency by DN-Huwe1 than wild-type Dvl1 or a Dvl1 mutant in which the lysines in the PDZ domain (PDZ3KR) were mutated (Fig. 4E). Mutation of the lysines within the DEP domain (DEP6KR) also decreased the amount of Huwe1-dependent ubiquitylation (Fig. 4E). However, given that the DEP domain is dispensable for Wnt/b-catenin signaling (28, 29), the ubiquitylation of the DIX domain is Fig. 4. Huwe1 promotes Dvl ubiquitylation. (A) Western blotting for FLAG-Dvl1 after pulldown of His-ubiquitin–modified proteins in HEK293T cells transfected with FLAG-Dvl1, His-ubiquitin, and DN-Huwe1 or DN-Huwe1(CA). (B) Western blotting of lysates from cells transfected as in (A), minus the CA Huwe1 mutant, and treated with control medium or Wnt3a CM for 30 or 60 min before lysis. (C) Western blotting of lysates from the cells described in (B) treated with vehicle or PF-670462 (CK1 inhibitor II). (D) Schematic representation of ubiquitylated lysine (K) residues of Dvl1 isolated from controls or cells expressing HA-DN-Huwe1. (E) Western blotting for FLAG-Dvl1 after pull-down of Hisubiquitin–modified proteins in HEK293T cells transfected with HA-DN-Huwe1 and FLAGDvl1 mutants containing lysine-to-arginine mutants in the PDZ (PDZ3KR), DEP (DEP6KR), or DIX (DIX7KR) domains. (F) Schematic representation of ubiquitylated lysine residues of Dvl conjugated to ubiquitin isolated from cells expressing HA-DN-Huwe1 or control cells. (G) Abundance of ubiquitylation (by HA tag) in endogenous Dvl3 immunoprecipitates from HEK293T cells transfected with wild-type HA-ubiquitin or a HA-ubiquitin mutant that only enables the formation of K63-linked or K48-linked polyubiquitin chains (K63 only and K48 only) and treated with Wnt3a or control CM. Western blotting data are from three independent experiments. R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 5 MS no: RA2004985/S/CELL BIOL likely to be most relevant for the function of Huwe1 in the Wnt/b-catenin pathway. The formation of high–molecular weight Dvl species after Huwe1 expression indicates that Huwe1 may promote polyubiquitylation of Dvl. To characterize the ubiquitin chains that are ligated to Dvl, we again used an MS-based approach. This time, we analyzed ubiquitinderived peptides and determined which lysine residues were ubiquitylated (table S5). In His-ubiquitin– andFLAG-Dvl1–transfected samples, we only detected peptides with the diglycine signature corresponding to Lys48 (K48) linkages (Fig. 4F). However, when we expressed DNHuwe1, there were also peptides for K63 and K11 ubiquitin linkages (Fig. 4F), indicating that Huwe1 promotes K63and K11-linked polyubiquitylation of Dvl. To investigate whether Wnt3a induces the formation of K63-linked polyubiquitin chains on Dvl, we performed a ubiquitylation assay with ubiquitin mutants that only enable the formation of K48- or K63-linked polyubiquitin chains, respectively (30). Stimulation with Wnt3a increased the amount of Dvl3 decorated with K63-linked polyubiquitin chains, whereas there was no change in the amount of K48-modified Dvl3 (Fig. 4G). Given the similarity in the ubiquitin modifications induced by Wnt3a and Huwe1, we propose that Huwe1 mediates the Wnt3a-induced K63-linked polyubiquitylation of Dvl. Huwe1 does not target Dvl for degradation but influences Dvl multimerization We set out to elucidate the mechanism by which the Huwe1dependent ubiquitylation of Dvl inhibits Wnt signaling. First, we investigated whether Huwe1 targets Dvl for degradation. We expressed FLAG-Dvl1 in HEK293T cells together with HA-DN-Huwe1 or KLHL12, a protein that recruits the Cullin-3 ubiquitin ligase toward Dvl and targets Dvl for proteasomal degradation (7). We inhibited protein translation by incubating the cells with cycloheximide and determined FLAG-Dvl1 protein expression and stability by Western blotting. We found no reduction in the amount and stability of FLAG-Dvl1 when HA-DN-Huwe1 was expressed (F5 Fig. 5A), but we observed a clear reduction in the amount and stability of FLAG-Dvl when KLHL12 was coexpressed. Additionally, we did not observe increased abundance of endogenous Dvl when we knocked down Huwe1 (Fig. 2B). These observations agree with our finding that DN-Huwe1 promotes the formation of K63-linked polyubiquitin chains on Dvl, a modification that in general does not target the substrate protein for proteasomal degradation (10). The formation of DIX domain–dependent multimeric aggregates is essential for Dvl signaling activity (5, 31). Because Huwe1 promoted the ubiquitylation of the DIX domain (Fig. 4, D and E), we investigated whether Huwe1 influences the ability of Dvl to multimerize. To this end, we examined the interaction between different Dvl isoforms by coimmunoprecipitation. Knockdown of Huwe1 increased the coimmunoprecipitation between endogenous Dvl2 and Dvl3 (Fig. 5B), whereas overexpression of HA-tagged DNHuwe1 decreased the coimmunoprecipitation between FLAGtagged Dvl1 and HA-tagged Dvl2 (Fig. 5D, lanes 1 and 2). Together, these results indicate that Huwe1 inhibited Dvl multimerization. This was supported by the observation that overexpression of HA-tagged DN-Huwe1 reduced the interaction between cyan fluorescent protein (CFP)–conjugated Dvl3 and yellow fluorescent protein (YFP)–conjugated Dvl3 in a fluorescence resonance energy transfer (FRET)–based assay (Fig. 5C and fig. S5A). To investigate whether Huwe1 inhibited Dvl multimerization through ubiquitylation of the DIX domain, we tested the coimmunoprecipitation between HA-tagged Dvl2 and the FLAG-tagged Dvl1-DIX7KR mutant (8). This mutant activated the Wnt pathway (assessed by a TopFlash reporter) and localized to punctae in the cytoplasm (fig. S5, B and C), but was only weakly ubiquitylated by Huwe1 (Fig. 4E). We found that overexpression of Fig. 5. Huwe1 does not target Dvl for degradation but inhibits Dvl polymerization.(A) HEK293T cells transfected with FLAG-Dvl1 and control vector, KLHL12, or HA-DN-Huwe1 expression constructs were treated with cycloheximide (CHX) for the indicated time before lysis and FLAG-Dvl1 detection. (B) Knockdown of Huwe1 increases coimmunoprecipitation between endogenous Dvl2 and Dvl3. OVCAR-5 cells were transfected with control or Huwe1 siRNA, and endogenous Dvl2 was immunoprecipitated. Dvl3present inthecomplex was visualized by Westernblotting. (C) FRETefficiency between Dvl3-YFP and Dvl3-CFP in HEK293T cells transfected with control or HA-DN-Huwe1 expression constructs. For typical FRET images, see fig. S4A. (D) Dvl1 was immunoprecipitated with anti-FLAG antibody, and complex composition was determined by Western blotting. (E) Structural model of two dimerized DIX domains of Dvl1. Ubiquitylated lysine (K) residues are indicated in yellow. Mutations known to prevent DIX domain polymerization (M1, M2, and M4) are indicated in red; b sheets involved in the head (b2)–tail (b4) interaction are labeled. (F) Model of Huwe1 function in Wnt/b-catenin signaling, described in the text. Yellow circles, phosphate; orange circles, ubiquitin; blue hexagons, Wnt. R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 6 MS no: RA2004985/S/CELL BIOL HA-DN-Huwe1 appeared not to decrease the coimmunoprecipitation between FLAG-tagged Dvl1-DIX7KR and HA-tagged Dvl2 (Fig. 5D, lanes 6 and 7), demonstrating that the inhibitory effect of Huwe1 on Dvl multimerization is mediated through ubiquitylation of the DIX domain. Templatebased homology modeling of the human Dvl3 DIX domain showed that a number of lysines in the DIX domain that are ubiquitylated by Huwe1 are structurally close, or identical to, residues that have previously been identified as being essential for DIX domain–dependent polymerization (Fig. 5E). Specifically, Lys34 (corresponding to Lys44 in Dvl2) neighbors a fully conserved phenylalanine, which is mutated in the M1 aggregation mutant (Dvl2-F43S) (5). The M1 mutation directly disrupts the “head” side of the polymerization interface. Second, Lys60 (corresponding to Lys68 in Dvl2) is part of the b4 sheet forming the tail side of the DIXDIX polymerization interface. Moreover, mutation in a corresponding residue in Dvl2 (K68A, corresponding to the aggregation mutant M2) (5) leads to a clear multimerization defect of Dvl2. We therefore propose that ubiquitylation of these residues inhibits DIX domain polymerization through steric effects. DISCUSSION Here, we identified the HECT domain–containing E3 ubiquitin ligase Huwe1 as an evolutionarily conserved inhibitor of the Wnt/b-catenin pathway. Our results from epistasis experiments in C. elegans and biochemical studies in mammalian cells suggest that Huwe1 acts by inhibiting Dvl signaling activity. The cytoplasmic protein Dvl functions at a key position in the Wnt/bcatenin pathway, interacting with Fz and Axin to facilitate the formation of signalosomes at the plasma membrane (32). The signaling activity of Dvl is promoted by the kinase CK1e, which in response to Wnt signaling phosphorylates Dvl to induce binding of Dvl to Axin (3, 4, 21, 31). We have previously shown that in addition to this activating function, CK1e also triggers a negative feedback loop that inhibits Wnt signaling, but the mechanism of this inhibition was not determined (3). Our results demonstrate that CK1e induces the binding of Huwe1 to Dvl and provides a molecular mechanism for this inhibition (Fig. 5F). We found that Huwe1 polyubiquitylated lysine residues within the DIX domain of Dvl. Consistent with our finding that Huwe1 mainly promoted K63- and K11-linked polyubiquitylation of Dvl, Huwe1 did not target Dvl for proteasomal degradation. Instead, we found that Huwe1 inhibited the multimerization of Dvl. As the formation of multimeric complexes is essential for Dvl activity (5, 31), our results are consistent with a model in which Huwe1 inhibits Wnt/b-catenin signaling by preventing Dvl multimerization (Fig. 5F). Two DUBs—Cyld and USP14—were recently identified that remove K63-linked polyubiquitin chains from Dvl. Cyld removes polyubiquitin (including K63-linked polyubiquitin) chains from the DIX domain of Dvl (8). Inhibition of Cyld function and the resulting hyperubiquitylation of Dvl activate Wnt/b-catenin signaling in mammalian cells (8). These results support the notion that K63-linked polyubiquitylation is an important regulatory mechanism of Dvl activity, but do not explain why Huwe1 and Cyld both act as inhibitors of Wnt/b-catenin signaling. Cyld and Huwe1 may control the ubiquitylation of different lysine residues within the DIX domain to either promote signaling activity or inhibit signaling by preventing Dvl multimerization. Alternatively, a tight balance in the dynamic addition and removal of K63-linked polyubiquitin chains may be required for the regulation of Dvl activity. USP14, which also removes K63-linked polyubiquitin chains from Dvl (34), was identified as a positive regulator of the Wnt pathway. This observation indicates that it could act as a counterpart to the Huwe1-dependent negative regulation of Wnt signaling. Whether USP14 removes the polyubiquitin chains that are generated by Huwe1, however, remains to be established. Our study, as well as the work of others, illustrates the importance of ubiquitin-mediated regulation of Dvl signaling in the Wnt pathway. Several E3 ubiquitin ligase systems modify Dvl. Thus, Dvl is targeted for proteasomal degradation by NEDL1, an E3 ligase associated with amyotrophic lateral sclerosis (35), by KLHL12, an adaptor protein linking Dvl to the Cullin-3 E3 ubiquitin ligase (7), and by Itch, which is a HECT domain E3 ligase (9). Thus far, eel-1 in worms or Huwe1 in mammals represents the only Dvl ubiquitin ligase or DUB with a role in Wnt signaling that is evolutionarily conserved in both vertebrates and invertebrates. This indicates that Huwe1 may represent an ancestral mechanism of Dvl regulation. Last but not least, Huwe1 was recently identified as an important mediator of Wnt pathway–dependent intestinal tumorigenesis in a mouse transposon insertional mutagenesis screen (36). Furthermore, sequence information in the COSMIC database (37) shows that Huwe1 is frequently mutated in colon cancer. Our results on the function of Huwe1 as an inhibitor of Wnt/b-catenin signaling—the deregulation of which is the primary cause of intestinal tumorigenesis (20)—may provide an important mechanistic explanation for a role of Huwe1 in intestinal cancer. MATERIALS AND METHODS C. elegans strains and culture C. elegans strains were cultured at 20°C under standard conditions as described (38). The mutants and transgenes used were vps-29(tm1320), egl-20(hu105), mig-14(ga62), mig-14(mu71), mig-5(rh147), bar-1(ga80), pop-1(hu9), eel-1(ok1575), muIs32[Pmec-7::gfp], muIs35[Pmec- 7::gfp], huIs7[Phs::DNbar-1], muIs53[Phs::egl-20], heIs63[Pwrt-2::H2b::gfp; Pwrt2::PH::gfp], huEx273[Pegl-17::eel-1(RNAi)], and huEx274[Pegl- 20::eel-1(RNAi)]. C. elegans RNAi screen A list of E2 ubiquitin–conjugating genes, DUBs, RING domain–containing genes, and HECT domain–containing genes was obtained from Kipreos (39) (table S1). We isolated RNAi clones targeting these genes from the Vidal or Ahringer RNAi libraries (40, 41). RNAi clones were grown overnight in LB supplemented with ampicillin (50 mg/ml). Bacterial cultures were grown for 3 days on NGM plates supplemented with ampicillin, tetracycline, fungizone, and isopropyl-b-D-thiogalactopyranoside. Two vps-29(tm1320);muIs32 L4 larvae were placed on the bacterial lawn and allowed to give rise to progeny. The position of the QL.d cells relative to the vulva was scored when the progeny reached the young adult stage (42). If RNAi clones modulated the vps-29 QL.d migration phenotype to >80% or <11%, the RNAi experiment was repeated twice. RNAi clones that consistently enhanced or suppressed the vps-29 phenotype were considered a hit. The RNAi clones were confirmed by sequencing. C. elegans tissue-specific RNAi and heat shock experiments Tissue-specific RNAi was performed as described (43). egl-17 and egl-20 promoter sequences and a 1.1-kb fragment spanning exon 13 of eel-1 were amplified from C. elegans genomic DNA. The eel-1 fragment was fused to the promoter fragments in sense and antisense orientations using a polymerase chain reaction (PCR) approach. The PCR products were injected in vps-29(tm1320); muIs32 at 2.5 ng/µl with Pmyo2:mCherry injection marker (7 ng/µl) and pBluescript DNA (200 ng/µl), resulting in transgenic lines huEx273[Pegl-17::eel-1(RNAi)] and huEx274[Pegl-20::eel-1(RNAi)]. R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 7 MS no: RA2004985/S/CELL BIOL Embryos isolated by bleaching from huIs7 or muIs53 animals grown on eel-1 or control double-stranded RNA–expressing bacteria were allowed to hatch overnight in M9 buffer. About 100 L1 larvae were heatshocked in 100 µl of M9 at 33°C for 5 min, and the position of the Q cell descendants AVM and PVM was determined when the animals reached the young adult stage. Single-molecule mRNA FISH Single-molecule mRNA FISH was performed as described (15). Imaging was performed with a Leica DM6000 microscope equipped with a Leica DFC360FX camera, 100× objective, and Tx2 filter cube. Images were acquired at 1024 × 1024 resolution and 2 × 2 binned before analysis using ImageJ. mab-5 mRNA spots were manually counted in the Q neuroblasts using heIs63 as a marker to outline the cells. Cell culture, plasmids, and transfections MEFs and HEK293T cells were propagated in Dulbecco’s modified Eagle’s medium with 10% fetal bovine serum (FBS), 5% L-glutamine, and 5% penicillin/streptomycin. OVCAR-5 cells were grown in RPMI medium with 10% FBS, 5% L-glutamine, and 5%penicillin/streptomycin.Cellswere seeded in 24-well plates on coverslips for FRETand immunocytochemistry and 10-cm dishes for immunoprecipitation. Cells were transfected with 2 µl of polyethylenimine per microgram of DNA. Forty-eight hours after transfection, cells were harvested and processed. The following previously described expression constructs were used: Dvl2-Myc (44), FLAG-hDvl3 and deletions (7), hCK1e (wild-type and P3 dominant-negativeAQ4 ) (45), FLAGmDvl1 deletions and His-ubiquitin lysine mutants (8), HA-ubiquitin mutants (30), HA-Dvl2(31), HA-DN-Huwe1(21),FLAG-Huwe1 andFLAG-Huwe1 (CA)(27),b-catenin,S33b-catenin,D43b-catenin(20),Lrp6DN(46),Super8X TopFlash (47), and VSV-KLHL12 (7). Dvl3-EYFP, Dvl3-ECFP, and FLAG-Dvl1/Dvl3 Pro-rich were generated by Gateway cloning. Cells were stimulated with mouse rWnt3a (R&D Systems) for 3 hours if not stated otherwise. Control stimulationswere donewith0.1% bovineserumalbumin (BSA) in phosphate-buffered saline (PBS). CK1e inhibition was performed with D4476 (Calbiochem) and/or PF-670462 (Tocris) dissolved in dimethyl sulfoxide or with 1 ml of Lipofectamine 2000 (Invitrogen) per well to increase cell penetration. RNA interference HEK 293T cells were transfected with Huwe1 endoribonuclease-prepared siRNA (esiRNA) using Lipofectamine 2000. Huwe1 esiRNA (300 ng per well) was mixed with Opti-MEM (Gibco) and 0.5 µl of Lipofectamine 2000 and incubated at room temperature for 30 min, after which the mixture was added to trypsinized cells in 24-well plates. The medium was changed after 12 hours. Dual luciferase assay HEK293T cells were seeded in 24-well plates. After 24 hours, cells were transfected with 0.2 µg of Super8X TopFlash construct and 0.2 mg of Renilla luciferase construct per well and the indicated expression plasmids. Cells were lysed 24 hours later, and luciferase activity was measured with the Dual-Luciferase Reporter Assay System (Promega) on an MLX luminometer (Dynex Technologies) and normalized by Renilla luciferase measurements. Data are shown as means ± SD of at least three independent experiments. Quantitative reverse transcription PCR RNA was isolated using TRIzol (Ambion 15596-026) according to the manufacturer’s instructions. Reverse transcription was performed with M-MuLV Reverse Transcriptase (Thermo Fisher Scientific), oligo(dT)18 primers (Thermo Fisher Scientific), and 1 µg of RNA for first-strand comple- mentaryDNAsynthesis.SampleswererunastriplicatesonaLightCycler480 II (Roche). b-Actin was used as an internal control. Primers used were hu- manACTB(b-actin;forward:5′-GATTCCTATGTGGTCGACGAG-3′,reverse: AGGGCTGGGGTGTTGAAGGTCTC) and human AXIN2 (forward: 5′TCCATACCGGAGGATGCTGA-3′, reverse: 5′-TTCATACATCGG- GAGCACCG-3′). Antibodies, immunoprecipitation, and Western blotting Cells were lysed 24 to 48 hours after transfection in NP-40 lysis buffer containing 50 mM tris (pH 7.4), 150 mM NaCl, 1 mM EDTA, 0.5% NP-40, and protease inhibitors (Roche) for 20 min at 4°C and centrifuged at 13,000 rpm for 30 min at 4°C. Supernatant was incubated with the antibodies against mouse monoclonal Dvl3 (Santa Cruz Biotechnology), mouse monoclonal Dvl2 (Santa Cruz Biotechnology), mouse monoclonal FLAG M2 (Sigma), rabbit polyclonal HA (Abcam), or rabbit polyclonal Huwe1 (Bethyl Laboratories) for 30 min on ice followed by incubation with protein G–Sepharose beads AQ5(GE Healthcare) overnight at 4°C. Samples were washed five times in NP-40 lysis buffer and analyzed by Western blot with antibodies against mouse monoclonal Myc, goat polyclonal actin, or goat polyclonal Ck1e (all from Santa Cruz Biotechnology), rabbit polyclonal phospho-Ser643 Dvl3 (Moravian Biotechnology), or mouse monoclonal a-tubulin (Sigma), followed by horseradish peroxidase–conjugated antibodies against mouse (GE Healthcare) or rabbit (Sigma) immunoglobulin G (IgG). Immunocytochemistry HEK293 cells were seeded onto gelatin-coated coverslips in 24-well plates, and transfection was carried out after 24 hours. Transfected cells were fixed after 24 hours in 4% paraformaldehyde for 15 min and blocked with PBTA (3% BSA, 0.25% Triton, 0.01% NaN3) for 1 hour at room temperature, followed by incubation in primary antibodies overnight at 4°C. Then, cells were washed in PBS; incubated with secondary antibodies conjugated to Alexa 488, Alexa 568, or Alexa 594 (Invitrogen) for 1 hour at room temperature; washed twice with PBS; and mounted in 4′,6-diamidino-2-phenylindole. Ubiquitylation assay HEK293T cells were transfected with expression constructs for FLAGDvl1, His-ubiquitin, and HA-DN-Huwe1; washed 48 hours later in PBS supplemented with 10 mM N-ethylmaleimide (NEM); lysed in denaturing guanidium buffer [10 mM tris (pH 8), 0.1 M Na2HPO4, 0.1 M NaH2PO4, 20 mM imidazole, 6 M guanidium-HCl, and 10 mM b-mercaptoethanol]; and sonicated. Ubiquitylation assays were performed as described using His pull-down (8) or a denaturing immunoprecipitation protocol. Samples for denaturing immunoprecipitation were lysed in 0.5% SDS, boiled for 8 min, and sonicated. Then, 1 ml of NP-40 lysis buffer supplemented with protease inhibitors, phosphatase inhibitors, and NEM was added to each sample. Samples were centrifuged at 16.1g for 30 min. Sixty microliters of supernatant was reserved for total cell lysate assays, and the rest was used for incubation with the corresponding antibody. Samples were processed and analyzed as in standard immunoprecipitation assays. For His pulldown, the lysate was centrifuged (10 min at 14,000 rpm), and His-tagged proteins were isolated from the supernatant with Ni-NTA beads (Qiagen) for 4 hours at room temperature. The beads were washed in guanidium buffer and three times in ureum buffer before elution of His-tagged proteins in 1× Laemmli sample buffer supplemented with 10 mM tris (pH 7), 0.1 M Na2HPO4, 0.1 M NaH2PO4, 8 M urea, and 200 mM imidazole. MS analysis Samples from the ubiquitylation assays were run on SDS–polyacrylamide gel electrophoresis to separate the ubiquitylated complexes. Gels were R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 8 MS no: RA2004985/S/CELL BIOL stained with 0.25% Coomassie blue stain (R250). Corresponding onedimensional (1D) bands were excised and destained, followed by reduction and alkylation. Gel pieces were then subjected to digestion by trypsin for 2 hours at 40°C. Immunoprecipitated Dvl was digested directly in solution. Liquid chromatography (LC)–MS/MS analyses of peptide mixtures were done using the RSLCnano system connected to an Orbitrap Elite mass spectrometer (Thermo Fisher Scientific). Before LC separation, tryptic digests were concentrated and desalted using a trapping column (100 mm × 30 mm) filled with 3.5-mm X-Bridge BEH 130 C18 sorbent (Waters). After washing with 0.1% formic acid (FA)AQ6 , the peptides were eluted (flow 300 nl/min) from the trapping column onto a Acclaim Pepmap100 C18 analytical column (2-µm particles, 75 mm × 250 mm; Thermo Fisher Scientific) by the following gradient program: mobile phase A—0.1% FA in water; mobile phase B—ACN/methanol/2,2,2-trifluoroethanol (6:3:1, v/v/v) containing 0.1% FA; the gradient elution started at 2% of mobile phase B and increased from 2 to 45% during the first 90 min (11% in the 30th, 25% in the 60th, and 45% in the 90th min), then increased linearly to 95% of mobile phase B in the next 5 min and remained at this state for the next 15 min. MS data were acquired in a data-dependent strategy with dynamic precursor exclusion selecting up to the top 20 of precursors on the basis of precursor abundance in the survey scan [350 to 1700 mass/charge ratio (m/z), resolution 120,000]. Low-resolution collision-induced dissociation (CID) MS/MS spectra were acquired in rapid CID scan mode. FRET analysis Cells were seeded on gelatin-coated coverslips in 24-well plates and transfected after 24 hours with 0.1 µg of Dvl3-ECFP, 0.1 µg of Dvl3EYFP, and 0.3 µg of HA-DN-Huwe1 using polyethylenimine. Plasmids containing fused EYFP and ECFP together were used as additional controls. The next day, cells were fixed in 4% paraformaldehyde for 15 min and blocked with PBTA (3% BSA, 0.25% Triton, 0.01% NaN3) for 1 hour at room temperature, followed by incubation in antibodies against HA overnight at 4°C. Cells were then washed in 0.1 M Pipes (pH 6.9) and incubated with Alexa 594–conjugated antibody against rabbit IgG (Invitrogen) at room temperature for 1 hour. Cells were then washed twice with Pipes solution, once with 2 mM MgCl2, and once with EGTA, followed by mounting. FRET between Dvl3-ECFP (donor) and Dvl3-EYFP (acceptor) was measured with a Leica confocal microscope. CFP, YFP, and Alexa 594 were excited using 455-, 514-, and 590-nm lasers, respectively. Region of interest was set on Dvl punctae. Dvl3-YFP punctae were irreversibly bleached for 15 to 30 s. Ratios of donor intensities before and after photobleaching were calculated to obtain FRETefficiencies. To avoid heterogeneity in donor and acceptor expression amounts, we performed experiments three times measuring FRET from 15 to 20 punctae per cell from 4 to 5 cells. Image quantification Western blots from three independent experiments were scanned with an Epson perfection 4990 photo scanner. Images were digitally inverted, and band intensities were measured using ImageJ software. Intensities are represented as arbitrary units in each experiment. 3D structure modeling The 3D model of hDvl3 DIX domain was generated by template-based homology modeling using the program PHYRE2 (48), and identified ubiquitin-modified residues and residues corresponding to M1, M2, and M4 DIX mutants (5) were mapped and visualized to this 3D model using the CHIMERA program (49). SUPPLEMENTARY MATERIALS www.sciencesignaling.org/cgi/content/full/7/317/ra26/DC1 Fig. S1. eel-1 is a negative regulator of Wnt signaling in C. elegans. Fig. S2. eel-1/Huwe1 is a negative regulator of Wnt signaling that acts upstream of b-catenin in the Wnt pathway. Fig. S3. Huwe1 interacts with Dvl and promotes Dvl ubiquitylation. Fig. S4. Huwe1 promotes Dvl ubiquitylation. Fig. S5. Huwe1 inhibits Dvl multimerization. Table S1. List of DUBs, E2 enzymes, and E3 ubiquitin ligases that were screened for enhancement or suppression of the vps-29(tm1320)–induced QL.d migration phenotype. Table S2. Effect of eel-1(RNAi) on the EGL-20/Wnt–dependent migration of the QL.d in various Wnt pathway mutants. Table S3. MS analysis of binding partners of Dvl immunoprecipitated from MEFs. Table S4. MS analysis of ubiquitylated lysine residues of Dvl1. Table S5. MS analysis of ubiquitin chain linkage ligated to Dvl1. REFERENCES AND NOTES 1. B. T. MacDonald, K. Tamai, X. He, Wnt/b-catenin signaling: Components, mechanisms, and diseases. Dev. Cell 17, 9–26 (2009). 2. J. Bilic, Y. L. Huang, G. Davidson, T. Zimmermann, C. M. Cruciat, M. Bienz, C. Niehrs, Wnt induces LRP6 signalosomes and promotes Dishevelled-dependent LRP6 phosphorylation. Science 316, 1619–1622 (2007) AQ7. 3. O. Bernatik, R. S. Ganji, J. P. Dijksterhuis, P. Konik, I. Cervenka, T. Polonio, P. Krejci, G. Schulte, V. Bryja, Sequential activation and inactivation of Dishevelled in the Wnt/bcatenin pathway by casein kinases. J. Biol. Chem. 286, 10396–10410 (2011). 4. C. M. Cruciat, C. Dolde, R. E. de Groot, B. Ohkawara, C. Reinhard, H. C. Korswagen, C. Niehrs, RNA helicase DDX3 is a regulatory subunit of casein kinase 1 in Wnt–bcatenin signaling. Science 339, 1436–1441 (2013). 5. T. Schwarz-Romond, M. Fiedler, N. Shibata, P. J. Butler, A. Kikuchi, Y. Higuchi, M. Bienz, The DIX domain of Dishevelled confers Wnt signaling by dynamic polymerization. Nat. Struct. Mol. Biol. 14, 484–492 (2007). 6. T. Schwarz-Romond, C. Merrifield, B. J. Nichols, M. Bienz, The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J. Cell Sci. 118, 5269–5277 (2005). 7. S. Angers, C. J. Thorpe, T. L. Biechele, S. J. Goldenberg, N. Zheng, M. J. MacCoss, R. T. Moon, The KLHL12–Cullin-3 ubiquitin ligase negatively regulates the Wnt–b-catenin pathway by targeting Dishevelled for degradation. Nat. Cell Biol. 8, 348–357 (2006). 8. D. V. F. Tauriello, A. Haegebarth, I. Kuper, M. J. Edelmann, M. Henraat, M. R. Canninga-van Dijk, B. M. Kessler, H. Clevers, M. M. Maurice, Loss of the tumor suppressor CYLD enhances Wnt/b-catenin signaling through K63-linked ubiquitination of Dvl. Mol. Cell 37, 607–619 (2010). 9. W. Wei, M. Li, J. Wang, F. Nie, L. Li, The E3 ubiquitin ligase ITCH negatively regulates canonical Wnt signaling by targeting Dishevelled protein. Mol. Cell. Biol. 32, 3903–3912 (2012). 10. Z. J. Chen, L. J. Sun, Nonproteolytic functions of ubiquitin in cell signaling. Mol. Cell 33, 275–286 (2009). 11. C. Gao, Y. G. Chen, Dishevelled: The hub of Wnt signaling. Cell. Signal. 22, 717–727 (2010). 12. M. Silhankova, H. C. Korswagen, Migration of neuronal cells along the anterior– posterior body axis of C. elegans: Wnts are in control. Curr. Opin. Genet. Dev. 17, 320–325 (2007). 13. M. J. Lorenowicz, M. Macurkova, M. Harterink, T. C. Middelkoop, R. de Groot, M. C. Betist, H. C. Korswagen, Inhibition of late endosomal maturation restores Wnt secretion in Caenorhabditis elegans vps-29 retromer mutants. Cell. Signal. 26, 19–31 (2014). 14. M. Harterink, F. Port, M. J. Lorenowicz, I. J. McGough, M. Silhankova, M. C. Betist, J. R. T. van Weering, R. G. H. P. van Heesbeen, T. C. Middelkoop, K. Basler, P. J. Cullen, H. C. Korswagen, A SNX3-dependent retromer pathway mediates retrograde transport of the Wnt sorting receptor Wntless and is required for Wnt secretion. Nat. Cell Biol. 13, 914–923 (2011). 15. A. Raj, P. Van Den Bogaard, S. A. Rifkin, A. Van Oudenaarden, S. Tyagi, Imaging individual mRNA molecules using multiple singly labeled probes. Nat. Methods 5, 877–879 (2008). 16. T. R. Myers, I. Greenwald, Wnt signal from multiple tissues and lin-3/EGF signal from the gonad maintain vulval precursor cell competence in Caenorhabditis elegans. Proc. Natl. Acad. Sci. U.S.A. 104, 20368–20373 (2007). 17. J. E. Gleason, H. C. Korswagen, D. M. Eisenmann, Activation of Wnt signaling bypasses the requirement for RTK/Ras signaling during C. elegans vulval induction. Genes Dev. 16, 1281–1290 (2002). 18. K. Ewan, B. Pajak, M. Stubbs, H. Todd, O. Barbeau, C. Quevedo, H. Botfield, R. Young, R. Ruddle, L. Samuel, A. Battersby, F. Raynaud, N. Allen, S. Wilson, B. Latinkic, P. Workman, E. McDonald, J. Blagg, W. Aherne, T. Dale, A useful approach to identify novel small-molecule inhibitors of Wnt-dependent transcription. Cancer Res. 70, 5963–5973 (2010). R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 9 MS no: RA2004985/S/CELL BIOL 19. K. Brennan, J. M. Gonzalez-Sancho, L. A. Castelo-Soccio, L. R. Howe, A. M. C. Brown, Truncated mutants of the putative Wnt receptor LRP6/Arrow can stabilize b-catenin independently of Frizzled proteins. Oncogene 23, 4873–4884 (2004). 20. P. J. Morin, A. B. Sparks, V. Korinek, N. Barker, H. Clevers, B. Vogelstein, K. W. Kinzler, Activation of b-catenin-Tcf signaling in colon cancer by mutations in b-catenin or APC. Science 275, 1787–1790 (1997). 21. S. Adhikary, F. Marinoni, A. Hock, E. Hulleman, N. Popov, R. Beier, S. Bernard, M. Quarto, M. Capra, S. Goettig, U. Kogel, M. Scheffner, K. Helin, M. Eilers, The ubiquitin ligase HectH9 regulates transcriptional activation by Myc and is essential for tumor cell proliferation. Cell 123, 409–421 (2005). 22. W. Swiatek, I. C. Tsai, L. Klimowski, A. Pepler, J. Barnette, H. J. Yost, D. M. Virshup, Regulation of casein kinase Ie activity by Wnt signaling. J. Biol. Chem. 279, 13011–13017 (2004). 23. D. Chen, N. Kon, M. Li, W. Zhang, J. Qin, W. Gu, ARF-BP1/Mule is a critical mediator of the ARF tumor suppressor. Cell 121, 1071–1083 (2005). 24. A. J. Ross, M. Li, B. Yu, M. X. Gao, W. B. Derry, The EEL-1 ubiquitin ligase promotes DNA damage-induced germ cell apoptosis in C. elegans. Cell Death Differ. 18, 1140–1149 (2011). 25. X. Zhao, D. D’ Arca, W. K. Lim, M. Brahmachary, M. S. Carro, T. Ludwig, C. C. Cardo, F. Guillemot, K. Aldape, A. Califano, A. Iavarone, A. Lasorella, The N-Myc-DLL3 cascade is suppressed by the ubiquitin ligase Huwe1 to inhibit proliferation and promote neurogenesis in the developing brain. Dev. Cell 17, 210–221 (2009). 26. X. Zhao, J. I. T. Heng, D. Guardavaccaro, R. Jiang, M. Pagano, F. Guillemot, A. Iavarone, A. Lasorella, The HECT-domain ubiquitin ligase Huwe1 controls neural differentiation and proliferation by destabilizing the N-Myc oncoprotein. Nat. Cell Biol. 10, 643–653 (2008). 27. Q. Zhong, W. Gao, F. Du, X. Wang, Mule/ARF-BP1, a BH3-only E3 ubiquitin ligase, catalyzes the polyubiquitination of Mcl-1 and regulates apoptosis. Cell 121, 1085–1095 (2005). 28. J. D. Axelrod, J. R. Miller, J. M. Shulman, R. T. Moon, N. Perrimon, Differential recruitment of Dishevelled provides signaling specificity in the planar cell polarity and Wingless signaling pathways. Genes Dev. 12, 2610–2622 (1998). 29. M. Boutros, N. Paricio, D. I. Strutt, M. Mlodzik, Dishevelled activates JNK and discriminates between JNK pathways in planar polarity and wingless signaling. Cell 94, 109–118 (1998). 30. K. L. Lim, K. C. M. Chew, J. M. M. Tan, C. Wang, K. K. K. Chung, Y. Zhang, Y. Tanaka, W. Smith, S. Engelender, C. a Ross, V. L. Dawson, T. M. Dawson, Parkin mediates nonclassical, proteasomal-independent ubiquitination of synphilin-1: Implications for Lewy body formation. J. Neurosci. 25, 2002–2009 (2005). 31. T. Schwarz-Romond, C. Metcalfe, M. Bienz, Dynamic recruitment of axin by Dishevelled protein assemblies. J. Cell Sci. 120, 2402–2412 (2007). 32. J. Bilic, Y. L. Huang, G. Davidson, T. Zimmermann, C. M. Cruciat, M. Bienz, C. Niehrs, Wnt induces LRP6 signalosomes and promotes Dishevelled-dependent LRP6 phosphorylation. Science 316, 1619–1622 (2007). 33. V. Bryja, G. Schulte, N. Rawal, A. Grahn, E. Arenas, Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J. Cell Sci. 120, 586–595 (2007). 34. H. Jung, B. G. Kim, W. H. Han, J. H. Lee, J. Y. Cho, W. S. Park, M. M. Maurice, J. K. Han, M. J. Lee, D. Finley, E. H. Jho, Deubiquitination of Dishevelled by Usp14 is required for Wnt signaling. Oncogenesis 2, e64 (2013). 35. K. Miyazaki, T. Fujita, T. Ozaki, C. Kato, Y. Kurose, M. Sakamoto, S. Kato, T. Goto, Y. Itoyama, M. Aoki, A. Nakagawara, NEDL1, a novel ubiquitin-protein isopeptide ligase for Dishevelled-1, targets mutant superoxide dismutase-1. J. Biol. Chem. 279, 11327–11335 (2004). 36. H.N.March,A.G.Rust,N.A.Wright,J.Hoeve,J.DeRidder,M.Eldridge,L.VanDerWeyden, A. Berns, J. Gadiot, A. Uren, R. Kemp, M. J. Arends, L. F. A. Wessels, D. J. Winton, D. J. Adams, Insertional mutagenesis identifies multiple networks of cooperating genes driving intestinal tumorigenesis. Nat. Genet. 43, 1202–1209 (2011). 37. S. A. Forbes, N. Bindal, S. Bamford, C. Cole, C. Y. Kok, D. Beare, M. Jia, R. Shepherd, K. Leung, A. Menzies, J. W. Teague, P. J. Campbell, M. R. Stratton, P. A. Futreal, COSMIC: Mining complete cancer genomes in the Catalogue of Somatic Mutations in Cancer. Nucleic Acids Res. 39, D945–D950 (2011). 38. J. Lewis, J. Fleming, Basic culture methods. Methods Cell Biol. 48, 3–29 (1995). 39. E. T. Kipreos, Ubiquitin-mediated pathways in C. elegans. WormBook 1–24 (2005). 40. J. F. Rual, J. Ceron, J. Koreth, T. Hao, A. S. Nicot, T. Hirozane-Kishikawa, J. Vandenhaute, S. H. Orkin, D. E. Hill, S. van den Heuvel, M. Vidal, Toward improving Caenorhabditis elegans phenome mapping with an ORFeome-based RNAi library. Genome Res. 14, 2162–2168 (2004). 41. R. S. Kamath, A. G. Fraser, Y. Dong, G. Poulin, R. Durbin, M. Gotta, A. Kanapin, N. Le Bot, S. Moreno, M. Sohrmann, D. P. Welchman, P. Zipperlen, J. Ahringer, Systematic functional analysis of the Caenorhabditis elegans genome using RNAi. Nature 421, 231–237 (2003). 42. Q. Ch’ng, L. Williams, Y. S. Lie, M. Sym, J. Whangbo, C. Kenyon, Identification of genes that regulate a left-right asymmetric neuronal migration in Caenorhabditis elegans. Genetics 164, 1355–1367 (2003). 43. G. Esposito, E. Di Schiavi, C. Bergamasco, P. Bazzicalupo, Efficient and cell specific knockdown of gene function in targeted C. elegans neurons. Gene 395, 170–176 (2007). 44. J. S. Lee, A. Ishimoto, S. Yanagawa, Characterization of mouse Dishevelled (Dvl) proteins in Wnt/Wingless signaling pathway. J. Biol. Chem. 274, 21464–21470 (1999). 45. S. Foldynová-Trantírková, P. Sekyrová, K. Tmejová, E. Brumovská, O. Bernatík, W. Blankenfeldt, P. Krejcí, A. Kozubík, T. Dolezal, L. Trantírek, V. Bryja, Breast cancerspecific mutations in CK1e inhibit Wnt/b-catenin and activate the Wnt/Rac1/JNK and NFAT pathways to decrease cell adhesion and promote cell migration. Breast Cancer Res. 12, R30 (2010). 46. K. Tamai, X. Zeng, C. Liu, X. Zhang, Y. Harada, Z. Chang, X. He, A mechanism for Wnt coreceptor activation. Mol. Cell 13, 149–156 (2004). 47. M. T. Veeman, D. C. Slusarski, A. Kaykas, S. H. Louie, R. T. Moon, Zebrafish prickle, a modulator of noncanonical Wnt/Fz signaling, regulates gastrulation movements. Curr. Biol. 13, 680–685 (2003). 48. L. A. Kelley, M. J. E. Sternberg, Protein structure prediction on the Web: A case study using the Phyre server. Nat. Protoc. 4, 363–371 (2009). 49. E. F. Pettersen, T. D. Goddard, C. C. Huang, G. S. Couch, D. M. Greenblatt, E. C. Meng, T. E. Ferrin, UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004). Acknowledgments: We thank X. Wang, R. Moon, M. Maurice, A. Fire, S. Yanagawa, M. Bienz, X. He, and M. Eilers for expression vectors, and the Caenorhabditis Genetic Center (University of Minnesota, Minneapolis) for strains. Funding: This work was funded by the Dutch Cancer Society (HUBR 2008-4114) (R.E.A.G. and H.C.K.), Ministry of Education, Youth, and Sports of the Czech Republic (MSM0021622430), Czech Science Foundation (204/09/H058, 204/09/0498, 13-31488P), EMBO (Installation Grant) (V.B. and O.B.), by the “CEITEC—Central European Institute of Technology” project (CZ.1.05/1.1.00/ 02.0068) from the European Regional Development Fund, by CEITEC open access project (LM2011020), funded by the Ministry of Education, Youth, and Sports of the Czech Republic under the activity “Projects of major infrastructures for research, development and innovations” (K.Š. and Z.Z.), and by Cancer Research UK, Merck Serono, Breast Cancer Campaign, and Tenovus (T.D.). Author contributions: R.E.A.d.G. and H.C.K. designed and carried out the C. elegans experiments. K.Š., B.L.-L., and T.D. performed functional characterization of Huwe1. R.S.G., O.B., R.E.A.d.G., and V.B. designed and carried out cell-based experiments. K.S., Z.Z., and V.M.D. performed MS analysis. R.E.A.d.G., R.S.G., H.C.K., and V.B. wrote the manuscript. Competing interests: The authors declare that they have no competing financial interests. Data and materials availability: C. elegans RNAi data have been submitted to Wormbase (http://www.wormbase.org) and are also available at http://www.sci.muni.cz/ofiz/wp-content/uploads/2012/11/De-Groot-2014-Sci-SignalSuppl.Table-1.pdf. MS data (tables S4 and S5) have been submitted to the PRIDE archive (http://www.ebi.ac.uk/pride/archive/), accession number XXXXXX AQ8. Data from table S3 are available at http://www.sci.muni.cz/ofiz/wp-content/uploads/2012/11/De-Groot-2014-SciSignal-Suppl.Table-3.pdf. Requests for materials should be addressed to H.C.K. and V.B. Submitted 6 December 2013 Accepted 27 February 2014 Final Publication 18 March 2014 10.1126/scisignal.2004985 Citation: R. E. A. de Groot, R. S. Ganji, O. Bernatik, B. Lloyd-Lewis, K. Seipel, K. Šedová, Z. Zdráhal, V. M. Dhople, T. Dale, H. C. Korswagen, V. Bryja, Huwe1-mediated ubiquitylation of dishevelled defines a negative feedback loop in the Wnt signaling pathway. Sci. Signal. 7, ra26 (2014). R E S E A R C H A R T I C L E www.SCIENCESIGNALING.org 18 March 2014 Vol 7 Issue 317 ra26 10 MS no: RA2004985/S/CELL BIOL Supplementary materials Huwe1-mediated ubiquitylation of Dishevelled defines a negative feedback loop in the Wnt signaling pathway Reinoud E.A. de Groot1,8 , Ranjani Sri Ganji2,8 , Ondrej Bernatik2,3 , Bethan Lloyd- Lewis4# , Katja Seipel4 , Kateřina Šedová5,6 , Zbyněk Zdráhal5,6 , Vishnu M. Dhople7 , Trevor Dale4 , Hendrik C. Korswagen1 *, Vitezslav Bryja2,4 * Figure S1: eel-1 is a negative regulator of Wnt signaling in C. elegans (A) Hits from the screen for genes with a role in ubiquitylation that enhance or suppress the QL.d phenotype of vps-29(tm1320) mutants (data are presented mean +/SD and include results from 3 experiments, n>50 per experiment, * indicates p<0.05 (Student’s t-test)). (B) Migration of QL.d cells in vps-29(tm1320) with eel-1 knockdown by egl-20 promoter-directed expression of eel-1 double-stranded RNA (dsRNA). Data are mean +/- SD from 3 independent experiments with more than 30 worms in each. Figure S2 eel-1/Huwe1 is a negative regulator of Wnt signaling that acts upstream of -catenin in the Wnt pathway. (A) Huwe1 knockdown increases TCF reporter activity in 7DF3 (Dvl2-ER) cells after estradiol treatment. (B) Huwe1 knockdown using an independent siRNA increases TCF reporter activity in 7DF3 (Dvl2-ER) cells after estradiol treatment. (C) Knock-down of Huwe1 increases Topflash activity in U2OS cells *p=0.044 **p=0.00034 (Students t-test). Figure S3. Huwe1 interacts with Dvl and promotes Dvl ubiquitylation. (A-C) HEK293T cells were transfected with FLAG-Dvl truncation constructs and HA-∆NHuwe1 as indicated. Co-IP efficiency between FLAG-Dvl3 and HA-∆N-Huwe1 was determined by anti-FLAG immunoprecipitation and subsequent Western blotting. (A.U.: arbitrary units) (D) HEK293T cells were transfected with a FLAG-Dvl1 and HA-∆N-Huwe1 and treated with control or Wnt3a conditioned medium. Cells were lysed and the interaction between FLAG-Dvl1 and HA-∆N-Huwe1 was assayed by co-immunoprecipitation. (E) Co-localization of FLAG-Huwe1 or FLAG-Huwe1(CA) and HA-Dvl2 in the cytoplasmic Dvl punctae. Scale bar, 20 μM Figure S4. Huwe1 promotes Dvl ubiquitylation. (A) Full length Huwe1 or Huwe1(CA) were transfected with HA-Dvl3 and His-ubiquitin After immunoprecipitation of His-ubiquitin modified proteins, HA-Dvl3 was detected by Western blotting. (B) HEK293T cells were transfected with FLAG-Dvl1 truncation constructs, His-ubiquitin and HA-∆N-Huwe1. After pull-down of His-ubiquitin modified proteins, FLAG-Dvl1 was detected by Western blotting. Figure S5. Huwe1 inhibits Dvl multimerization. (A) Example of measurement of FRET efficiency between Dvl proteins. (B) Punctate subcellular localization of wildtype FLAG-Dvl1 and the FLAG-Dvl1DIX7KR mutant. (C) Topflash reporter activity in HEK293T cells expressing wild type Dvl1 or the Dvl1-DIX7KR mutant with or without CK1ε. Table S1. List DUBs, E2 enzymes and E3 ubiquitin ligases that were screened for enhancers or suppressors of the vps-29(tm1320) QL.d migration phenotype. _____________________________________________________________________________________ DUB C04E6.5 DUB C08B11.7 DUB C34F6.9 DUB E01B7.1 DUB F07A11.4 DUB F14D2.7 DUB F21D5.2 DUB F28F8.6 DUB F29C4.5 DUB F30A10.10 DUB F35B3.1 DUB F38B7.5 DUB F46E10.8 DUB F59E12.6 DUB H19N07.2 DUB H34C03.2 DUB K02C4.3 DUB K08B4.5 DUB K09A9.4 DUB R10E11.3 DUB T05H10.1 DUB T22F3.2 DUB T24B8.7 DUB T27A3.2 DUB Y106G6H.12 DUB Y40G12A.1 DUB Y50C1A.1 DUB Y71A12B.9 DUB ZK328.1 E2 B0403.2 E2 C06E2.3 E2 C28G1.1 E2 D1022.1 E2 F25H2.8 E2 F29B9.6 E2 F40G9.3 E2 F49E12.4 E2 F58A4.10 E2 M7.1 E2 R01H2.6 E2 R09B3.4 E2 Y110A2AR.2 E2 Y54E5B.4 E2 Y54G2A.31 E2 Y69H2.6 E2 Y71G12B.15 E2 Y87G2A.9 E2 Y94H6A.6 HECT C34D4.14 HECT D2085.4 HECT F36A2.13 HECT F45H7.6 HECT Y39A1C.2 HECT Y48G8AL.1 HECT Y65B4BR.4 HECT Y67D8C.5 Ubox F59E10.2 Ubox T05H10.5 Ubox T09B4.10 RING B0281.3 RING B0281.8 RING B0393.6 RING B0416.4 RING C01B7.6 RING C01G6.4 RING C02B8.6 RING C06A5.8 RING C06A5.9 RING C11H1.3 RING C12C8.3 RING C15F1.5 RING C16C10.5 RING C16C10.7 RING C17E4.3 RING C17H11.6 RING C18B12.4 RING C18H9.7 RING C26B9.6 RING C30F2.2 RING C32D5.10 RING C32D5.11 RING C32E8.1 RING C34E10.4 RING C36A4.8 RING C45G7.4 RING C49H3.5 RING C52E12.1 RING C53A5.6 RING C53D5.2 RING C55A6.1 RING C56A3.4 RING D2089.2 RING F08B12.2 RING F08G12.5 RING F10D7.5 RING F10G7.10 RING F11A10.3 RING F16A11.1 RING F19G12.1 RING F23B2.10 RING F26F4.7 RING F26G5.9 RING F32A6.3 RING F35G12.9 RING F36F2.3 RING F40G9.12 RING F40G9.14 RING F42C5.4 RING F42G2.5 RING F43C11.7 RING F43G6.8 RING F45G2.6 RING F46F2.1 RING F47G9.4 RING F53F8.3 RING F53G2.7 RING F54B11.5 RING F55A11.7 RING F55A3.1 RING F56D2.2 RING F58B6.3 RING F58E6.1 RING H05L14.2 RING H10E21.5 RING K01G5.1 RING K02B12.8 RING K04C2.4 RING K09F6.7 RING K11D12.9 RING K12B6.8 RING M02A10.3 RING M110.3 RING M142.6 RING M88.3 RING R02E12.4 RING R05D3.4 RING R06F6.2 RING R10A10.2 RING T01C3.3 RING T01G5.7 RING T02C1.1 RING T05A12.4 RING T08D2.4 RING T13A10.2 RING T20F5.6 RING T20F5.7 RING T22B2.1 RING T23F6.3 RING T24D1.2 RING T26C12.3 RING W02A11.3 RING W04H10.3 RING Y105C5B.11 RING Y105E8A.14 RING Y38H8A.2 RING Y45F10B.8 RING Y45F10B.9 RING Y47D3B.11 RING Y47G6A.14 RING Y52E8A.2 RING Y53G8AM.4 RING Y55F3AM.6 RING Y57A10B.1 RING Y71F9AL.10 RING Y73C8C.7 RING Y73C8C.8 RING Y75B8A.10 RING Y7A9C.1 RING ZC13.1 RING ZK1240.1 RING ZK1240.2 RING ZK1240.3 RING ZK1240.6 RING ZK287.5 RING ZK637.14 RING ZK809.7 SKR C06A8.4 SKR C42D4.6 SKR C52D10.6 SKR C52D10.7 SKR C52D10.8 SKR C52D10.9 SKR F13A7.9 SKR F44G3.6 SKR F46A9.4 SKR F46A9.5 SKR F47H4.10 SKR F54D10.1 SKR K08H2.1 SKR R12H7.3 SKR R12H7.5 SKR Y105C5B.13 SKR Y37H2C.2 SKR Y47D7A.1 SKR Y47D7A.8 Table S2. Effect of eel-1(RNAi) on the EGL-20/Wnt dependent migration of the QL.d in various Wnt pathway mutants Control RNAi eel-1 RNAi % QL.d anterior Wild type 0 0 mig-14(mu71)/Wls 100 86 vps-29(tm1320)/retromer 33 8 egl-20(hu105)/Wnt 100 100 mig-5(rh147)/Dvl 100 100 bar-1(ga80)/-catenin 100 100 pop1(hu9)/TCF 29 26 The final position of QL.paa (PVM) was scored as anterior or posterior to the vulva in young adult hermaphrodites using a mec-7::gfp (muIs32) reporter transgene or using Nomarski optics. In each case, n>50. Table S3. Mass-spectrometric analysis of binding partners of Dvl immunopresipitated from mouse embryonic fibroblasts Table S4. Mass-spectrometric analysis of ubiquitylated lysine residues of Dvl1 Table S5. Mass-spectrometric analysis of ubiquitin chain linkage ligated to Dvl1 Sample: Dvl1 +  N-Huwe1 Mass: 8560 Score: 475 Sequences: 9 pI = 6.5 88% Ubiquitin (gi2627129 polyubiquitin) Observed Mr(calc) ppm Score Peptide 730.8968 1459.7783 0.50 72 R.LIFAGKQLEDGR.T + UBI_dT K 48 Di-Gly enrichment Observed Mr(calc) ppm Score Peptide 690.3895 1378.7643 0.19 32 -.MQIFVKTLTGK.T + UBI_dT K 6 698.3866 1394.7592 -0.37 21 -.MQIFVKTLTGK.T + Oxidation (M); UBI_dT K 6 730.8967 1459.7783 0.33 87 R.LIFAGKQLEDGR.T + UBI_dT K 48 459.9943 1835.9489 -0.48 24 K.AKIQDKEGIPPDQQR.L + UBI_dT K 29 or K 33 701.0389 2100.0950 -0.09 22 K.TITLEVEPSDTIENVKAK.I + UBI_dT K 27 or K 29 1122.6013 2243.1910 -1.28 47 R.TLSDYNIQKESTLHLVLR.L + UBI_dT K 63 1201.6381 2401.2588 1.16 58 K.TLTGKTITLEVEPSDTIENVK.A + UBI_dT K 11 841.7784 2522.3129 0.25 24 R.LIFAGKQLEDGRTLSDYNIQK.E + UBI_dT K 48 Sample: Dvl1 Mass: 8560 Score: 243 Sequences: 6 pI = 6.6 85% Ubiquitin (gi2627129 polyubiquitin) Observed Mr(calc) ppm Score Peptide 730.8967 1459.7783 0.33 72 R.LIFAGKQLEDGR.T + UBI_dT K 48 Di-Gly enrichment Observed PTM Mr(calc) ppm Score Peptide 730.8967 1459.7783 0.33 73 R.LIFAGKQLEDGR.T + UBI_dT K 48 Vítězslav Bryja, 2014    Attachments      #19      M.B.C Kilander1 , J. Petersen1 , J. Dahlström, R. Sri Ganji, J. Schuster, N. Dahl, V. Bryja, G.  Schulte.  Preassembly  of  human  Frizzled  6  with  heterotrimeric  Gαi2  is  selectively  impaired  by  the  pathogenic  Frizzled  6  Arg511Cys  mutation  and  is  regulated  by  Disheveled. FASEB J. (fj.13‐246363. Published online February 7, 2014).  1  equal contribution      Impact factor (2012): 5.704  Times cited (without autocitations, WoS, Feb 21st 2014): 0  Significance:  Using  life  imaging  this  study  shows  that  alpha  subunits  of  trimeric  G  proteins are precoupled to Fzd6, a Wnt receptor. This complex rapidly dissociates  following  Wnt5a  treatment  and  is  tightly  controlled  by  Dishevelled.  The  first  mechanistic explanation of the role Galpha in Fzd signaling.  Contibution of the author/author´s team: Biochemical analysis of Dvl‐Frizzled‐Galpha  interaction.            The FASEB Journal • Research Communication Disheveled regulates precoupling of heterotrimeric G proteins to Frizzled 6 Michaela B. C. Kilander,*,1 Julian Petersen,*,1 Kjetil Wessel Andressen,†,‡,§ Ranjani Sri Ganji,ʈ Finn Olav Levy,†,‡,§ Jens Schuster,¶ Niklas Dahl,¶ Vitezslav Bryja,ʈ,# and Gunnar Schulte*,ʈ,2 *Section of Receptor Biology and Signaling, Department of Physiology and Pharmacology, Karolinska Institutet, Stockholm, Sweden; † Department of Pharmacology, Institute of Clinical Medicine, University of Oslo and Oslo University Hospital, Oslo, Norway; ‡ K. G. Jebsen Cardiac Research Centre and § Center for Heart Failure Research, Faculty of Medicine, University of Oslo, Oslo, Norway; ʈ Faculty of Science, Institute of Experimental Biology, Masaryk University, Brno, Czech Republic; ¶ Department of Immunology, Genetics, and Pathology, Science for Life Laboratory, Biomedicinskt Centrum (BMC), Uppsala University, Uppsala, Sweden; and # Department of Cytokinetics, Institute of Biophysics, Academy of Sciences, Prague, Czech Republic ABSTRACT Frizzleds (FZDs) are classified as G-protein-coupling receptors, but how signals are initiated and specified through heterotrimeric G proteins is unknown. FZD6 regulates convergent extension movements, and its C-terminal Arg511Cys mutation causes nail dysplasia in humans. We investigated the functional relationship between FZD6, Disheveled (DVL), and heterotrimeric G proteins. Live cell imaging combined with fluorescence recovery after photobleaching (FRAP) revealed that inactive human FZD6 precouples to G␣i1 and G␣q but not to G␣oA,G␣s, and G␣12 proteins. G-protein coupling is measured as a 10–20% reduction in the mobile fraction of fluorescently tagged G proteins on chemical receptor surface cross-linking. The FZD6 Arg511Cys mutation is incapable of G-protein precoupling, even though it still binds DVL. Using both FRAP and Förster resonance energy transfer (FRET) technology, we showed that the FZD6-G␣i1 and FZD-G␣q complexes dissociate on WNT-5A stimulation. Most important, G-protein precoupling of FZD6 and WNT- 5A-induced signaling to extracellular signal-regulated kinase1/2 were impaired by DVL knockdown or overexpression, arguing for a strict dependence of FZD6– G-protein coupling on DVL levels and identifying DVL as a master regulator of FZD/G-protein signaling. In summary, we propose a mechanistic connection between DVL and G proteins integrating WNT, FZD, G-protein, and DVL function.—Kilander, M. B. C., Petersen, J., Andressen, K. W., Ganji, R. S. Levy, F. O., Schuster, J., Dahl N., Bryja, V., Schulte, G. Disheveled regulates precoupling of heterotrimeric G proteins to Frizzled 6. FASEB J. 28, 000–000 (2014). www.fasebj.org Key Words: GPCR ⅐ WNT-5A ⅐ GNAI1 ⅐ GNAQ Wingless/Int-1 (WNT) family lipoglycoproteins, which activate class Frizzled (FZD1–10) receptors (1–4), are of central importance in embryonic development and human disease (5). The mechanisms of signal initiation from WNT binding to FZD and its interaction with intracellular signaling compounds and subsequent signaling specification are as yet poorly understood (2, 4, 6). Two central players, which act directly downstream of FZDs to initiate several signaling branches, are the phosphoprotein Disheveled (DVL; ref. 7) and heterotrimeric G proteins (8). Ever since the discovery of FZDs and their 7-transmembrane-spanning architecture, it has been surmised that they, at least under certain circumstances, can act as G-protein-coupled receptors (GPCRs; 9, 10). Furthermore, gain- and loss-of-function approaches in diverse experimental systems have indicated a functional role of heterotrimeric G proteins in WNT/FZD signaling (11– 17). More recently, experiments in mammalian cells and tissue preparations have confirmed that WNTs can evoke GDP/GTP exchange at heterotrimeric G proteins at physiological stoichiometry of the involved signaling components, to mediate responses (e.g., proinflammatory 1 These authors contributed equally to this work. 2 Correspondence: Section of Receptor Biology and Signaling, Department of Physiology and Pharmacology, Karolinska Institutet, Nanna Svartz väg 2, S-171 77 Stockholm, Sweden. E-mail: gunnar.schulte@ki.se doi: 10.1096/fj.13-246363 This article includes supplemental data. Please visit http:// www.fasebj.org to obtain this information. Abbreviations: CB1R, cannabinoid 1 receptor; CL, crosslinking; dcFRAP, dual-color fluorescence recovery after photobleaching; DIX, Disheveled axin; DVL, Disheveled; ERK1/2, extracellular signal-regulated kinase 1/2; FRAP, fluorescence recovery after photobleaching; FRET, Förster resonance energy transfer; FZD, Frizzled; GFP, green fluorescent protein; GPCR, G-protein-coupled receptor; i3, third intercellular loop; MYR, myristoylated; P-ERK1/2, phosphorylated extracellular signal-regulated kinase 1/2; PTX, pertussis toxin; ROI, region of interest; SFRP1, secreted Frizzled-related protein 1; WNT, Wingless/Int-1 10892-6638/14/0028-0001 © FASEB The FASEB Journal article fj.13-246363. Published online February 7, 2014. transformation of microglia cells (17–21). However, underlying, mechanistic details of FZD–G-protein coupling and subsequent signaling specification remain unclear. The interface for G-protein interaction on GPCRs lies mainly within the third intracellular loop (i3) and the C terminus (22), and thus it overlaps substantially with regions in FZDs important for DVL interaction (23, 24). FZDmediated recruitment of DVL is based on at least 3 different mechanisms: interaction between the unconventional PDZ ligand domain KTXXXW in the FZD C terminus with the PDZ domain in DVL (23); protein–protein interaction between a discontinuous FZD motif in i3 with the DEP (DVL, EGL-10, and Pleckstrin) domain and the C-terminal domains of DVL (24); and pH-dependent electrochemical interaction of the DEP domain of DVL with acidic membrane lipids (25). The complex regulation of FZD–DVL interaction and the topological overlap of putative FZD–Gprotein and FZD–DVL binding sites raise the question of whether steric hindrance or close cooperation plays a role in FZD–DVL and FZD–G-protein interaction, signal integration, and signal specification. There is some controversy concerning whether GPCRs are precoupled to heterotrimeric G proteins or the GPCR–G-protein interaction is based on random collision (8, 26–28); there is experimental evidence for both scenarios (28–32). Collision coupling postulates receptor–G-protein interaction only on receptor activation. Precoupling of GPCRs with G proteins, however, more readily explains receptor–G-protein selectivity and the rapid responses observed in cells (8). We set out to characterize the FZD6–G-protein coupling mode and a putative regulatory role of DVL therein. Our cellular imaging data show that FZD6 is precoupled to G␣i1 and G␣q and that the receptor–Gprotein complexes are dissociated by agonist stimulation. Furthermore, with gain- and loss-of-function experiments, we found that DVL is required, on the one hand, for FZD6–G-protein precoupling but that, on the other hand, high expression levels of DVL are inhibitory. On the basis of our results, we propose a bimodal regulatory role of DVL to modulate FZD6-mediated G-protein precoupling and signaling. MATERIALS AND METHODS Cell culture and transfections HEK293 cells (American Type Culture Collection, Manassas, VA, USA) were cultured in DMEM supplemented with 10% FBS, 1% penicillin/streptomycin, 1% l-glutamine (all from Invitrogen, Carlsbad, CA, USA) in a humidified CO2 incubator at 37°C. All cell culture plastics were from Sarstedt (Nümbrecht, Germany) unless otherwise specified. For live cell imaging and immunochemical and immunoblot analyses, the cells were seeded on ECM gel-coated (1:300; SigmaAldrich, Stockholm, Sweden) glass-bottomed dishes (MatTek, Ashland MA, USA, or Greiner Bio One, Frickenhausen, Germany; 4-chamber 35-mm glass-bottomed dishes). The cells were transfected with either Lipofectamine 2000:DNA/ RNA 2:1 (Life Technologies, Gaithersburg, MD, USA), 24–48 h before analysis, or CaPO4, 48–72 h before analysis. FZD6-green fluorescent protein (GFP) and FZD6-R511CGFP were constructed as described previously (33). GFPtagged human FZD6 and FZD6-R511C were recloned into the pmCherry-N1 vector by using the restriction enzymes BglII and AgeI. G␣s/␣i1-GFP were provided by Mark Rasenick (University of Chicago, Chicago, IL, USA; ref. 34); G␣q/oAVenus, ␥-Venus were from Nevin A. Lambert (Georgia Health Sciences University, Augusta, GA, USA; refs. 29, 35, 36); G␣12-mCherry was cloned as an N-terminal fusion to the G protein, according to G␣13-GFP10 (37). DVL1-FLAG was from Madelon M. Maurice (University Medical Center, Utrecht, The Netherlands); DVL2-GFP was from Robert J. Lefkowitz (Duke University Medical Center, Durham, NC, USA); DVL2-MYC was from S. A. Yanagawa (Kyoto University, Kyoto, Japan); DVL3-FLAG and deletion mutants were from Randall T. Moon (University of Washington School of Medicine, Seattle, WA, USA); DVL2-CFP was from C. Niehrs (German Cancer Research Center, Heidelberg, Germany); cannabinoid 1 receptor (CB1R)-mCherry was from Zsolt Lenkei (Ecole Supérieure de Physique et Chimie Industrielle– Paristech, Paris, France); and MYR-mCherry was from Jyrki Kukkonen (Turku University Hospital, Turku, Finland). All constructs were confirmed by sequencing. For human DVL silencing, pan-DVL siRNA (AAGUCAACAAGAUCACCUUCU) targeting position 1450–1468 of human DVL1, isoform 1; position 1375–1393 of human DVL1, isoform 2; position 1474–1492 of DVL2; and position 1441– 1459 of DVL3) or Xeragon (Qiagen, Sollentuna, Sweden) control nonsilencing siRNA (AAUUCUCCGAACGUGUCACGU) were added simultaneously with DNA plasmids (38). For all experiments, the cells were transfected at a 3:1:1:1 ratio of receptor:G␣:␤:␥ plasmids or receptor:DVL plasmids at a ratio of 3:1. For human DVL silencing, pan-DVL siRNA was added simultaneously with DNA plasmids (38). For detection of phosphorylated extracellular signal-regulated kinase 1/2 (P-ERK1/2), cells were seeded in 48-well plates. siRNA was transfected for 72 h and DVL overexpression for 24 h. The cells were serum-deprived overnight before stimulation with WNT-5A. Pertussis toxin (PTX) was applied together with starvation medium. Cell lysates were analyzed by polyacrylamide gel electrophoresis and immunoblot analysis. Fluorescence recovery after photobleaching (FRAP), surface cross-linking (CL), and cellular imaging Cells on 35-mm glass-bottomed dishes were placed in a heated (37°C) CO2 (5%) chamber on the stage of a 710 laser-scanning microscope (Zeiss, Jena, Germany). Images were acquired with ϫ40, 1.2 NA C-Apochromat (Zeiss) objective, and 488- and 561-nm laser lines were used to excite the GFP/Venus and mCherry fluorophores, respectively. Photobleaching was performed by 100% laser illumination of a 2.08- ϫ 2.08-␮m area placed over the cell plasma membrane. Fluorescence was monitored pre- and postbleach with lowintensity illumination for 160 s. Average pixel intensity was measured with ZEN2009/2012 software (Zenwaves Software Technologies, Tokyo, Japan), corrected for background fluctuations and bleaching artifacts, and normalized to prebleached intensity (39). For details, see below. For dual-color FRAP (dcFRAP) experiments, cell plasma membrane proteins were immobilized using avidin-biotin CL (35). Briefly, the cells were incubated in 0.5 mg/ml NHSsulfo-LC-LC-biotin and 0.1 mg/ml avidin (both from ThermoFisher Scientific, Kungens Kurva, Sweden) for 15 min each at room temperature and rinsed 3 times each, before, between, and after incubations. Washing and incubation were performed in CL buffer (150 mM NaCl, 2.5 mM KCl, 10 mM HEPES, 12 mM glucose, 0.5 mM CaCl2, and 0.5 mM MgCl2, adjusted to pH 8.0). 2 Vol. 28 May 2014 KILANDER ET AL.The FASEB Journal ⅐ www.fasebj.org To visualize the subcellular localization of the receptor, G protein, and DVL, alone or in combination (Supplemental Figs. S1 and S3), living HEK293 cells expressing FZD6mCherry, G␣i1-GFP, and/or DVL2-CFP were documented with a confocal microscope (LSM710; Zeiss). Calculations for FRAP experiments The recovered mobile fraction (Fm) was calculated as follows: Fm ϭ (IP Ϫ I0)/(II Ϫ I0), where II is the initial intensity measured before bleaching, I0 is the immediate postbleaching fluorescence intensity, and IP is the postbleaching intensity. The mobile fraction was determined by taking the average of the fractions obtained between 85 and 105 s. Acceptor photobleaching Förster resonance energy transfer (FRET) HEK293 cells were seeded on Matrigel (BD Biosciences, Stockholm, Sweden)-coated coverslips in a 24-well plate, transfected with FZD6-mCherry, ␥-Venus, and untagged ␤ and G␣ subunits. The cells were left untreated or stimulated with WNT-5A or PTX before fixation in 4% paraformaldehyde. FRET between FZD6-mCherry and ␥-Venus was performed on an LSM710 (Zeiss), with a ϫ63 oil-immersion objective (PlanApochromat, 1.4 NA; Zeiss) by acceptor photobleaching with 100% laser power of a 561-nm diode laser for 20 s. Signal intensity was routinely reduced by 80–90%. Images were acquired before and after photobleaching with excitation/ emission ranges of 488/493–545 nm (Venus) and 561/562– 681 nm (mCherry). Quantification of pre- and postbleaching Venus emission was determined with region of interest (ROI) analysis in Ն10 individual cells per experiment and condition, with the ZEN2009/2012 software. Data were background corrected and adjusted for fluctuations in intensity with an unbleached ROI as reference. FRET efficiency was calculated as E ϭ [1 Ϫ (Ipre/Ipost)] ϫ 100%. Immunoblot analysis Cells were lysed in 1ϫ loading buffer (10% glycerol; 1% SDS; 100 mM Tris/HCl, pH 7.4; and bromphenol blue) and electrophoretically separated on 8% SDS-polyacrylamide gels. Proteins were electrotransferred to PVDF membranes by using wet blotting. The membranes were blocked in 3% nonfat milk; the primary antibodies used were mouse antiDVL1 (1:500; sc-133525; Santa Cruz Biotechnology, Santa Cruz, CA, USA), rabbit anti-DVL2 (1:1000; 3216; Cell Signaling Technologies, Danvers, MA, USA), mouse anti-DVL3 (1:500; sc-8027; Santa Cruz Biotechnology), mouse anti-FLAG (1:500; F1804, Sigma-Aldrich), mouse anti-MYC (1:500; sc-40, Santa Cruz Biotechnology), rabbit anti-P-ERK1/2 (1:2000; 9101L; Cell Signaling Technology), and mouse anti-␤-actin (1:40 000, A5441; Sigma-Aldrich). The Western Lightning ECL detection kit (PerkinElmer, Wellesley, MA, USA) was used for detection of HRP-conjugated secondary antibodies (Pierce Biotechnology, Rockford, IL, USA, and Sigma-Aldrich). Immunoblots were quantified by densitometry using ImageJ software (U.S. National Institutes of Health, Bethesda, MD, USA). P-ERK1/2 was normalized to ␤-actin signals. Statistical analysis Statistical and graphic analyses were performed with Prism 5 software (GraphPad, San Diego, CA, USA). Data were analyzed by 1-way analysis of variance (ANOVA), with Bonferroni’s post hoc adjustment for multiple comparisons, or Student’s t test. Curve fitting of FRAP data was performed with a 2-phase association nonlinear function according to the least-square fit. All experiments were repeated Ն3 times; FRAP and FRET data are based on 20–70 ROIs/data point from Ն3 independent experiments. Values of P Ͻ 0.05 were considered significant. Data in FRAP curves and bar graphs (FRAP/ FRET) are presented as means Ϯ sem. Protein sequence alignment of the C termini of human FZD1–10 was performed with the STRAP interactive structure-based sequence alignment program (http://www.bioinformatics.org/strap/) according to online instructions. RESULTS FZD6, G proteins, and DVL colocalize in the plasma membrane When the scaffold protein DVL is overexpressed in cells, the predominant distribution pattern is punctate (Supplemental Fig. S1). The protein accumulates in membranefree aggregates on DVL axin (DIX) domain-mediated multimerization (40). This pattern is disrupted on coexpression of class Frizzled cell surface receptors, which results in a distinct membrane association of DVL, as shown for FZD6 in Fig. 1 and Supplemental Fig. S1. Coexpression of G␣i1-GFP, FZD6-mCherry, and DVL2CFP in HEK293 cells resulted in membrane localization of all 3 molecules (Fig. 1A). In this study, we addressed whether this colocalization is the basis for functional interaction by asking whether FZD6 can indeed couple to heterotrimeric G proteins and what role the DVL protein plays in FZD–G-protein coupling. FZD6 is precoupled to heterotrimeric G␣i1 proteins First, to assess FZD6–G-protein interaction, we implemented a dcFRAP approach that allows simultaneous assessment of lateral mobility of 2 fluorescently tagged proteins (refs. 29, 35 and Fig. 1B–E). The experiments are based on immobilization of transmembrane proteins by chemical surface CL. Intracellular proteins, such as the heterotrimeric G proteins or DVL, are not directly affected by surface CL in the dcFRAP assay (Supplemental Fig. S2A–H); their mobility is retarded only when they interact with immobilized transmembrane proteins (35). Indeed, CL with sulfo-NHS-LC-LCbiotin and avidin strongly reduced FZD6 lateral diffusion, whereas the effects on G␣i1-GFP protein mobility were weaker but distinct (Fig. 1B–D). Data quantification revealed an almost complete immobilization of FZD6-mCherry compared with approximately a onefifth decrease in G␣i1-GFP mobile fraction (Fig. 1E). To assure that the observed effects were FZD6-dependent, to exclude general effects of surface CL on G-protein mobility, and to control for the possibility that endogenously expressed and immobilized proteins (e.g., other GPCRs) compromise the interpretation of G␣i1GFP retardation, we combined the expression of myristoylated mCherry (MYR-mCherry) and G␣i1-GFP as a no-receptor control. The mobility of neither MYRmCherry nor G␣i1-GFP was affected by surface immo- 3DVL AS MASTER REGULATOR OF FZD–G-PROTEIN COUPLING bilization, excluding CL and cross-coupling artifacts. In addition, we excluded that receptor internalization distorts the dcFRAP measurements in the time frame of 0–15 min (Supplemental Fig. S2I). To determine whether FZD6-G␣i1 precoupling can be disrupted by agonist treatment, as would be predicted by the ternary complex model (41), we stimulated HEK293 cells expressing FZD6-mCherry and G␣i1GFP with WNT-5A, both before and after surface CL (Fig. 1F). The lack of CL-induced reduction of the mobile fraction of G␣i1-GFP shows that WNT-5A (300 ng/ml; 15 min) led to dissociation of the FZD6-G␣i1 complex. Similarly, pretreatment with PTX (100 ng/ml overnight), which ADP ribosylates G␣i/o proteins, impeded FZD–G-protein precoupling (Fig. 1G). To exclude that endogenously expressed WNT proteins provide FZD6-mCherry with a constitutive activating input, we pretreated the cells with secreted Frizzled-related protein 1 (SFRP1), which acts through sequestration of WNTs (Fig. 1H). Recombinant SFRP1 at 10 ␮g/ml, a dose that efficiently blocks WNT-induced signaling through heterotrimeric G proteins (17, 19, 42), did not affect FZD6-G␣i1 precoupling, arguing against WNTinduced receptor–G-protein complex formation. Figure 1. dcFRAP in combination with cell surface CL reveals FZD6-G␣i precoupling. A) Confocal imaging revealed that FZD6-mCherry, G␣i1-GFP, and DVL2-CFP in HEK293 cells colocalized in the cell membrane. HEK293 cells expressing the individual vectors and FZD6-G␣i and FZD6-DVL2 combinations are shown in Supplemental Fig. S1. Scale bar ϭ 10 ␮m. B) Photomicrographs of HEK293 cells cotransfected with FZD6-mCherry and G␣i-GFP after surface CL are shown before (prebleach), directly after (bleach), and 100 s after photobleaching (postbleach). Arrows mark the bleached stretch of the cell membrane. Scale bar ϭ 2 ␮m. Receptor model clarifies the experimental setup: FZD6-mCherry coexpression with G␣i-GFP. C, D) FRAP curves for fluorescence recovery of FZD6-mCherry (C) and G␣i-GFP (D), before (gray) and after (red) surface CL. E) Summary of C, D. For the CL experiments, open bars show quantitation of the mobile fraction of the receptor protein. Solid bars provide information on the intracellular protein (i.e., G␣i1-GFP). Hatched bars (red) indicate surface CL. Color coding is consistent throughout the figures. The same experimental approach was used in combination with WNT-5A, PTX and SFRP1: F, G) Acute stimulation with WNT-5A (300 ng/ml; 15 min; F) and pretreatment with PTX (100 ng/ml overnight; G) dissociated the FZD6-G␣i1 complex in living cells, visualized by lack of retardation of G␣i1-GFP mobility on surface CL of FZD6-GFP. H) SFRP1, acting by sequestering WNTs, did not affect FZD6-G␣i1 precoupling. Error bars ϭ sem. ns, not significant. ***P Ͻ 0.001 (nϭ3). 4 Vol. 28 May 2014 KILANDER ET AL.The FASEB Journal ⅐ www.fasebj.org FZD6 is precoupled to G␣i1 and G␣q, but not to G␣oA, G␣s, and G␣12 proteins Because the structural basis of receptor–G-protein selectivity is only poorly understood, we also investigated the possibility that FZD6 can precouple to other representatives of the 4 families of heterotrimeric G proteins: G␣i/o, G␣s, G␣q/11, and G␣12/13 (Supplemental Fig. S3). Systematic analysis of FZD6-G␣oA, -G␣s, -G␣q, and -G␣12 mobility by dcFRAP in cotransfection with untagged ␤␥ subunits revealed that FZD6 preferentially precouples to G␣i1 and G␣q rather than to G␣s, G␣oA, or G␣12 (Fig. 2A–D). The question arose of whether E Figure 2. FZD6 differentially couples to members of the heterotrimeric G-protein family. A–D) Analogous to the data in Fig. 1C–E, we investigated the effect of surface CL on mobility of G␣oA-Venus (A), G␣s-GFP (B), G␣q-Venus (C), and G␣12-mCherry (D). FRAP curves and summarizing bar graphs representative of Ն3 independent experiments are shown. Receptor models show the experimental setup. E) dcFRAP methodology was used to analyze FZD6-G␣q coupling in unstimulated and WNT-stimulated (WNT-5A, 300 ng/ml for 5, 10, and 15 min) cells, indicating that receptor–G-protein dissociation was rapid and reversible. Error bars ϭ sem. ns, not significant. **P Ͻ 0.01, ***P Ͻ 0.001 (nϭ3). 5DVL AS MASTER REGULATOR OF FZD–G-PROTEIN COUPLING FZD6-G␣q complexes also dissociate on WNT-5A treatment. Therefore, we first analyzed the lateral mobility of FZD6-mCherry and G␣q-Venus in the presence of untagged ␤␥ subunits, in the absence and presence of 300 ng/ml WNT-5A for 15 min, as we did for G␣i1. No difference in precoupling was observed compared with that in unstimulated cells. However, on kinetic analysis (5, 10, and 15 min after treatment with WNT-5A) we observed a rapid and reversible dissociation, which argues that FZD6-G␣q precoupling was disrupted by agonist treatment and that the precoupled state was reestablished after 15 min (Fig. 2E). Analysis of the FZD6–G-protein complex and its agonist-induced dissociation by FRET To support the idea of FZD–G-protein interaction, we used FRET methodology. Acceptor photobleaching FRET experiments in fixed HEK293 cells expressing FZD6-mCherry, ␥-Venus, and untagged G␣i1 and ␤ subunits showed substantial FRET between FZD6mCherry and ␥-Venus (Fig. 3A), which decreased significantly on WNT-5A (300 ng/ml; 15 min) or PTX treatment (100 ng/ml; overnight) (Fig. 3B) indicating that both agonist treatment and PTX uncoupled the G protein from the receptor. Similar data were obtained when the ␣ subunit was replaced with G␣q, and FRET efficiency was compared in unstimulated and WNT-5Atreated (300 ng/ml; 5 min) cells (Fig. 3C). The FZD6 R511C mutation is selectively impaired in G-protein precoupling A pathogenic Arg511Cys mutation in human FZD6 (FZD6-R511C), which causes nail dysplasia, is dysfunctional in WNT-5A signaling (33, 43). Arg511 is highly conserved among FZD isoforms (Fig. 4A and Supplemental Fig. S4), and it is close to the DVL-binding sequence (KTXXXW; aa 498–503 in human FZD6). Further, Arg511 is located in a basic region (Fig. 4A) reminiscent of the C-terminal regions present in several class A GPCRs and is necessary for GPCR–G-protein interaction (29). To address the role of the FZD6R511C mutation on G-protein precoupling, we used dcFRAP with FZD6-R511C-mCherry and either G␣i1GFP (Fig. 4B, C) or G␣q-Venus (Fig. 4D, E). The mobile fraction of G␣i1-GFP and G␣q-Venus was unaffected by CL, indicating that the mutation disturbs FZD6-R511CmCherry’s precoupling with the G proteins. After establishing that FZD6 R511C is dysfunctional regarding precoupling to G␣i1 and G␣q, we investigated its interaction with DVL. Quantification of FZD–DVL interactions is notoriously difficult, because cell lysis disrupts the electrochemical DVL–membrane lipid interactions necessary for FZD–DVL recruitment. To circumvent cell lysis and to establish a quantitative assay to measure FZD–DVL interactions, we used dcFRAP with cells expressing either FZD6-mCherry or FZD6R511C-mCherry, together with DVL2-GFP. These experiments revealed substantial FZD–DVL interaction, regardless of the Arg511Cys mutation (Fig. 4F–I). In both wild-type and mutated FZD6, CL reduced the mobile fraction of both the membrane-expressed receptor and the pool of membrane-associated DVL2GFP, thus arguing for strong interaction of FZD6 and FZD6-R511C with DVL2-GFP. To confirm that reduction of DVL2-GFP mobility on CL was specific for FZD6, we combined MYR-mCherry and DVL2-GFP as a noreceptor control (Supplemental Fig. S2E–H). DVL is a master regulator of FZD6–G-protein precoupling To further investigate the role of DVL for G␣i1 and G␣q coupling to FZD6, we used loss- and gain-offunction approaches. First, we downregulated DVL1–3 in HEK293 cells expressing FZD6-mCherry by using a pan-DVL siRNA designed to target all 3 human isoforms of DVL (Fig. 5). To investigate the effects of modified DVL levels on G-protein precoupling to FZD6mCherry we used the dcFRAP approach combined with pan-DVL siRNA or DVL1–3 overexpression, creating a cellular system with low (pan-DVL siRNA), intermediate (endogenous), and high (overexpression) DVL levels. In these experiments, reduced DVL1–3 levels disrupted FZD6-G␣i1 (Fig. 5A) and FZD6-G␣q (Fig. 5F) protein precoupling. Moreover, irrespective of the DVL isoform transfected, excess DVL1–3 counteracted FZD6-G␣i1 (Fig. 5B–D) and FZD6-G␣q (Fig. 5G, H) precoupling. When visualizing this complex data set in Figure 3. Analysis of FZD6–G-protein precoupling and complex dissociation by FRET. A) Receptor model indicates the FRET occurring between FZD6-mCherry and ␥-Venus. B, C) Acceptor photobleaching FRET between FZD6-mCherry and ␥-Venus in the presence of untagged ␤ and G␣i1 (B) or G␣q (C) indicates proximity of receptor and ␣ subunit supportive of direct physical interaction. FRET efficiency is strongly reduced on WNT-5A (300 ng/ml; 15 min, B; 5 min, C) or PTX (B) treatment, proving complex dissociation/rearrangement on agonist treatment and PTX-mediated receptor–G␣i1 protein uncoupling. Data are from Ն20 cells/condition from 3 independent experiments. Error bars ϭ sem. ns, not significant. ***P Ͻ 0.001. 6 Vol. 28 May 2014 KILANDER ET AL.The FASEB Journal ⅐ www.fasebj.org a single graph correlating G␣i1 and G␣q coupling with the expression levels of DVL3, it becomes obvious that DVL levels dramatically influenced coupling efficiency of heterotrimeric G proteins to FZD6 (Fig. 6). Most important, we controlled for the FZD selectivity of the inhibitory effects of high DVL levels by assessing G␣i1 precoupling of a non-FZD GPCR, such as CB1R (44), in cells with normal DVL levels and in cells overexpressing DVL3. Even though we could detect a clear CB1R-G␣i1 precoupling comparable to that observed with FZD6, altered DVL levels did not affect CB1R-G␣i1 precoupling (Fig. 7). To identify the structural domains of DVL involved in the negative regulation of FZD–G-protein precoupling, we measured FZD–G-protein dcFRAP in cells coexpressing a set of FLAG-tagged DVL3 deletion mutants (Fig. 8A). Strikingly, only constructs containing the DIX domain interfered with G-protein precoupling. The DIX domain of DVL is known to mediate the DVL–DVL oligomerization essential in forming high-molecularweight DVL complexes and signalosomes (45, 46). WNT-5A-induced G-protein signaling to extracellular signal-regulated kinase1/2 is tightly regulated by DVL levels So far, we know that FZD–G-protein precoupling relies on balanced levels of DVL (Figs. 5 and 6). Since FZD6–G-protein coupling was measured with overexpressed proteins, we asked whether deregulation of DVL levels also affects WNT-induced signaling through heterotrimeric G proteins at endogenous receptor expression levels. Similar to observations in primary microglia cells, where WNT-5A induced a PTX-sensitive, Figure 4. Naturally occurring, pathogenic Arg511Cys FZD6 (FZD6-R511C) mutation is not precoupled to G␣i1 or G␣q. A) Protein sequence alignment of the C termini of human FZD1–10. Red highlights the conserved KTXXXW DVL-binding sequence. Basic amino acids are blue. Green highlights the conserved residue resembling Arg511 in FZD6. B–E) Analysis of FZD6-R511CmCherry in a dcFRAP setting shows that the mutated FZD6 did not precouple to G␣i1-GFP (B, C) or G␣q-Venus (D, E). Lack of FZD6-R511C-mCherry and G␣i1-GFP and G␣q-Venus interaction is visualized by reduced G-protein retardation after CL. F, G) Quantitative analysis of the mobile fractions of FZD6-mCherry and DVL2-GFP in dcFRAP experiments before and after surface CL indicates interaction between the 2 proteins in living cells visualized by the retardation of DVL2 mobility after CL. H, I) dcFRAP analysis of FZD6-R511C-mCherry and DVL2-GFP shows that the receptor mutation did not affect the FZD–DVL interaction in living cells. For color coding, see Fig. 1. Error bars ϭ sem. ns, not significant (nՆ3). ***P Ͻ 0.001. 7DVL AS MASTER REGULATOR OF FZD–G-PROTEIN COUPLING G␣i/o-dependent phosphorylation of extracellular signal-regulated kinases 1/2 (P-ERK1/2; ref. 17), we detected a time-dependent and PTX-sensitive increase in P-ERK1/2 in HEK293 cells stimulated with 300 ng/ml WNT-5A (Fig. 9A, B). Basal P-ERK1/2 levels were normalized to 100%, and the maximum WNT-5Ainduced response was 164.5 Ϯ 18.3% (meanϮsem) at 10 min of exposure to WNT-5A. Whereas PTX reduced the WNT-5A-induced response to 113.8 Ϯ 4.5%, overexpression of DVL1-FLAG resulted in only a 102.3 Ϯ 7.9% P-ERK1/2 response. Maximum ERK1/2 phosphorylation is observed at 10 min, returning toward baseline 20 min after WNT-5A exposure. When overexpressing FLAG-DVL1 and on pan-DVL siRNA treatment (100.4Ϯ11.8%), the WNT-5A-induced P-ERK1/2 response was completely abrogated (Fig. 9A, B), thus arguing that a balance in DVL levels is essential, not only for receptor–G-protein precoupling, but also for the initiation of downstream signaling. Transfection with control siRNA did not affect the WNT-5A-induced P-ERK1/2 response (154.6Ϯ12.5%). DISCUSSION We used cellular imaging to define the mechanisms of FZD–G-protein coupling in unprecedented detail and resolution. We identified FZD6 as a GPCR that is precoupled to the G␣i and G␣q proteins and pinFigure 5. Perturbation of cellular DVL levels abrogates FZD6-G␣i1 and FZD6-G␣q precoupling. A–J) FZD6-G␣i1 (A) and FZD6-G␣q (F) precoupling was disrupted by down-regulation of all DVL isoforms with pan-DVL siRNA, suggesting that DVL is essential for FZD6-G␣i1 protein precoupling. Overexpression of DVL1 (B, G), DVL2 (C, H), and DVL3 (D, I) prevented the CL-induced retardation of G␣i1-GFP (B–D) and G␣q-GFP (F–I), suggesting that high local concentration of DVL disrupts FZD6-G␣ precoupling. Receptor schemes (E, J) clarify the experimental approach used in B–D and F–I, respectively: dcFRAP with FZD6-mCherry and G␣i1-GFP (B–D) or G␣q-Venus (F–I). K, L) Immunoblot analysis confirms knockdown (K) and overexpression (L) of DVL isoforms. Endogenous DVL was detected with isoform-selective antibodies. Overexpressed DVL was detected with antibodies against the epitope tag (FLAG/MYC) fused to DVL isoforms. ␤-Actin was the loading control. For color coding, see Fig. 1. Error bars ϭ sem. ns, not significant (nՆ3). ***P Ͻ 0.001. Figure 6. DVL as a master regulator of FZD–G-protein coupling. A) Data sets from Figs. 1A, 2G, and 5A, D, F, I are compiled in a single bar graph relating FZD–G-protein coupling efficiency (left y axis) to cellular DVL levels (right y axis; red). Error bars ϭ sem. B) Representative blot used for the assessment of DVL3 expression from A in cells with low (pan-DVL siRNA), intermediate (endogenous) and high (DVL3-FLAG overexpression) DVL levels. ␤-Actin was the loading control. Each red dot in A represents DVL3 levels from 1 experiment; DVL levels are relative, with endogenous DVL3 levels set to 1. 8 Vol. 28 May 2014 KILANDER ET AL.The FASEB Journal ⅐ www.fasebj.org pointed DVL as a master regulator of functional FZD– G-protein coupling. Based on our data, we suggest a higher-order complex model, integrating WNT, FZD, heterotrimeric G proteins, and DVL as essential components for WNT/FZD signaling (Fig. 9C), where a delicate balance between heterotrimeric G proteins and DVL levels determines signaling outcome. DVL, on the one hand, is required for FZD–G-protein coupling, but, on the other hand, it has a negative regulatory function. So far, the mechanistic details of this bimodal, somewhat paradoxical action of DVL on FZD6–G-protein precoupling remain obscure; further studies are needed to determine how DVL can be required for GPCR–G-protein precoupling and also act as a negative regulator. Even though the underlying mechanisms of FZD’s interaction with the membrane-tethered heterotrimeric G proteins are still a matter of debate (2, 4, 47), functional data obtained over the years strongly argue that FZDs can signal as GPCRs (2, 4, 9, 11, 13, 14, 16, 19, 20, 43, 48–51). In addition, a bioinformatic analysis suggested that FZD6 couples to the G␣i/o family of heterotrimeric G proteins (52), predictions that were later experimentally confirmed (21). The dcFRAP assay, which has been used to investigate GPCR–G-protein interactions (29, 35), presents a powerful tool to assay dynamic protein–protein interactions. To further support FZD–G-protein interaction, we complemented our dcFRAP analysis with FRET measurements (Fig. 3), which provided evidence for close interaction between Figure 8. DIX domain of DVL is responsible for the negative regulation of FZD6–G-protein precoupling. A) dcFRAP analysis in FZD6-mCherry–G␣i1-GFP, and FZD6-mCherry–G␣q-Venus–transfected cells coexpressing DVL3-FLAG deletion mutants revealed a selective, negative effect of the DVL DIX domain on FZD6–G-protein coupling. Error bars ϭ sem. ns, not significant. *P Ͻ 0.05, **P Ͻ 0.01, ***P Ͻ 0.001. B) Summary of the data shown in the bar graphs. For color coding, see Fig. 1. Figure 7. CB1R-G␣i1 protein precoupling is not disturbed by DVL overexpression. A, B) To control for the FZD selectivity of the DVL-mediated block of GPCR–G-protein precoupling, we used dcFRAP to assess CB1R precoupling to G␣i1 in the absence (A) and presence (B) of DVL3 overexpression. C) Scheme clarifies the experimental approach of CB1R-mCherry–G␣i1-GFP dcFRAP. For color coding, see Fig. 1. Error bars ϭ sem (nϭ3). ***P Ͻ 0.001. 9DVL AS MASTER REGULATOR OF FZD–G-PROTEIN COUPLING receptor and G protein, FZD6-G␣i1/FZD6-G␣q precoupling, and the agonist-/PTX-induced disruption of the receptor–G-protein complex. In contrast to the dcFRAP assay, FRET analysis argues for proximity and therefore in favor of conventional FZD-␤␥–G␣i1 and FZD-␤␥–G␣q interaction in a precoupled state, since the required distance for resonance energy transfer to occur is Ͻ10 nm (53); thus, the complementary FRET experiments validate our conclusions from the dcFRAP experiments. The other way around, a decrease in FRET does not necessarily support physical dissociation of two interacting G proteins but could also monitor structural rearrangement. The dcFRAP assay in combination with agonist stimulation, on the other hand, strongly supports dissociation of the precoupled receptor–G-protein complex, since the CL-based retardation of the intracellular partner originates from interaction with the immobilized transmembrane receptor (35). In addition, the inhibitory effect of PTX on FZD6-G␣i1 precoupling strongly argues for direct GPCR–G-protein interaction, since ADP ribosylation of the far C terminus of the G protein interferes with complex formation. WNT-induced dissociation of an FZD–G-protein-DVL complex has been detected biochemically in L929 cells (54), however, without providing any mechanistic information on either the role of DVL or the FZD isoform involved. Further, the differential coupling of FZD6 to some, but not all, heterotrimeric G proteins expressed at comparable levels serves as an internal control for the specificity of the dcFRAP assay, which apparently is not compromised by forced coexpression. Interestingly, the mutant FZD6-R511C selectively interacts with DVL but not heterotrimeric G proteins. This phenomenon could, in addition to the localization phenotype of the mutant, explain the dysfunction of this receptor in nail development (33). On the other hand, this phenotype may allow some speculation on the receptor’s structure. Introduction of a Cys at position 511 could very well induce an additional lipid anchor, which might distort helix 8. This cytoplasmic domain on the FZD terminus is located in the amino acid stretch of the KTXXXW sequence, as concluded from the recently published crystal structure of Smoothened, a closely related class Frizzled receptor (55), as well as from structural analysis of the FZD1 C-terminal peptide in detergent micelles (56). This hypothetical distortion induced by the additional Cys could affect G-protein precoupling, in agreement with the role of helix 8 for G-protein coupling and selectivity in other GPCRs (57). However, this mutation and the putative distortion of helix 8 apparently do not lead to a substantial decrease in FZD–DVL affinity, which may be explained by the multimodal interaction between DVL and FZD. Classic GPCRs, such as CB1Rs, are known to exist in a precoupled state (8, 30, 45, 58–60). Based on the strong FZD–G-protein association, as measured by dcFRAP, and the PTX sensitivity of FZD6-G␣i1 interaction, we conclude that we measure FZD–G-protein precoupling as a mode of receptor interaction with the GDP-bound, inactive G protein. This precoupling between FZD6 and G␣i1/G␣q appears rather stable, since it can be resolved by dcFRAP measurements, indicating a low degree of dynamics in the absence of receptor activation. The sensitivity to PTX argues either that the C terminus of the G protein is not permanently buried Figure 9. WNT-5A-induced ERK1/2 phosphorylation is G protein dependent and subject to regulation by DVL. A) WNT-5A (300 ng/ml) induced a time-dependent increase in P-ERK1/2 in native HEK293 cells. PTX completely blocked WNT-5A-induced P-ERK1/2, whereas overexpression of DVL1-FLAG completely abrogated WNT-5A-induced P-ERK1/2. Bar graph summarizes optical density data (P-ERK1/2 normalized to ␤-actin; nՆ3). *P Ͻ 0.05, **P Ͻ 0.01. B) Representative immunoblots. C) We propose a higher-order complex model, where DVL levels dictate FZD6–G-protein coupling and thereby functionally integrate G-protein signaling (e.g., to ERK1/2). Stoichiometry of the receptor–G-protein–DVL complex remains unclear. 10 Vol. 28 May 2014 KILANDER ET AL.The FASEB Journal ⅐ www.fasebj.org in the receptor (since PTX would not be able to access it there) or that the interaction shows slow dynamics over time, allowing PTX to access the C terminus during the overnight exposure. We further suggest for FZD6 that agonist (WNT-5A) stimulation promotes an open and active receptor conformation similar to that described for the ␤-adrenergic receptor coupled to G␣s (61). In this state, the affinity between receptor and G protein is transiently increased. The high-affinity complex then promotes GDP release, and subsequent GTP binding leads to complex dissociation, which we monitor by dcFRAP. That ADP ribosylation of the G␣i1 C terminus by PTX prevents FZD–G-protein precoupling argues for a role of the C terminus in this precoupling mode, even in the absence of a fully open, activated, agonist-bound receptor structure. The selective, inhibitory effect of DVL overexpression on FZD–G-protein, but not CB1R–G-protein, precoupling (Fig. 7) indicates that the FZD–G-protein complex is structurally and functionally different from the GPCR (CB1R)–G-protein complex. In addition, dysfunction of WNT-5A-induced ERK1/2 signaling in cells with altered DVL expression levels emphasizes that not only is the FZD–G-protein precoupling perturbed, but also interference with precoupling cannot be compensated by collision coupling-initiated signaling (Fig. 9). Notably, tight regulation of intracellular DVL levels impairs FZD–G-protein precoupling and WNT-induced G-protein signaling, not only in a recombinant overexpression cell system, but also in cells with physiological receptor/G-protein stoichiometry. Based on these findings, one might even speculate that fine-tuning subcellular DVL concentrations presents a means of compartmentalizing FZD/G-protein signals in living cells. In addition, the DVL-mediated regulation of FZD–G-protein precoupling identifies DVL as the first regulator of GPCR–G-protein precoupling, thereby uncovering a novel regulatory mechanism for GPCR signaling. Domain mapping with DVL deletion mutants showed the DIX domain to be responsible for the negative regulation of FZD–G-protein precoupling (Fig. 7). This domain was previously assigned a dominant role in WNT/␤-catenin signaling, signalosome formation, and DVL-mediated low-density lipoprotein receptor–related protein 6 phosphorylation. On the other hand, the DIX domain is not known for mediating WNT/␤-cateninindependent signaling (7, 62). Since the DIX domain mediates DVL–DVL interaction and polymerization, this feature may in fact also be responsible for the inhibition of FZD–G-protein coupling. It could be possible, for example, that FZD-associated endogenous DVL is bound by overexpressed DIX-domain containing DVL molecules (overexpressed mutants or wildtype proteins) in a way that sterically blocks G-protein coupling. FZD–DVL interaction is notoriously difficult to assess and quantify. With the FZD–DVL dcFRAP measurements (Fig. 4), we present for the first time a quantitative method opening new possibilities for measurements of WNT-induced dynamics in DVL localization and its interaction with membrane proteins. In summary, we have shed light on the longstanding question of FZD–G-protein coupling. The identification of the dual role of DVL as a necessary and negatively regulating factor for FZD–G-protein precoupling may explain previous difficulties in mechanistically pinpointing G protein coupling of FZDs in the context of WNT signaling and the ambiguous data on the involvement of G proteins and DVL, as well as their relative localization to each other in several cellular systems (54, 63). Nevertheless, further studies are needed to determine FZD complex stoichiometry and dynamics, to fully understand signal initiation and specification by Class FZD receptors. The authors thank Madelon M. Maurice (University Medical Center, Utrecht, The Netherlands), Robert J. Lefkowitz (Duke University Medical Center, Durham, NC, USA), Randall T. Moon (University of Washington School of Medicine, Seattle, WA, USA), Jyrki Kukkonen (Turku University Hospital, Turku, Finland), Mark Rasenick (University of Chicago, Chicago, IL, USA), Nevin A. Lambert (Georgia Health Sciences University, Augusta, GA, USA), and Zsolt Lenkei (Ecole Supérieure de Physique et Chimie Industrielle–Paristech, Paris, France) for reagents. The work was supported by grants from Karolinska Institutet, the Swedish Research Council (K2008-68P-20810-01-4, K2008-333 68X-20805-01-4, and K2012-67X-20805-05-3), the Swedish Cancer Society (CAN 2008/539, 2011/690), the Knut and Alice Wallenberg Foundation (KAW2008.0149), the Board of Doctoral Education at Karolinska Institutet (to M.B.C.K. and J.P.), Engkvist’s Foundations, Foundation Lars Hiertas Minne, Swedish Royal Academy of Sciences/Foundation Hierta-Retzius Fond, the Swedish Foundation for International Cooperation in Research and Higher Education (STINT), the Czech Science Foundation (204/09/ H058, 204/09/0498, and 13-32990S), the Ministry of Education, Youth and Sports of The Czech Republic (MSM0021622430), and the Program KI-MU (CZ.1.07/2.3.00/20.0180), cofinanced by the European Social Fund and the state budget of the Czech Republic. The authors declare no conflicts of interest. REFERENCES 1. Nusse, R. (2003) Wnts and Hedgehogs: lipid-modified proteins and similarities in signaling mechanisms at the cell surface. Development 130, 5297–5305 2. Schulte, G. (2010) International Union of Basic and Clinical Pharmacology LXXX: the Class Frizzled receptors. Pharmacol. Rev. 62, 632–667 3. Schulte, G., and Bryja, V. (2007) The Frizzled family of unconventional G-protein-coupled receptors. Trends Pharmacol. Sci. 28, 518–525 4. Dijksterhuis, J. P., Petersen, J., and Schulte, G. (2013) WNT/ Frizzled signaling: receptor-ligand selectivity with focus on FZD-G protein signaling and its physiological relevance. [E-pub ahead of print] Br. J. Pharmacol. doi: 10.1111/bph.12364 5. Barker, N., and Clevers, H. (2006) Mining the Wnt pathway for cancer therapeutics. Nat. Rev. Drug Discov. 5, 997–1014 6. Angers, S., and Moon, R. T. (2009) Proximal events in Wnt signal transduction. Nat. Rev. Mol. Cell Biol. 10, 468–477 7. Gao, C., and Chen, Y. G. (2010) Dishevelled: the hub of Wnt signaling. Cell. Signal. 22, 717–727 8. Oldham, W. M., and Hamm, H. E. (2008) Heterotrimeric G protein activation by G-protein-coupled receptors. Nat. Rev. Mol. Cell Biol. 9, 60–71 9. Koval, A., Purvanov, V., Egger-Adam, D., and Katanaev, V. L. (2011) Yellow submarine of the Wnt/Frizzled signaling: sub- 11DVL AS MASTER REGULATOR OF FZD–G-PROTEIN COUPLING merging from the G protein harbor to the targets. Biochem. Pharmacol. 82, 1311–1319 10. Malbon, C. C., Wang, H., and Moon, R. T. (2001) Wnt signaling and heterotrimeric G-proteins: strange bedfellows or a classic romance? Biochem. Biophys. Res. Commun. 287, 589–593 11. Slusarski, D. C., Corces, V. G., and Moon, R. T. (1997) Interaction of Wnt and a Frizzled homologue triggers G-protein-linked phosphatidylinositol signalling. Nature 390, 410–413 12. Slusarski, D. C., Yang-Snyder, J., Busa, W. B., and Moon, R. T. (1997) Modulation of embryonic intracellular Ca2ϩ signaling by Wnt-5A. Dev. Biol 182, 114–120 13. Liu, T., Liu, X., Wang, H., Moon, R. T., and Malbon, C. C. (1999) Activation of rat frizzled-1 promotes Wnt signaling and differentiation of mouse F9 teratocarcinoma cells via pathways that require Galpha(q) and Galpha(o) function. J. Biol. Chem. 274, 33539–33544 14. Liu, X., Liu, T., Slusarski, D. C., Yang-Snyder, J., Malbon, C. C., Moon, R. T., and Wang, H. (1999) Activation of a frizzled-2/ beta-adrenergic receptor chimera promotes Wnt signaling and differentiation of mouse F9 teratocarcinoma cells via Galphao and Galphat. Proc. Natl. Acad. Sci. U. S. A. 96, 14383–14388 15. Sheldahl, L. C., Park, M., Malbon, C. C., and Moon, R. T. (1999) Protein kinase C is differentially stimulated by Wnt and Frizzled homologs in a G-protein-dependent manner. Curr. Biol. 9, 695–698 16. Katanaev, V. L., Ponzielli, R., Semeriva, M., and Tomlinson, A. (2005) Trimeric G protein-dependent frizzled signaling in Drosophila. Cell 120, 111–122 17. Halleskog, C., Dijksterhuis, J. P., Kilander, M. B., BecerrilOrtega, J., Villaescusa, J. C., Lindgren, E., Arenas, E., and Schulte, G. (2012) Heterotrimeric G protein-dependent WNT- 5A signaling to ERK1/2 mediates distinct aspects of microglia proinflammatory transformation. J. Neuroinflammation 9, 111 18. Kilander, M. B., Halleskog, C., and Schulte, G. (2011) Recombinant WNTs differentially activate ␤-catenin-dependent and -independent signalling in mouse microglia-like cells. Acta Physiol. (Oxf.) 203, 363–372 19. Kilander, M. B. C., Dijksterhuis, J. P., Ganji, R. S., Bryja, V., and Schulte, G. (2011) WNT-5A stimulates the GDP/GTP exchange at pertussis toxin-sensitive heterotrimeric G proteins. Cell. Signal. 23, 550–554 20. Koval, A., and Katanaev, V. L. (2011) Wnt3a stimulation elicits G-protein-coupled receptor properties of mammalian Frizzled proteins. Biochem. J. 433, 435–440 21. Katanaev, V. L., and Buestorf, S. (2009) Frizzled Proteins are bona fide G protein-coupled receptors. Nat. Prec. doi: hdl: 10101/npre.2009.2765.1 22. Wess, J. (1998) Molecular basis of receptor/G-protein-coupling selectivity. Pharmacol. Ther. 80, 231–264 23. Wong, H. C., Bourdelas, A., Krauss, A., Lee, H. J., Shao, Y., Wu, D., Mlodzik, M., Shi, D. L., and Zheng, J. (2003) Direct binding of the PDZ domain of Dishevelled to a conserved internal sequence in the C-terminal region of Frizzled. Mol. Cell 12, 1251–1260 24. Tauriello, D. V., Jordens, I., Kirchner, K., Slootstra, J. W., Kruitwagen, T., Bouwman, B. A., Noutsou, M., Rudiger, S. G., Schwamborn, K., Schambony, A., and Maurice, M. M. (2012) Wnt/beta-catenin signaling requires interaction of the Dishevelled DEP domain and C terminus with a discontinuous motif in Frizzled. Proc. Natl. Acad. Sci. U. S. A. 109, E812–E820 25. Simons, M., Gault, W. J., Gotthardt, D., Rohatgi, R., Klein, T. J., Shao, Y., Lee, H. J., Wu, A. L., Fang, Y., Satlin, L. M., Dow, J. T., Chen, J., Zheng, J., Boutros, M., and Mlodzik, M. (2009) Electrochemical cues regulate assembly of the Frizzled/Dishevelled complex at the plasma membrane during planar epithelial polarization. Nat. Cell Biol. 11, 286–294 26. Ayoub, M. A., Al-Senaidy, A., and Pin, J. P. (2012) Receptor-G protein interaction studied by bioluminescence resonance energy transfer: lessons from protease-activated receptor 1. Front. Endocrinol. (Lausanne) 3, 82 27. Neubig, R. R. (1994) Membrane organization in G-protein mechanisms. FASEB J. 8, 939–946 28. Hein, P., Frank, M., Hoffmann, C., Lohse, M. J., and Bunemann, M. (2005) Dynamics of receptor/G protein coupling in living cells. EMBO J. 24, 4106–4114 29. Qin, K., Dong, C., Wu, G., and Lambert, N. A. (2011) Inactivestate preassembly of G(q)-coupled receptors and G(q) heterotrimers. Nat. Chem. Biol. 7, 740–747 30. Nobles, M., Benians, A., and Tinker, A. (2005) Heterotrimeric G proteins precouple with G protein-coupled receptors in living cells. Proc. Natl. Acad. Sci. U. S. A. 102, 18706–18711 31. Gales, C., Van Durm, J. J., Schaak, S., Pontier, S., Percherancier, Y., Audet, M., Paris, H., and Bouvier, M. (2006) Probing the activation-promoted structural rearrangements in preassembled receptor-G protein complexes. Nat. Struct. Mol. Biol. 13, 778–786 32. Gales, C., Rebois, R. V., Hogue, M., Trieu, P., Breit, A., Hebert, T. E., and Bouvier, M. (2005) Real-time monitoring of receptor and G-protein interactions in living cells. Nat. Methods 2, 177– 184 33. Frojmark, A. S., Schuster, J., Sobol, M., Entesarian, M., Kilander, M. B., Gabrikova, D., Nawaz, S., Baig, S. M., Schulte, G., Klar, J., and Dahl, N. (2011) Mutations in Frizzled 6 cause isolated autosomal-recessive nail dysplasia. Am. J. Hum. Genet. 88, 852– 860 34. Yu, J. Z., and Rasenick, M. M. (2002) Real-time visualization of a fluorescent G(alpha)(s): dissociation of the activated G protein from plasma membrane. Mol. Pharmacol. 61, 352–359 35. Qin, K., Sethi, P. R., and Lambert, N. A. (2008) Abundance and stability of complexes containing inactive G protein-coupled receptors and G proteins. FASEB J. 22, 2920–2927 36. Digby, G. J., Lober, R. M., Sethi, P. R., and Lambert, N. A. (2006) Some G protein heterotrimers physically dissociate in living cells. Proc. Natl. Acad. Sci. U. S. A. 103, 17789–17794 37. Yagi, H., Tan, W., Dillenburg-Pilla, P., Armando, S., Amornphimoltham, P., Simaan, M., Weigert, R., Molinolo, A. A., Bouvier, M., and Gutkind, J. S. (2011) A synthetic biology approach reveals a CXCR4-G13-Rho signaling axis driving transendothelial migration of metastatic breast cancer cells. Sci. Signal. 4, ra60 38. Bryja, V., Schambony, A., Cajanek, L., Dominguez, I., Arenas, E., and Schulte, G. (2008) Beta-arrestin and casein kinase 1/2 define distinct branches of non-canonical WNT signalling pathways. EMBO Rep. 9, 1244–1250 39. Phair, R. D., Gorski, S. A., and Misteli, T. (2004) Measurement of dynamic protein binding to chromatin in vivo, using photobleaching microscopy. Methods Enzymol. 375, 393–414 40. Schwarz-Romond, T., Merrifield, C., Nichols, B. J., and Bienz, M. (2005) The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J. Cell Sci. 118, 5269–5277 41. De Lean, A., Stadel, J. M., and Lefkowitz, R. J. (1980) A ternary complex model explains the agonist-specific binding properties of the adenylate cyclase-coupled beta-adrenergic receptor. J. Biol. Chem. 255, 7108–7117 42. Halleskog, C., and Schulte, G. (2013) Pertussis toxin-sensitive heterotrimeric G(␣i/o) proteins mediate WNT/beta-catenin and WNT/ERK1/2 signaling in mouse primary microglia stimulated with purified WNT-3A. Cell. Signal. 25, 822–828 43. Cui, C. Y., Klar, J., Georgii-Heming, P., Frojmark, A. S., Baig, S. M., Schlessinger, D., and Dahl, N. (2013) Frizzled6 deficiency disrupts the differentiation process of nail development. J. Invest. Dermatol. 133, 1990–1997 44. Vásquez, C., and Lewis, D. L. (1999) The CB1 cannabinoid receptor can sequester G-proteins, making them unavailable to couple to other receptors. J. Neurosci. 19, 9271–9180 45. Bilic, J., Huang, Y. L., Davidson, G., Zimmermann, T., Cruciat, C. M., Bienz, M., and Niehrs, C. (2007) Wnt induces LRP6 signalosomes and promotes dishevelled-dependent LRP6 phosphorylation. Science 316, 1619–1622 46. Schwarz-Romond, T., Fiedler, M., Shibata, N., Butler, P. J., Kikuchi, A., Higuchi, Y., and Bienz, M. (2007) The DIX domain of Dishevelled confers Wnt signaling by dynamic polymerization. Nat. Struct. Mol. Biol. 14, 484–492 47. Malbon, C. C. (2011) Wnt signalling: the case of the ‘missing’ G-protein (published correction appears in Biochem J., 434, 575). Biochem. J. 433, e3–e5 48. Ahumada, A., Slusarski, D. C., Liu, X. X., Moon, R. T., Malbon, C. C., and Wang, H. Y. (2002) Signaling of rat Frizzled-2 through phosphodiesterase and cyclic GMP. Science 298, 2006– 2010 49. Liu, T., DeCostanzo, A. J., Liu, X., Wang, H., Hallagan, S., Moon, R. T., and Malbon, C. C. (2001) G protein signaling from 12 Vol. 28 May 2014 KILANDER ET AL.The FASEB Journal ⅐ www.fasebj.org activated rat frizzled-1 to the beta-catenin-Lef-Tcf pathway. Science 292, 1718–1722 50. Sheldahl, L. C., Slusarski, D. C., Pandur, P., Miller, J. R., Kuhl, M., and Moon, R. T. (2003) Dishevelled activates Ca2ϩ flux, PKC, and CamKII in vertebrate embryos. J. Cell Biol. 161, 769–777 51. Slusarski, D. C., Yang-Snyder, J., Busa, W. B., and Moon, R. T. (1997) Modulation of embryonic intracellular Ca2ϩ signaling by Wnt-5A. Dev. Biol. 182, 114–120 52. Wang, H. Y., Liu, T., and Malbon, C. C. (2006) Structurefunction analysis of Frizzleds. Cell. Signal. 18, 934–941 53. Lohse, M. J., Nuber, S., and Hoffmann, C. (2012) Fluorescence/ bioluminescence resonance energy transfer techniques to study G-protein-coupled receptor activation and signaling. Pharmacol. Rev. 64, 299–336 54. Liu, X., Rubin, J. S., and Kimmel, A. R. (2005) Rapid, Wntinduced changes in GSK3beta associations that regulate betacatenin stabilization are mediated by Galpha proteins. Curr. Biol. 15, 1989–1997 55. Wang, C., Wu, H., Katritch, V., Han, G. W., Huang, X. P., Liu, W., Siu, F. Y., Roth, B. L., Cherezov, V., and Stevens, R. C. (2013) Structure of the human smoothened receptor bound to an antitumour agent. Nature 497, 338–343 56. Gayen, S., Li, Q., Kim, Y. M., and Kang, C. (2013) Structure of the C-terminal region of the Frizzled receptor 1 in detergent micelles. Molecules 18, 8579–8590 57. Wess, J., Han, S. J., Kim, S. K., Jacobson, K. A., and Li, J. H. (2008) Conformational changes involved in G-protein-coupledreceptor activation. Trends Pharmacol. Sci. 29, 616–625 58. Hein, P., and Bünemann, M. (2009) Coupling mode of receptors and G proteins. Naunyn Schmiedebergs Arch. Pharmacol 379, 435–443 59. Neubig, R. R., Gantzos, R. D., and Thomsen, W. J. (1988) Mechanism of agonist and antagonist binding to alpha 2 adrenergic receptors: evidence for a precoupled receptorguanine nucleotide protein complex. Biochemistry 27, 2374– 2384 60. Philip, F., Sengupta, P., and Scarlata, S. (2007) Signaling through a G Protein-coupled receptor and its corresponding G protein follows a stoichiometrically limited model. J. Biol. Chem. 282, 19203–19216 61. Rasmussen, S. G., DeVree, B. T., Zou, Y., Kruse, A. C., Chung, K. Y., Kobilka, T. S., Thian, F. S., Chae, P. S., Pardon, E., Calinski, D., Mathiesen, J. M., Shah, S. T., Lyons, J. A., Caffrey, M., Gellman, S. H., Steyaert, J., Skiniotis, G., Weis, W. I., Sunahara, R. K., and Kobilka, B. K. (2011) Crystal structure of the b2 adrenergic receptor-Gs protein complex. Nature 477, 549–555 62. Boutros, M., and Mlodzik, M. (1999) Dishevelled: at the crossroads of divergent intracellular signaling pathways. Mech. Dev. 83, 27–37 63. Bikkavilli, R. K., Feigin, M. E., and Malbon, C. C. (2008) G alpha(o) mediates WNT-JNK signaling through Dishevelled 1 and 3, RhoA family members, and MEKK 1 and 4 in mammalian cells. J. Cell Sci. 121, 234–245 Received for publication November 22, 2013. Accepted for publication January 23, 2014. 13DVL AS MASTER REGULATOR OF FZD–G-PROTEIN COUPLING   Supplementary Figure S1: Localization of FZD6, Gi1 and DVL overexpressed in HEK293 cells. (A) Confocal photomicrographs show FZD6-mCherry, Gi1-GFP or DVL2-CFP expression in transiently transfected and living HEK293 cells. Note the membranous localization of FZD6 and Gi1 and punctate distribution pattern in DVL2-expressing cells, when each plasmid is transfected alone. (B) FZD6-mCherry recruits DVL2-CFP to the membrane and co-expression prevents DVL from punctate aggregation in the cytosol. Further, both FZD6-mCherry and Gi1-GFP co-localize in the plasma membrane of HEK293 cells Triple transfection is shown in Fig. 1a in the main body of the paper. Scale bar = 10 µm.   Supplementary Figure 2: MYR-mCherry “no receptor” controls for specificity of dcFRAP experiments. In order to control for specificity of FZD effects and possible crosslinking artifacts, dual color FRAP experiments were performed in HEK293 cells co-expressing MYR-mCherry and the Gi1-GFP (A-D) or MYR-mCherry and DVL2-GFP (e-h) serving as “no receptor” control. Before and after crosslinking (CL), the lateral mobility of both proteins was assessed. Panels (A,E) compare the mobile fractions of MYR-mCherry and Gi1-GFP/DVL2-GFP indicating that CL affects neither of them. This control is especially important to support the specificity of FZD-G protein and FZD-DVL2 interactions in the dual color FRAP and excludes interaction with endogenously expressed, cross-linked membrane proteins. Schemes indicate the replacement of FZD6-mCherry with MYR-mCherry in combination with Gi1-GFP or DVL2-GFP for dcFRAP experiments. (I) Changes in receptor mobility upon WNT stimulation could be affected by receptor internalization. The photomicrographs show a time series performed on a FZD6-mCherry expressing HEK293 cells in the absence of stimulation and with 5 and 15 min WNT-5A (300 ng/ml) stimulation. No distinct receptor internalization was observed in this time frame during which dcFRAP measurements were performed. 1   Supplementary Figure S3: Subcellular localization of FZD6 and heterotrimeric G proteins in HEK293 cells. In the main body of the paper (Fig. 1 and Fig. 2) we address whether FZD6mCherry precouples to heterotrimeric G proteins. We have used fluorescently-tagged Gsubunits of representatives of the four families of G proteins (Gs/Gi/Gq/G12) in combination with tagged FZD6 for double color FRAP analysis. The photomicrographs show confocal images from FZD6mCherry/GFP and fluorescently-tagged G subunits (in coexpression with  subunits) in living cells. Both FZD6 and the G subunits were expressed in the membrane, a prerequisite for their physical interaction. Size bar = 10 µm.   Supplementary Fig. 4: Protein sequence alignment of FZD C-terminal sequences from different species show evolutionary conservation of the arginine resembling Arg511 in human FZD6. C terminal sequences for Class Frizzled receptors from Xenopus laevis (XFz), Drosophila melanogaster (Dfz), and Caenorhabditis elegans (MIG-1, MOM-5, CFZ-2, LIN-17) are shown. Basic amino acids are labeled in blue. Red highlights the conserved KTXXXW DVL-binding sequence. Green pinpoints the amino acid resembling the Arg511 in human FZD6. The C terminal sequences are numbered from 1, resembling the first amino acid after the seventh transmembrane helix. Note the occurrence of basic regions in most of the FZDs in Xenopus laevis as well as Drosophila melanogaster Dfz1, 2. Sequence alignment was done with STRAP Interactive Structure based Sequences Alignment Program available at http://www.bioinformatics.org/strap/       Vítězslav Bryja, 2014    Attachments      #20      O.  Bernatík1 ,  K.  Šedová1 ,  R.  Sri  Ganji,  I.  Červenka,  L.  Trantírek,  Z.  Zdráhal,  V.  Bryja.  Functional  analysis  of  Dishevelled‐3  phosphorylation  identifies  distinct  mechanisms  driven by Casein Kinase 1ε and Fzd5 (J. Biol. Chem., in revision).      Impact factor (2012): 4.651  Times cited (without autocitations, WoS, Feb 21st 2014): 0  Significance:  In  this  study  we  have  for  the  first  time  described  all  (or  almost  all)  phosphorylation  events  triggered  by  CK1ε  on  Dishevelled‐3.  Further  functional  analysis  confirmed  in  much  greater  detail  the  dual  role  of  CK1  in  Dvl  biology  postulated  by  us  earlier  (see  study  #12).  Moreover,  we  uncovered  unexpected  fundamental differences between phosphorylation triggered by CK1 and Frizzled,  a Wnt receptor.  Contibution of the author/author´s team: The study has been designed and performed  in our lab. The mass spectrometry analysis has been done in collaboration with  proteomic facility headed by Z. Zdráhal.          Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 1 Functional analysis of Dishevelled-3 phosphorylation identifies distinct mechanisms driven by Casein Kinase 1 and Frizzled5 Ondřej Bernatík1,2! , Kateřina Šedová3,4! , Ranjani Sri Ganji1 , Igor Červenka1 , Lukáš Trantírek5,6 , Zbyněk Zdráhal3,4 , Vitězslav Bryja1,2 * 1 Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlarska 2, 61137 Brno, Czech Republic; 2 Department of Cytokinetics, Institute of Biophysics Academy of Sciences of the Czech Republic, Kralovopolska 135, 61265 Brno, Czech Republic; 3 National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Kamenice 5, 62500 Brno, Czech Republic; 4 Research Group–Proteomics and 5 Structural Biology Program, Central European Institute of Technology (CEITEC), Masaryk University, Kamenice 5, 62500 Brno, Czech Republic; 6 Cellular Protein Chemistry, Utrecht University, Padualaan 8, 3584 CH Utrecht, Netherlands, Running title: Phosphorylation of Dvl3 by CK1ε: MS/MS analysis ! Authors contributed equally to this work *Address correspondence to:Vitezslav Bryja Ph.D., Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlarska 2, 611 37, Brno, Czech Republic, phone: +420-549493291, Email: bryja@sci.muni.cz Key words: Dishevelled-3, Casein kinase 1ε, Frizzled 5, phosphorylation, MS/MS, Wnt/-catenin pathway Background: Phosphorylation of Dishevelled (Dvl) by casein kinase 1ε (CK1ε) is a key event in Wnt signal transduction. Results: Dvl3 residues phosphorylated by CK1ε were identified by proteomics and analyzed functionally. Conclusion: Individual phosphorylation events control different aspects of Dvl biology. Significance: CK1ε and Fzd5, a Wnt receptor, act on Dvl via distinct mechanism, suggesting that CK1ε is not directly downstream of Frizzled5. ABSTRACT: Dishevelled-3 (Dvl3), a key component of the Wnt signaling pathways, acts downstream of Frizzled (Fzd) receptors and gets heavily phosphorylated in response to pathway activation by Wnt ligands. Casein kinase (CK) 1ε was identified as the major kinase responsible for Wnt-induced Dvl3 phosphorylation. Currently it is not clear, which Dvl residues are phosphorylated and what is the consequence of individual phosphorylation events. In the present study, we employed mass spectrometry to analyze in a comprehensive way the phosphorylation of human Dvl3 induced by CK1ε. Our analysis revealed more than 50 phosphorylation sites on Dvl3; only minority of these sites was found dynamically induced after co-expression of CK1ε and, surprisingly, phosphorylation of one cluster of modified residues was downregulated. Dynamically phosphorylated sites were analyzed functionally. Mutations within PDZ domain (S280A and S311A) reduced the ability of Dvl3 to activate TCF/LEF-driven transcription. In contrast, mutations of clustered S/T in the Dvl3 C-terminus prevented ability of CK1ε to induce electrophoretic mobility shift of Dvl3 and its even subcellular localization. Surprisingly, mobility shift and subcellular localization changes induced by Fzd5, a Wnt receptor, were in all these mutants indistinguishable from wild type Dvl3. In summary, our data on the molecular level (i) support previous assumption that CK1ε acts via phosphorylation of distinct residues as the activator as well as shut off signal (via negative effects on Dvl polymerization) of downstream Wnt/-catenin signaling, and (ii) suggest that CK1ε acts on Dvl via different mechanism than Fzd5 and is not directly downstream of Fzd receptor complex. Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 2 INTRODUCTION Wnt/-catenin pathway is a conserved signaling machinery, which coordinates cell proliferation, cell fate and cell differentiation during development and during maintenance of homeostasis in the adult tissues (1,2). Dishevelled (Dvl) with three isoforms in vertebrates (Dvl1, Dvl2 and Dvl3) is a key cytoplasmic protein required for signal transduction of the canonical Wnt signal. Dvl is a modular protein composed of three domains, N terminal DIX (Dishevelled, Axin), central PDZ (Postsynaptic density, Discs large, Zonula ocludens), and DEP (Dvl, Egl-10, Pleckstrin). The linking regions between the domains do not seem to have well defined tertiary structure, and might thus act as “hinge” regions, allowing Dvl to achieve various conformations. Although the reports about the function of individual domains are not always in full agreement, it is commonly acknowledged that DIX and PDZ domains are necessary for transduction of canonical Wnt signal, whereas PDZ and DEP are necessary for other, so called non-canonical Wnt pathways (3-5). The part of Dvl located Cterminally from all three domains has no clear attributed function, although it was shown that it has inhibitory effect on canonical Wnt pathway (6). Upon stimulation of cells by Wnt-3a, which is the best defined ligand of Wnt/β-catenin pathway, Dvl electrophoretic mobility gets reduced due to multiple phosphorylations (7). Wnt-3a-induced activation of endogenous Dvl leads to a variant of Dvl with retarded electrophoretic mobility named phosphorylated and shifted (PS)-Dvl (8-10). It has been clearly demonstrated that both the activation of Dvl in the Wnt/β-catenin pathway (11-15) and Wntinduced PS-Dvl formation is dependent on casein kinase 1 (CK1)/ activity (8,10). From the several protein kinases identified as Dvl binding partners, only CK1δ and CK1ε (for simplicity referred in further text as CK1ε) were shown to be required for Wnt-induced Dvl shift (11,16) and at the same time capable of inducing PS-Dvl when overexpressed (13). Serines (S), threonines (T) and tyrosines (Y), residues, which can be modified by phosphorylation, represent roughly15% of Dvl aminoacid (aa) residues. Despite the fact that many phosphorylation sites have been identified in Dvl (17-24), it is currently unknown, which Dvl residues are dynamically phosphorylated by CK1ε. In order to overcome this caveat we have used mass-spectrometry (MS) based approach to identify in an unbiased way the phosphorylation sites of human (h) Dvl3 in absence and presence of overexpressed CK1ε. We have identified more than 50 phosphorylated S and T residues on Dvl3. Majority of these sites were found to be constitutively phosphorylated. Using criteria such as the site conservation, phosphorylation dynamics and the position in Dvl3 we have selected several phosphorylation sites/clusters for mutation and subsequent functional analysis. Among these we have identified phosphorylation sites which (i) contribute to the electrophoretic mobility of Dvl3, (ii) control Dvl3-driven activation of the Wnt/β-catenin pathway and (iii) are required for even subcellular localization of Dvl3. EXPERIMENTAL PROCEDURES Cell Culture, Transfection, and Treatments. - HEK-293t cells were propagated in DMEM/10% FCS/2 mM L-glutamine/50 units/ml penicillin/50 units/ml streptomycin. Cells (40,000–60.000 per well) were seeded in 24-well plates. The next day, cells were transfected using polyethylenimine (PEI) in a stoichiometry of 2.5 μl PEI per 1 μg of DNA. Cells were harvested for immunoblotting or immunocytofluorescence 24 h after transfection, only for WB analysis of Fzd induced Dvl shift samples were collected after 12h. The following plasmids have been described previously: FLAG-Dvl3 wt (25), FLAG Dvl1 wt (26), and FLAG-Dvl2 wt (27), CK1ε (28), V5-Fzd5 (29). Mutagenesis was performed using Quikchange XL kit following manufacturer’s instructions (Stratagene #200518). All mutations were verified by sequencing. Dual Luciferase Assay, Western blotting, pS643 Dvl3 antibody production and immunoprecipitation - For the luciferase reporter assay, cells were transfected with 0.1 μg Super8X TopFlash construct and 0.1 μg pRLTK luc (Renilla) luciferase construct per well in a 24-well plate, 24 hours after seeding. Cells were then transfected with corresponding plasmids, and processed 24 hours after transfection. For the TopFlash assay, a Promega dual luciferase assay kit was used according to manufacturer’s instructions. Relative luciferase units (RLU) of Firefly luciferase were measured on a MLX luminometer (Dynex Technologies) Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 3 and normalized to the Renilla luciferase signal. Immunoblotting and sample preparation was performed as previously described (30) The following antibodies were used: FLAG M2 (F1804, Sigma-Aldrich), CK1ε (sc-6471; Santa Cruz Biotechnology), V5 (R960-25, Invitrogen), FLAG (F7425, Sigma-Aldrich). Anti-hDvl3-pS643 (pS643) antibody was prepared by immunizing rabits by HHSLAS(pS)LRSHH peptide. Immunization and production of antibody was performed on a service basis by Moravian Biotechnology (http://www.moravian-biotech.com/). For MS/MS based identification of phosphorylation HEK293t cells were seeded on 15-cm dishes and transfected with corresponding plasmids 24h after seeding. Immunoprecipitation protocol used was modified from (31). In short, 2 ml of cold lysis buffer supplemented with protease inhibitors (Roche, 11836145001), phosphatase inhibitors (Calbiochem, 524625), 0.1 mM DTT and 10 mM NEM (N-ethylmaleimide) (Sigma-Aldrich E3876) was used for lysis of one 15 cm dish. Lysate was collected after 20 minutes of lysis on 4o C, and was cleared by centrifugation at 16.1 RCF for 20 minutes. Three μg of antibody were used per sample. Samples were incubated with the antibody for 40 min, then 45 μl of G protein sepharose beads (GE Healthcare, 17- 0618-05) equilibrated in the lysis buffer were added to each sample. Samples were incubated on the carousel overnight, washed 6 times with lysis buffer, 40 μl of 2x Laemmli buffer was added and samples were boiled. The antibody used for immunoprecipitation was FLAG M2 (F1804; Sigma) Immunocytofluorescence - HEK-293t cells were seeded (2x105 cells/well) on gelatine coated coverslips in 24-well plates. Cells were transfected the next day and 24 hours later were fixed in fresh 4% paraformaldehyde, permeabilized with 0.05% Triton-X100, blocked with PBTA (3% BSA, 0,25% Triton, 0,01% NaN3) for 1 hour and incubated overnight with primary antibodies. Next day, coverslips were washed in PBS, and incubated with secondary antibodies conjugated to Alexa Fluor 488 (Life Technologies A11001) or Alexa Fluor 594 (Life Technologies A11058), washed with PBS, stained with DAPI (1:5000) and mounted on microscopic slides. Cells were visualized using Olympus IX51 fluorescent microscope. Hundred positive cells per condition were analyzed and scored based on the pattern of intracellular Dvl3 distribution. 3D predictions, folding, kinase motives Analysis and modeling of Dvl3 PDZ domain 3D structure was performed using Chimera software (32). Folding prediction was generated by PONDR-FIT prediction software (33). Phosphorylation prediction was performed using GPS 2.1 program using predefined values for high, medium and low stringency of the prediction (34). LC-MS/MS analysis - Samples from immunoprecipitation were separated on a 8% SDS-PAGE, fixed with solution A (50% methanol, 10% acetic acid) then stained with Coomassie (0.1% Coomassie briliant blue (Sigma, BO149), 20% methanol, 10% acetic acid) for 2 hours. Gel was destained using solution A. Corresponding 1-D bands were excised. After destaining, the proteins in gel pieces were incubated with 10 mM DTT at 56°C for 45 min. After removal of DTT excess samples were incubated with 55 mM IAA at room temperature in darkness for 30 min, then alkylation solution was removed and gel pieces were hydrated for 45 min at 4 °C in digestion solution (5 ng/μl trypsin, sequencing grade, Promega, Fitchburg, Wisconsin, USA, in 25 mM AB). The trypsin digestion was performed for 2 hours at 40°C on Thermomixer (750 rpm; Eppendorf, Hamburg, Germany). Subsequently, the tryptic digests were cleaved by chymotrypsin (5 ng/μl, sequencing grade, Roche, Basel, Switzerland, in 25 mM AB) for 2 hours at 30 °C or 40°C. Digested peptides were extracted from gels using 50% ACN solution with 2.5% formic acid and concentrated in speedVac concentrator (Eppendorf, Hamburg, Germany). The aliquot (1/10) of concentrated sample was directly analysed by LC-MS/MS for protein identification. The rest of sample was used for phosphopeptide analysis. Sample was diluted with acidified acetonitrile solution (80% ACN, 2% FA). Phosphopeptides were enriched using Pierce Magnetic Titanium Dioxide Phosphopeptide Enrichment Kit (Thermo Scientific, Waltham, Massachusetts, USA) according to slightly modified manufacturer protocol (samples were mixed with binding solution in ratio 1:2 prior loading and one additional washing step with binding solution Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 4 after phosphopeptides capture was implemented). Eluates were concentrated under vacuum and then diluted in 10 μl of 0.1% FA solution before LC-MS/MS analysis. Liquid chromatography-tandem mass spectrometry (LC–MS/MS) analysis was performed using reverse phase RSLCnano system (Dionex, Sunnyvale, CA, USA) on-line coupled with an HCT Ultra PTM Discovery System ion trap mass spectrometer equipped with ETD source II (Bruker Daltonik, Bremen, Germany) or with an Orbitrap Elite hybrid spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). (i) LC-MS/MS with HCT Ultra PTM Discovery System. Samples (10 μl) were injected in loading solution (0.1% FA) on trapping column (100 μm × 30 mm) filled with 4-μm Jupiter Proteo sorbent (Phenomenex, Torrance, CA, USA) where the analytes were desalted and concentrated. Subsequently, the analytes were eluted from the trapping column using an acetonitrile/water gradient (350 nl/min) onto a fused-silica capillary column (100 μm × 180 mm), on which they were separated. The column was filled with 3.5-μm X-Bridge BEH 130 C18 sorbent (Waters, Milford, MA, USA). The mobile phase A and B consisted of 0.1% formic acid in water and in acetonitrile, respectively. The gradient elution started at 4 % of mobile phase B and increased to 30% B after 30 min and to 60% B in the next 10 minutes, then 90% B was reach in one minute and system remained at this state for 5 min, finally proportion of phase B was reduced to 4% B in 1 min. Equilibration of the precolumn and the column lasting 25 min was done prior to sample injection to sample loop. The analytical column outlet was directly connected to the nanoelectrospray ion source. Mass spectrometer was operated in the positive ion mode with collision induced dissociation (CID) followed, in case of phosphopeptides, by data-depended electron transfer dissociation (ETD) fragmentation; triggered by detection of neutral loss of phosphoric acid in CID spectrum. Nitrogen was used as nebulizing as well as drying gas. The pressure of nebulizing gas was 8 psi. The temperature and flow rate of drying gas were set to 300 ºC and 6 l/min, respectively. The capillary voltage was 4.0 kV. Ion charge control (ICC) controlling the filling of ion trap was set to 200,000. The mass spectrometer scanned in m/z range of 300 –1500 for MS and 100-2500 for MS/MS operation. The data were processed with DataAnalysis 4.0 and BioTools 3.0 software (Bruker Daltonik) (ii) LC-MS/MS with Orbitrap Elite hybrid spectrometer. Prior to LC separation, tryptic digests were online concentrated and desalted using trapping column (100 μm × 30 mm) filled with 3.5-μm X-Bridge BEH 130 C18 sorbent (Waters, Milford, MA, USA). After washing of trapping column with 0.1% FA, the peptides were eluted (300 nl/min) from the trapping column onto a Acclaim Pepmap100 C18 column (2 µm particles, 75 μm × 250 mm; Thermo Fisher Scientific, Waltham, MA, USA) by the following gradient program (mobile phase A: 0.1% FA in water; mobile phase B: ACN:methanol:2,2,2-trifluoroethanol (6:3:1; v/v/v) containing 0.1% FA): the gradient elution started at 2% of mobile phase B and increased from 2% to 45% during the first 90 min (11% in the 30th, 25% in the 60th and 45% in 90th min), then increased linearly to 95% of mobile phase B in the next 5 min and remained at this state for the next 15 min. Equilibration of the trapping column and the column was done prior to sample injection to sample loop. The analytical column outlet was directly connected to the Nanospray Flex Ion Source (Thermo Fisher Scientific, Waltham, MA, USA). MS data were acquired in a data-dependent strategy selecting up to top 10 precursors based on precursor abundance in the survey scan (350-1700 m/z). Mass spectrometer was operating in the positive ion mode with collision induced dissociation (CID) followed, in case of neutral loss detection (32.7, 49.0, 65.3 and 98; with m/z tolerance of -1.5 and +0.5 Da), by electron transfer dissociation (ETD) fragmentation with supplemental activation (energy 20). The resolution of the survey scan was 120 000 (400 m/z) with a target value of 1×106 ions, one microscan and maximum injection time of 200 ms. Low resolution CID or ETD MS/MS spectra were acquired with a target value of 10 000 ions in rapid scan mode with m/z range adjusted according to actual precursor mass and charge. MS/MS acquisition in the linear ion trap was carried out in parallel to the survey scan in the Orbitrap analyser by using the preview mode. The maximum injection time for MS/MS was 50 ms. ETD reaction time was 100 ms (doubly charged precursors, adjusted according to charge state). Dynamic exclusion was enabled for 45 s after one MS/MS spectra acquisition Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 5 and early expiration was disabled. The isolation window for MS/MS fragmentation was set to 2 m/z. Database searching. The processed data were searched with MASCOT (version 2.2 or higher, Matrix Science, Boston, Massachusetts, USA) against NCBInr database (non-redundant, taxonomic restriction Mammalia) with settings corresponding to trypsin/chymotrypsin specifity (two miss-cleavages allowed) and optional modifications: oxidation (M), carbamidomethylation (C), phosphorylation (S, T, Y). Mass tolerances for peptides and MS/MS fragments were 0.5 Da in case of HCT Ultra system (correction for one 13 C atom) and 5 ppm and 0.5 Da, respectively, in case of Orbitrap system.The significance threshold was set to p < 0.05 and p < 0.01 for HCT Ultra and Orbitrap Elite generated data, respectively. All phosphorylation sites were manually confirmed in profile MS/MS spectra for generation of phosphorylation variants, when necessary. phoshoRS feature was used for phosphorylation localization in case of Orbitrap Elite data. RESULTS MS/MS-based identification of Dvl3 phosphorylation in the presence and absence of CK1ε - CK1ε (together with the closely related CK1δ) is the only Dvl kinase, which is both required and sufficient for the shift of Dvl electrophoretical mobility (Fig. 1A) and activation of the Wnt/-catenin pathway (eg. analyzed using the TopFlash reporter (35)) at the same time (Fig. 1B). Other kinases such as CK2, PAR1, RIPK4, PKC or Abl known to bind/activate Dvl upon overexpression do not lead to this phenotype (11,16,36-40). In order to identify Dvl3 phosphorylation sites controlled by CK1ε, we overexpressed FLAG-tagged human Dvl3 in HEK293t cells alone or together with CK1ε. Overexpressed Dvl3 was immunoprecipitated, separated on 1D SDSPAGE and stained using Coomassie blue (for typical gel see Fig. 1C). Bands were then excised from the gel and phosphorylation was analyzed by LC-MS/MS. Identification of phosphorylations in Dvl3/Dvl3+CK1ε samples was repeated in seven independent experiments. All significantly identified phosphorylation sites are summarized in Table 1 and Figure 2A. Out of more than fifty detected phosphorylation sites, only 9 were found uniquely in samples activated by CK1ε (Fig. 2, in red). Additional 7 serines were detected more frequently in presence of CK1ε, but were also occasionally seen in the control samples (Fig. 2, in orange). We propose that phosphorylation of these 16 sites contribute to the Dvl3-associated phenotypes induced by CK1ε. Interestingly, phosphorylation of residues S175, S192, S197, S202 and S203 in the region between DIX and PDZ domain, which is rich in basic aminoacids (the basic region), of Dvl3 was less abundant after CK1ε coexpression. These data suggest that some Dvl phosphorylation events may be inhibitory and mutually exclusive with others. Phosphorylation of the remaining 30 sites (Fig. 2A, indicated in blue) was not affected by CK1ε and seems to be constitutive in the used culture conditions. Most of the phosphorylation sites were found on residues conserved among Dvl isoforms and species (Dvl1, Dvl2 and Dvl3 of Homo sapiens, Mus musculus, Xenopus laevis and Xenopus tropicalis were considered for the analysis) but some were unique for Dvl3 – this was the case of S611, S612, S633 S636, S639, S642 and S643 in the His-rich region of the Dvl3 C-terminus. For complete information about all identified phosphorylation sites see Table1. The MS/MS-based analysis demonstrated a complex pattern of Dvl3 phosphorylation and showed that many phosphorylation events take place independently of exogenous CK1. In order to classify individual phosphorylated motives, we searched the sequence of Dvl3 for substrate motives of known Dvl kinases, specifically CK1, CK2, protein kinase C (PKC) and Polo-like kinase 1 (PLK1) using GPS 2.1 software (34). High stringency searches did not result in the prediction of kinases responsible for all phosphorylation sites; we thus present in Table 1 the results of searches with lower stringency. Surprisingly, this analysis showed that only two (S61, S211) out of 16 phosphorylation events induced by CK1 (indicated in orange or red in Fig. 2) appeared in a canonical CK1 motif (see Table 1). Moreover, based on the software prediction, sequences containing S280, S350 and S636 are not recognized by any of the considered Dvl kinases (CK1, CK2, PKC and PLK1). This suggests that either CK1 recognizes atypical motives or alternatively that its phosphorylation Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 6 sites are primed for subsequent phosphorylation by so far unidentified kinases. Our analysis does not provide definitive answer to that question. Majority of the identified phosphorylated residues were found outside of the DIX, PDZ and DEP domains (Fig. 2A). The only exceptions from this rule are infrequently detected T15, S48 and S61 in the DIX and CK1ε-induced S280 and S311 in the PDZ domain. These three structured regions are linked by intrinsically unstructured regions (see PONDR-FIT prediction for hDvl3 in Fig. 2B) (33), which contained almost all phosphorylation sites (Table 1 and Fig. 2A). We have thus tested in silico whether or not the phosphorylations change the secondary structure and folding of the protein. Mimicking phosphorylation by changing phosphorylated S/T residues in the unstructured regions for aspartic acid (D) (Dvl3 S-D) resulted in a slight increase of the score for the unstructured regions (Fig. 2B). However, the overall distribution of structured/unstructured regions was not massively affected and it is unlikely that even robust Dvl3 phosphorylation at multiple sites affects the secondary structure of the protein. However, we cannot exclude minor changes or phosphorylation-induced formation of novel tertiary contacts. Basal Dvl-induced TCF/LEF-driven transcription is significantly reduced by the mutation of PDZ domain residues S280 and S311 - In order to study the function of the repeatedly identified (detected more than once) phosphorylations, we decided to mutate the S/Ts, which were selected based on the following priority criteria: (i) unknown function, (ii) induction/repression by CK1ε, (iii) sequence conservation, and/or (iv) presence in the structured regions with the well-defined function. In the first step we functionally analyzed the phosphorylation of two individual serines: S280 and S311. These fully conserved residues are present in the central part of the PDZ domain, which is critically required for the function in the Wnt/β-catenin pathway. We mutated S280 and S311 of human Dvl3 to either alanine (A), to prevent phosphorylation or to glutamic acid (E), to mimic constitutive phosphorylation and tested the ability of these mutants to activate the Wnt/β-catenin pathway by TopFlash assay. As we show in Fig. 3A, Wnt/β-catenin pathway induction by Dvl3 (S280A) was decreased whereas induction by Dvl3 (S280E) was increased when compared to wt Dvl3. S280 is conserved among all Dvl isoforms. In order to test the level of conservation of S280 function, we have mutated corresponding serines in Dvl1 (S282A) and Dvl2 (S298A). As shown in Fig. 3B, the mutations of serines corresponding to S280 (hDvl3) to alanine in Dvl1 and Dvl2 also significantly reduced their ability to promote downstream signaling. These data suggest that phosphorylation of S280 is required for the efficient Dvl activity in Wnt/β-catenin pathway across the whole Dvl family. In contrast to S280 both phospho-mimicking and phospho-preventing mutation of S311 (S311A or S311E) resulted in a decreased activation of Wnt/β-catenin pathway. The function of S311 phosphorylation, however, seems to be synergistic with S280 as shown by the analysis double mutants. Mutation of S280/S311 to A or E almost completely abolished the activation of Wnt/β-catenin pathway by Dvl3 (Fig. 3A). Interestingly, when we co-expressed CK1ε (Fig. 3C) we have not observed any significant differences among individual mutants although S280A and S280/311A mutants were slightly hypoactive whereas S280E was hyperactive. Data in Fig. 3A-C suggested an important role of S280 and S311 for the downstream canonical Wnt signaling. Specifically, phosphorylation at S280 and subsequent electrostatic changes might be relevant because S280A decreases whereas phospho-mimicking S280E increases the signaling efficiency of Dvl3. The observed functional phenotype might be explained by the fact that the S280 directly borders with binding area common to several proteins known to modulate Wnt signaling, namely Idax (41), Fzd7 (42), and TMEM88 (43) (Fig. 3D). Considering that the binding area for these proteins is predominantly positively charged, the introduction of a strong negative charge at its border is expected to produce notable impact on binding affinity between the Dvl PDZ domain and abovementioned proteins (Fig. 3E). Phosphorylation of residues in the Dvl3 C-terminus causes the electrophoretic mobility changes induced by CK1ε but not by Fzd5 - As the second step we have mutated additional four clusters of S/Ts, which were chosen based on the criteria defined above. The Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 7 mutated residues included the cluster of S202/S203/S204/T205 from the basic region (Cluster1; C1) and three C-terminally located clusters S630/S633/S636 (Cluster2; C2), S639/S642/S643 (Cluster3; C3), and S611/S612 (Cluster4; C4) (Fig. 4A). The cluster C1 was selected because the level of phosphorylation of this region decreased after CK1ε co-expression and these sites are conserved in all considered Dvl isoforms. The clusters C2, C3 and C4 are conserved only in Dvl3 but phosphorylation of these residues was strongly induced after CK1ε co-expression. The four clusters of residues were mutated to alanine. Dvl can be detected on Western blot as two bands designated Dvl and P (phosphorylated)Dvl. Co-expression of CK1ε promotes the formation of novel even more slowly migrating form of Dvl3 named PS-(phosphorylated and shifted)-Dvl (see Fig. 1A), which corresponds to endogenous Dvl form induced by Wnts (8,9,16,30). When we analyzed individual mutants, we found that they show rather comparable electrophoretic mobility (Fig. 4B). The only exception was Dvl3-C1-SA with electrophoretic migration slightly more retarded than that of wt. This is in agreement with original MS/MS data, which suggest that phosphorylation of C1 residues opposes the effects of CK1ε. In contrast, co-expression of CK1ε led to the formation of PS-Dvl3 in all Dvl3 variants with the exception of Dvl3-C2, C2+C3 and C2+C3+C4 SA mutants (Fig. 4C). The effects are not due to decreased ability of CK1ε to bind to Dvl3 because wt Dvl3 and its mutants C1 S-A and C2+3+4 S-A were able to bind to CK1ε with similar efficiency (Fig. 4D). The ability to activate TCF/LEF-driven transcription (Topflash), which reflects the ability to activate Wnt/β-catenin pathway, was not significantly affected in any of the cluster mutants (Fig. 4E). These observations suggest that residues in clusters C2, C3 and C4 are required for PS-Dvl formation but not for Dvl3 activity in the Wnt/β-catenin signaling. It has been shown that not only CK1ε but also Frizzled (Fzd) overexpression can lead to the electrophoretic mobility shift of Dvl. Wnt receptors Fzd are known to recruit Dvl to cytoplasmatic membrane and to trigger Dvl phosphorylation (5,29). In the next step we thus tested the ability of Fzd5, selected for this study, to control electrophoretic migration of Dvl3 mutants. Although Fzd5 was able to induce phosphorylation-dependent shift of Dvl3, the shift was never as prominent as PSDvl3 induced by CK1ε (Fig. 4F). Moreover, the intensity of Dvl3 phosphorylation was comparable in all Dvl3 mutants including Dvl3 C2+3+4 S-A (Fig. 4F). This opened the surprising possibility that CK1ε and Fzd5induced phosphorylation is qualitatively different. In order to test this possibility, we introduced to hDvl3 K435M mutation in the DEP domain, which corresponds to fly Dsh1 mutant (K417 in dDsh), which cannot interact with Frizzled (5). Surprisingly, shift of Dvl3K435M caused by CK1ε is comparable if not identical to that of wt Dvl3 (Fig. 4C) whereas Fzd5 is not able to trigger any detectable change in the mobility of Dvl3-K435M, in striking contrast to other mutants used in this study (Fig. 4F). This suggests that the phosphorylation events triggered by Fzd5 and CK1ε are not identical. Phospho-preventing mutations in Dvl3 Cterminus interfere with CK1ε-induced but not Fzd5-induced changes in Dvl3 subcellular localization - Dvl3 is usually found in dynamic multiprotein aggregates (44,45) called Dvl dots or puncta. After coexpression of CK1ε, Dvl intracellular localization changes from punctate to even (for example see Fig. 5A); a phenomenon, which is associated with decreased polymerization of Dvl molecules (16). In contrast, high levels of Fzd receptor or Fzd co-expression (5,29,46) lead to membrane recruitment of Dvl (Fig. 5A). In the next step we thus tested whether Dvl3 S/T-A mutants affect subcellular localization of Dvl3. As we show in Fig. 5B all Dvl3 mutants showed in HEK293t predominantly punctate localization pattern. Interestingly, following overexpression of CK1ε, C1 mutant and K435M acquired even localization similarly to wt Dvl3 whereas C2, C2+3 and C2+3+4 S-A mutants showed an additive deficit in their ability to achieve even subcellular distribution. Localization of C2+C3+C4 S-A mutant in presence and absence of overexpressed CK1ε was indistinguishable. Mutations within PDZ domain showed no detectable effect on localization pattern of Dvl3 following CK1ε overexpression. In contrast to CK1ε, none of the tested Dvl3 S-A mutants showed a deficit in the Fzd5mediated recruitment of Dvl3 to the plasma membrane (Fig. 5B). Dvl3 (K435M), which is analogous to Drosophila dDsh mutant Dsh1 (5) Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 8 and is expected to be deficient in membrane recruitment by Fzd, however, clearly failed to be recruited to plasma membrane by Fzd5 coexpression (Fig. 5B). This suggests that residues phosphorylated in response to CK1ε and mutated in this study are not required for Fzd5-induced membrane localization of Dvl3. In combination with data in Fig. 4 we propose that membrane recruitment of Dvl3 is required for phosphorylation events induced by Fzd5 but not by CK1ε. Phospho-S643-antibody recognizes CK1εphosphorylated and evenly distributed Dvl3 To get better insight into the subcellular localization of individual phospho pools of Dvl3 we have raised a phospho-specific antibody recognizing phosphorylated S643. We have selected phospho-serine 643 as a candidate epitope because it belongs to the cluster of residues phosphorylated after CK1ε coexpression, and it is the last phosphorylated residue in the cluster of pS633, pS636, pS639, pS642 and pS643. As we demonstrate on Western blotting anti-pS643-Dvl3 antibody strongly detected Dvl3 co-expressed with CK1ε, whereas it almost did not recognize FLAG-Dvl3 expressed alone or in combination with Fzd5 (Fig. 5C). Anti-pS643-Dvl3 antibody failed to detect Dvl3 (C2+3 SA) mutant where S643 is mutated to A643, which demonstrates that it is specific. Next we analyzed subcellular distribution of pS643-Dvl3 by immunocytochemistry in order to identify the pool of Dvl3 specifically modified by phosphorylation in the C-terminus. The staining showed a weak signal that was dramatically boosted by CK1ε overexpression causing even localization of Dvl3 (Fig. 5D). Importantly, Fzd5 overexpression leading to Dvl3 membrane recruitment did not increase the pS643-Dvl3 signal. This suggests that membranous, Fzd5-recruited, Dvl3 is not modified by CK1ε at S643. This observation correlates well with the analysis of Dvl3 electrophoretic shift (Fig. 4C) and demonstrates that phosphorylation of C2/C3/C4 residues is part of the mechanism, which promotes even localization of Dvl3 and is distinct from the signaling events induced by Fzd5. DISCUSSION In this study, we performed unbiased proteomic identification of Dvl3 phosphorylation pattern induced by CK1ε. Previous studies in various experimental systems identified several phosphorylation sites spread throughout the structure of Dvl protein (18,20-22). Combination of all data clearly demonstrates that phosphorylation of Dvl protein is extensive and the number of phosphorylated sites extends 50. However, it is likely that the number of phosphorylated residues on a single Dvl molecule is lower because MS/MS analysis cannot distinguish individual differentially modified Dvl pools. This assumption is supported by our observation, that despite general increase in the number of phosphorylated residues, some phosphorylation sites in the basic region of Dvl3 disappeared after CK1 co-expression. To our surprise, the majority of the detected phosphorylation events were not promoted by CK1. This suggests that Dvl3 gets heavily phosphorylated prior to Wnt-induced CK1mediated activation. This is in good agreement with our previous work where we show that CK2 mediates such basal state of Dvl phosphorylation, which is then required for dynamic phosphorylation by CK1ε and for further downstream signaling in the Wnt/catenin pathway (16). Several kinases including PKC, CK2, PLK1, RIPK4, and Abl have been implicated previously in Dvl binding and phosphorylation (11,19,37-40) and our in silico analysis of phosphorylation sites supports possible involvement of these kinases in the basal state Dvl phosphorylation. Our findings support earlier assumption that CK1ε has a dual function in Dvl biology (16). It seems that CK1ε acts as (i) the activator of downstream Wnt/-catenin signaling via phosphorylation of distinct S/T residues (phosphorylation of S280/S311 in the PDZ domain) as well as (ii) the inactivating kinase affecting Dvl polymerization via phosphorylation of the residues (C2+C3+C4) in the C-terminus of Dvl3. The role of Dvl Cterminus in the Wnt/β-catenin signaling was established by our previous work, where we proposed that it acts as the CK1-controlled negative regulator (6,16). Here we identified specific residues phosphorylated by CK1ε in the C-terminus of Dvl3 (clusters C2, C3 and C4). We demonstrate that these residues control PS-Dvl formation and are critical for CK1εinduced changes in Dvl3 subcellular localization. Although the clusters C2, C3 and Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 9 C4 are conserved only in Dvl3, multiple S/T residues are present also in the C-terminal region of Dvl1 and Dvl2. This opens the possibility that function of phosphorylated Cterminus is conserved despite the differences in the primary sequence. This possibility is supported by recent findings by GonzalezSancho and colleagues (17), which demonstrated that formation of PS-Dvl2 and its even localization depends on other residues of hDvl2 C-terminus. The residues are conserved in Dvl3 (S578 and S581 in hDvl3) but were found constitutively phosphorylated in our study. Our data support the line of recent evidence that Dvl C-terminus and the phosphorylation therein controls the activity of Dvl in Wnt/βcatenin pathway. It was reported that Dvl Cterminus binds several components of the noncanonical Wnt pathway such as NFAT (47) or Ror2 (6). This together with other reports (48) suggests that Dvl C-terminus can critically contribute to several branches of non-canonical Wnt signaling. In line with these observations, it was recently shown that the phosphorylation of the residues in the C-terminus of Dvl2 necessary for PS-Dvl2 formation is required for Dvl2-dependent neurite outgrowth in TC-32 cells, a process not mediated by Wnt/β-catenin pathway (17). It is not well understood how Dvl C-terminus exerts its function, however recent report proposes that C-terminus can intramolecularly interact with the PDZ domain (49). Phosphorylation in the C-terminus can then affect the intramolecular conformation similarly to the regulatory mechanism well described among protein kinases. This possibility is supported by earlier analysis of the interaction of Ror2 and Dvl3, which showed that Ror2 could efficiently recognize only PSDvl3 or Dvl3 C-terminus as such (6). Most comprehensive study to date, which attempted to identify Dvl phosphorylation, was performed in Drosophila model and used activation of Dvl by Frizzled overexpression (18). This study concluded that S/T phosphorylation of Dvl has no role neither in PCP nor canonical Wnt signaling in fly. In difference to that study we tested phosphorylation of Dvl3 by overexpression of the most relevant Dvl kinase – CK1ε. To our surprise, when we mutated the most dynamically phosphorylated Dvl3 residues in the C-terminus we found that they are deficient only in CK1ε-induced but not in Fzd5-induced mobility shift or subcellular localization changes. These distinct molecular features of CK1ε- and Fzd5-induced events were further confirmed by implementing anti-pS643-Dvl3 phospho-specific antibody, which recognized only Dvl3 modified by CK1ε but not by Fzd5. This data suggest that CK1ε is not directly downstream of Fzd receptor but rather represents independent, possibly parallel, molecular mechanism of activation/deactivation of Dvl. In summary, our findings shed more light onto complex molecular events on the level of Dvl3 phosphorylations triggered by CK1ε. On the single aminoacid level we show that CK1εmediated phoshorylation in the PDZ domain Ser280 and Ser311 is one of the steps in the chain of events leading to the Wnt/-catenin downstream activation. On the contrary, phosphorylations in the C-terminus contribute to the change of Dvl3 electrophoretic mobility and mediate negative effects of CK1ε on Dvl polymerization, which were postulated to be part of the negative feedback loop. REFERENCES: 1. Clevers, H. (2006) Wnt/beta-catenin signaling in development and disease. Cell 127, 469-480 2. MacDonald, B. T., Tamai, K., and He, X. (2009) Wnt/beta-catenin signaling: components, mechanisms, and diseases. Dev Cell 17, 9-26 3. Schwarz-Romond, T., Fiedler, M., Shibata, N., Butler, P. J., Kikuchi, A., Higuchi, Y., and Bienz, M. (2007) The DIX domain of Dishevelled confers Wnt signaling by dynamic polymerization. Nat Struct Mol Biol 14, 484-492 4. Gao, C., and Chen, Y. G. (2010) Dishevelled: The hub of Wnt signaling. Cell Signal 22, 717- 727 Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 10 5. Axelrod, J. D., Miller, J. R., Shulman, J. M., Moon, R. T., and Perrimon, N. (1998) Differential recruitment of Dishevelled provides signaling specificity in the planar cell polarity and Wingless signaling pathways. Genes Dev 12, 2610-2622 6. Witte, F., Bernatik, O., Kirchner, K., Masek, J., Mahl, A., Krejci, P., Mundlos, S., Schambony, A., Bryja, V., and Stricker, S. (2010) Negative regulation of Wnt signaling mediated by CK1-phosphorylated Dishevelled via Ror2. Faseb J 24, 2417-2426 7. Yanagawa, S., van Leeuwen, F., Wodarz, A., Klingensmith, J., and Nusse, R. (1995) The dishevelled protein is modified by wingless signaling in Drosophila. Genes Dev 9, 1087-1097 8. Bryja, V., Schulte, G., Rawal, N., Grahn, A., and Arenas, E. (2007) Wnt-5a induces Dishevelled phosphorylation and dopaminergic differentiation via a CK1-dependent mechanism. J Cell Sci 120, 586-595 9. Gonzalez-Sancho, J. M., Brennan, K. R., Castelo-Soccio, L. A., and Brown, A. M. (2004) Wnt proteins induce dishevelled phosphorylation via an LRP5/6- independent mechanism, irrespective of their ability to stabilize beta-catenin. Mol Cell Biol 24, 4757-4768 10. Bryja, V., Schulte, G., and Arenas, E. (2007) Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate bcatenin. Cell. Signal. 19, 610-616 11. Peters, J. M., McKay, R. M., McKay, J. P., and Graff, J. M. (1999) Casein kinase I transduces Wnt signals. Nature 401, 345-350 12. Sakanaka, C., Leong, P., Xu, L., Harrison, S. D., and Williams, L. T. (1999) Casein kinase iepsilon in the wnt pathway: regulation of beta-catenin function. Proc Natl Acad Sci U S A 96, 12548-12552 13. Kishida, M., Hino, S., Michiue, T., Yamamoto, H., Kishida, S., Fukui, A., Asashima, M., and Kikuchi, A. (2001) Synergistic activation of the Wnt signaling pathway by Dvl and casein kinase Iepsilon. J Biol Chem 276, 33147-33155 14. Cong, F., Schweizer, L., and Varmus, H. (2004) Casein kinase Iepsilon modulates the signaling specificities of dishevelled. Mol Cell Biol 24, 2000-2011 15. Foldynova-Trantirkova, S., Sekyrova, P., Tmejova, K., Brumovska, E., Bernatik, O., Blankenfeldt, W., Krejci, P., Kozubik, A., Dolezal, T., Trantirek, L., and Bryja, V. (2010) Breast cancer-specific mutations in CK1epsilon inhibit Wnt/beta-catenin and activate the Wnt/Rac1/JNK and NFAT pathways to decrease cell adhesion and promote cell migration. Breast Cancer Res 12, R30 16. Bernatik, O., Ganji, R. S., Dijksterhuis, J. P., Konik, P., Cervenka, I., Polonio, T., Krejci, P., Schulte, G., and Bryja, V. (2011) Sequential activation and inactivation of Dishevelled in the Wnt/beta-catenin pathway by casein kinases. J Biol Chem 286, 10396-10410 17. Gonzalez-Sancho, J. M., Greer, Y. E., Abrahams, C. L., Takigawa, Y., Baljinnyam, B., Lee, K. H., Lee, K. S., Rubin, J. S., and Brown, A. M. (2013) Functional consequences of Wntinduced dishevelled 2 phosphorylation in canonical and noncanonical Wnt signaling. J Biol Chem 288, 9428-9437 18. Yanfeng, W. A., Berhane, H., Mola, M., Singh, J., Jenny, A., and Mlodzik, M. (2011) Functional dissection of phosphorylation of Disheveled in Drosophila. Dev Biol 360, 132-142 19. Kikuchi, K., Niikura, Y., Kitagawa, K., and Kikuchi, A. (2010) Dishevelled, a Wnt signalling component, is involved in mitotic progression in cooperation with Plk1. EMBO J 29, 3470- 3483 20. Klimowski, L. K., Garcia, B. A., Shabanowitz, J., Hunt, D. F., and Virshup, D. M. (2006) Site-specific casein kinase 1epsilon-dependent phosphorylation of Dishevelled modulates beta-catenin signaling. The FEBS journal 273, 4594-4602 21. Klein, T. J., Jenny, A., Djiane, A., and Mlodzik, M. (2006) CKIepsilon/discs overgrown promotes both Wnt-Fz/beta-catenin and Fz/PCP signaling in Drosophila. Curr Biol 16, 1337- 1343 22. Wu, C., Wei, W., Li, C., Li, Q., Sheng, Q., and Zeng, R. (2012) Delicate analysis of posttranslational modifications on Dishevelled 3. Journal of proteome research 11, 3829-3837 23. Yokoyama, N., and Malbon, C. C. (2009) Dishevelled-2 docks and activates Src in a Wntdependent manner. J Cell Sci 122, 4439-4451 Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 11 24. Lee, K. H., Johmura, Y., Yu, L. R., Park, J. E., Gao, Y., Bang, J. K., Zhou, M., Veenstra, T. D., Yeon Kim, B., and Lee, K. S. (2012) Identification of a novel Wnt5a-CK1varepsilonDvl2-Plk1-mediated primary cilia disassembly pathway. EMBO J 31, 3104-3117 25. Angers, S., Thorpe, C. J., Biechele, T. L., Goldenberg, S. J., Zheng, N., MacCoss, M. J., and Moon, R. T. (2006) The KLHL12-Cullin-3 ubiquitin ligase negatively regulates the Wnt-betacatenin pathway by targeting Dishevelled for degradation. Nat Cell Biol 8, 348-357 26. Tauriello, D. V., Haegebarth, A., Kuper, I., Edelmann, M. J., Henraat, M., Canninga-van Dijk, M. R., Kessler, B. M., Clevers, H., and Maurice, M. M. (2010) Loss of the tumor suppressor CYLD enhances Wnt/beta-catenin signaling through K63-linked ubiquitination of Dvl. Mol Cell 37, 607-619 27. Narimatsu, M., Bose, R., Pye, M., Zhang, L., Miller, B., Ching, P., Sakuma, R., Luga, V., Roncari, L., Attisano, L., and Wrana, J. L. (2009) Regulation of planar cell polarity by Smurf ubiquitin ligases. Cell 137, 295-307 28. McKay, R. M., Peters, J. M., and Graff, J. M. (2001) The casein kinase I family in Wnt signaling. Dev Biol 235, 388-396 29. Tauriello, D. V., Jordens, I., Kirchner, K., Slootstra, J. W., Kruitwagen, T., Bouwman, B. A., Noutsou, M., Rudiger, S. G., Schwamborn, K., Schambony, A., and Maurice, M. M. (2012) Wnt/beta-catenin signaling requires interaction of the Dishevelled DEP domain and C terminus with a discontinuous motif in Frizzled. Proc Natl Acad Sci U S A 109, E812-820 30. Bryja, V., Schulte, G., and Arenas, E. (2007) Wnt-3a utilizes a novel low dose and rapid pathway that does not require casein kinase 1-mediated phosphorylation of Dvl to activate beta-catenin. Cell Signal 19, 610-616 31. Bryja, V., Cajanek, L., Pachernik, J., Hall, A. C., Horvath, V., Dvorak, P., and Hampl, A. (2005) Abnormal development of mouse embryoid bodies lacking p27Kip1 cell cycle regulator. Stem Cells 23, 965-974 32. Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E. C., and Ferrin, T. E. (2004) UCSF Chimera--a visualization system for exploratory research and analysis. J Comput Chem 25, 1605-1612 33. Xue, B., Dunbrack, R. L., Williams, R. W., Dunker, A. K., and Uversky, V. N. (2010) PONDR-FIT: a meta-predictor of intrinsically disordered amino acids. Biochim Biophys Acta 1804, 996-1010 34. Xue, Y., Ren, J., Gao, X., Jin, C., Wen, L., and Yao, X. (2008) GPS 2.0, a tool to predict kinase-specific phosphorylation sites in hierarchy. Mol Cell Proteomics 7, 1598-1608 35. Korinek, V., Barker, N., Morin, P. J., van Wichen, D., de Weger, R., Kinzler, K. W., Vogelstein, B., and Clevers, H. (1997) Constitutive transcriptional activation by a betacatenin-Tcf complex in APC-/- colon carcinoma. Science 275, 1784-1787 36. Ossipova, O., Dhawan, S., Sokol, S., and Green, J. B. (2005) Distinct PAR-1 proteins function in different branches of Wnt signaling during vertebrate development. Dev Cell 8, 829-841 37. Willert, K., Brink, M., Wodarz, A., Varmus, H., and Nusse, R. (1997) Casein kinase 2 associates with and phosphorylates dishevelled. Embo J 16, 3089-3096 38. Singh, J., Yanfeng, W. A., Grumolato, L., Aaronson, S. A., and Mlodzik, M. (2010) Abelson family kinases regulate Frizzled planar cell polarity signaling via Dsh phosphorylation. Genes Dev 24, 2157-2168 39. Huang, X., McGann, J. C., Liu, B. Y., Hannoush, R. N., Lill, J. R., Pham, V., Newton, K., Kakunda, M., Liu, J., Yu, C., Hymowitz, S. G., Hongo, J. A., Wynshaw-Boris, A., Polakis, P., Harland, R. M., and Dixit, V. M. (2013) Phosphorylation of Dishevelled by protein kinase RIPK4 regulates Wnt signaling. Science 339, 1441-1445 40. Kinoshita, N., Iioka, H., Miyakoshi, A., and Ueno, N. (2003) PKC delta is essential for Dishevelled function in a noncanonical Wnt pathway that regulates Xenopus convergent extension movements. Genes Dev 17, 1663-1676 41. London, T. B., Lee, H. J., Shao, Y., and Zheng, J. (2004) Interaction between the internal motif KTXXXI of Idax and mDvl PDZ domain. Biochem Biophys Res Commun 322, 326-332 42. Wong, H. C., Bourdelas, A., Krauss, A., Lee, H. J., Shao, Y., Wu, D., Mlodzik, M., Shi, D. L., and Zheng, J. (2003) Direct binding of the PDZ domain of Dishevelled to a conserved internal sequence in the C-terminal region of Frizzled. Mol Cell 12, 1251-1260 Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 12 43. Lee, H. J., Finkelstein, D., Li, X., Wu, D., Shi, D. L., and Zheng, J. J. (2010) Identification of transmembrane protein 88 (TMEM88) as a dishevelled-binding protein. J Biol Chem 285, 41549-41556 44. Schwarz-Romond, T., Merrifield, C., Nichols, B. J., and Bienz, M. (2005) The Wnt signalling effector Dishevelled forms dynamic protein assemblies rather than stable associations with cytoplasmic vesicles. J Cell Sci 118, 5269-5277 45. Smalley, M. J., Signoret, N., Robertson, D., Tilley, A., Hann, A., Ewan, K., Ding, Y., Paterson, H., and Dale, T. C. (2005) Dishevelled (Dvl-2) activates canonical Wnt signalling in the absence of cytoplasmic puncta. J Cell Sci 118, 5279-5289 46. Schulte, G., and Bryja, V. (2007) The Frizzled family of unconventional G-protein-coupled receptors. Trends Pharmacol Sci 28, 518-525 47. Huang, T., Xie, Z., Wang, J., Li, M., Jing, N., and Li, L. (2011) Nuclear factor of activated T cells (NFAT) proteins repress canonical Wnt signaling via its interaction with Dishevelled (Dvl) protein and participate in regulating neural progenitor cell proliferation and differentiation. J Biol Chem 286, 37399-37405 48. Ma, L., Wang, Y., Malbon, C. C., and Wang, H. Y. (2010) Dishevelled-3 C-terminal His single amino acid repeats are obligate for Wnt5a activation of non-canonical signaling. J Mol Signal 5, 19 49. Smietana, K., Mateja, A., Krezel, A., and Otlewski, J. (2011) PDZ domain from Dishevelled -- a specificity study. Acta Biochim Pol 58, 243-249 ACKNOWLEDGMENTS We would like to thank Randy Moon, Madelon Maurice, Jeff Wrana, Jonathan M. Graff and S. Yanagawa for providing plasmids. We would also like to thank Hana Pecuchova and Lenka Bryjova for excellent technical assistance and members of the lab for inspirational discussions about the results. This study was supported by the grants from the Ministry of Education, Youth, and Sports of the Czech Republic (MSM0021622430), and by the Czech Science Foundation (204/09/0498; 13- 31488P). The MS/MS analysis was supported by the project “CEITEC - Central European Institute of Technology” (CZ.1.05/1.1.00/02.0068) from European Regional Development Fund and by CEITEC Open Access project (LM2011020), funded by the Ministry of Education, Youth and Sports of the Czech Republic under the activity „Projects of major infrastructures for research, development and innovations”). LT thanks for support from VIDI - Career Development (Netherland’s Organisation for Scientific Research. Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 13 FIGURE LEGENDS Figure 1: Co-expression of CK1ε with FLAG-Dvl3 retards electrophoretic migration and induces phosphorylation-dependent shift of Dvl (PS-Dvl3) (A) and induces TCF/LEF dependent transcription as shown by TopFlash assay (B). C: Representative example of Coomassie Brilliant Blue R-stained gel used for MS/MS analysis. See details in Material and methods. Figure 2: (A) Dvl3 phosphorylation (summary of MS/MS data). Position and dynamics of detected phosphorylations is indicated by color coded lines: green – less frequent in CK1ε induced samples, blue – constitutive, orange – more often in CK1ε-induced samples, red – only in CK1ε induced samples. The yellow/violet lines indicate hDvl3 protein coverage in the MS/MS analysis (yellow – significant identification, violet – insignificant identification). C1, C2, C3, C4, S280 and S311 labels indicate the position of residues mutated in the study. (B) Secondary structure of Dvl3 as predicted by PONDR-FIT software (upper panel). Structured regions are indicated by values <0.5, unstructured by values >0.5. Accuracy of the prediction is confirmed by the identification of DIX, PDZ and DEP domains as regions with secondary structure. In silico mutation of all identified phosphorylated S/T sites in basic region and C-terminus for D did not lead to the massive changes in the secondary structure (lower panel). Figure 3: Phosphopreventing mutations of S280/S311 in the PDZ domain block activation of the Wnt/β-catenin pathway. (A-C): HEK 293 cells were transfected with indicated Dvl3 constructs, TopFlash and Renilla and CK1ε. (A) Mutations of S280 and S311 prevent efficient activation of Wnt/β-catenin by Dvl3. (B) The function of S280 is conserved in Dvl1 and Dvl2. Dvl1-S282A and Dvl2-S298A, corresponding to Dvl3 S280A, activated Wnt/β-catenin pathway less efficiently than wt Dvl1 and Dvl2 respectively. (C) In CK1ε treated samples no significant differences between the wt Dvl3 and the mutant Dvl3 were detected. Samples were analyzed by One-way ANOVA followed by Tukey post tests. * p<0.05, *** p<0.001, number of experiments ≥4. (D) Interaction sites of Idax (in yellow), Fzd7 (in blue) and Tmem88 (in green) proteins determined for mouse Dvl1 PDZ domain (PDB ID 1MC7). Phylogentically conserved S280 and S311 phosphorylation sites in human Dvl3 PDZ are shown in red. (E) Left panel: Schematic representation of secondary structure of mouse Dvl1 PDZ domain. Central panel: Modeled electrostatic surface potential of mDvl1 PDZ. Right panel: Electrostatic surface potential of mDvl1 PDZ bearing phosphomimicking mutations at positions corresponding to S280 and S311 of human Dvl3 PDZ. Phosphorylation sites are highlighted in magenta (magenta arrow). Calculations of electrostatic surface potential were performed in UCSF Chimera software Figure 4: Functional analysis of phosphorylation sites in clusters C1-C4. (A) Representation of Dvl structure with the positions of mutated residues. (B,C) HEK 293 cells were transfected with the indicated plasmids, and the electrophoretic migration of Dvl3 in absence (B) and presence (C) of CK1 ε was analyzed by Western blotting with anti-Flag antibody. Mutations in C1 decreased Dvl3 mobility (B) whereas mutations in C2, C2+3, C2+3+4 blocked PS-Dvl3 induction by CK1ε (C). The protein expression levels of individual mutants were comparable. (D) The ability of individual Dvl3 mutants to interact with CK1ε was tested by immunoprecipitation of CK1ε and Flag-Dvl3 and subsequent analysis by Western blotting. (E) HEK293t cells were transfected with the indicated plasmids and the ability to activate TCF/LEF-dependent transcription was assessed by TopFlash system. None of the C1, C2, C2+3, C2+3+4 mutants showed any statistically significant difference. Statistical differences were analyzed by One-way ANOVA followed by Tukey post tests. Number of experiments n≥3. (F) HEK293t cells were transfected with the indicated plasmids, and the electrophoretic mobility of Dvl3 Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 14 after V5-Fzd5 coexpression was analyzed by Western blotting. With the exception of Dvl3-K435M electrophoretic mobility of all Dvl3 mutants was indistinguishable from wild type. Figure 5: Analysis of subcellular localization of individual Dvl pools. (A) Representation of typical distribution patterns of Dvl3 – punctate (i), even (ii), and membrane (iii). (B) HEK293 cells were transfected with corresponding plasmids, fixed and Dvl3 was stained with anti-FLAG antibody. Distribution pattern of Dvl3 was analyzed in at least 100 cells. CK1ε is unable to promote even localization of Dvl3 in C2, C3 and C4 S-A mutants, whereas V5-Fzd5 changes the intracellular distribution of Dvl3 to membranous in all tested constructs with the exception of Dvl3-K435M, which served as positive control. (C, D) Phospho-S643-Dvl3 was detected using anti-pS643-Dvl3 specific antibody by Westen blotting (C) or immunocytochemistry (D). Total Dvl3 was detected by anti-FLAG antibody. (C) Anti-pS643-Dvl3 specific antibody does not detect Dvl3 C2+C3 S-A mutant, but it strongly recognizes Dvl3 co-expressed with CK1ε. (D) The signal of anti-pS643-Dvl3 antibody (red) is negligible for Dvl3 expressed either alone or in combination with Fzd5 but strong for evenly distributed Dvl3 after CK1ε co-expression. Total Dvl3 detected by anti-Flag antibody is shown in green. All confocal images were acquired using the same laser/detector settings and subsequently quantified using ImageJ software. Graphs show the overlap of fluorescence intensity peaks of individual channels along profiles indicated in the merged micrographs by a white line. Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 15 Table 1 Position 1 Dvl/Dvl+CK1ε2 Presence 3 Conservation 4 CK1 5 CK2 6 PKC 7 PLK1 8 Peptide 9 Note 10 15 0/1 ↑ Y o HLDGQETPYLVKLPL 48 1/0 ↓ Y X RPSYKFFFKSMDDDFGVV K 61 0/1 ↑ Dvl2 X FGVVKEEISDDNAKLPCF 116 1/1 ↔ Y X GIGDSRPPSFHPHAGGGS 125 6/6 ↔ Y X HPHAGGGSQENLDND 133 3/3 ↔ y X X QENLDNDTETDSLVS 135 2/3 ↖↗ y X X NLDNDTETDSLVSAQ 137 3/3 ↔ Y X DNDTETDSLVSAQRE 140 4/4 ↔ Y X X TETDSLVSAQRERPR 175 2/1 ↓ Y ~ RRREPGGYDSSSTLM 192 7/5 ↓ Y X ETTSFFDSDEDDSTS 197 3/1 ↓ Dvl2 X X X FDSDEDDSTSRFSSS 202 4/4 ↓ Y X X DDSTSRFSSSTEQSS C1 203 7/4 ↓ y X X DDSTSRFSSSTEQSS C1 204 2/1 ↔ Y o o o DDSTSRFSSSTEQSS C1 205 3/6 ↔ Y ~ ~ TSRFSSSTEQSSASRL C1 208 5/7 ↔ Y X FSSSTEQSSASRLMR 209 5/7 ↔ Y X FSSSTEQSSASRLMR 211 0/1 ↑ Y o o SSTEQSSASRLMRRHK 232 5/5 ↔ Dvl2 X X KVSRIERSSSFSSIT 234 6/5 ↔ Y X KVSRIERSSSFSSIT 237 6/4 ↔ Y o ERSSSFSSITDSTMS 280 0/4 ↑ Y GGIYIGSIMKGGA 311 0/4 ↑ Y X EINFENMSNDDAVRV 350 0/4 ↑ n CFTLPRSEPIRPID 421 3/2 ↔ n X AIVKAMASPESGLEV 512 2/3 ↔ Y X SLHDHDGSSGASDQD 516 4/5 ↔ Y o DGSSGASDQDTLA 564 1/1 ↔ Y o LGYSYGGGSASSQHSEGS 566 1/3 ↔ n X SYGGGSASSQHSEGS 567 3/5 ↖↗ Y o SYGGGSASSQHSEGS 570 6/4 ↔ Y X GSASSQHSEGSRSSG 573 5/6 ↔ Y X SSQHSEGSRSSGSNR 575 4/5 ↔ Y X HSEGSRSSGSNRSGS 576 4/5 ↔ Y X X HSEGSRSSGSNRSGS 578 3/4 ↔ Y X EGSRSSGSNRSGSDR 601 3/4 ↔ Y X GDSKSGGSGSESDHT 603 4/3 ↔ Y X X GDSKSGGSGSESDHT 605 1/1 ↔ Y o SGGSGSESDHTTRSSLR 611 1/3 ↖↗ Dvl1 X ESDHTTRSSLRGPRE C4 612 1/3 ↖↗ Dvl1 X ESDHTTRSSLRGPRE C4 622 2/6 ↔ n X GPRERAPSERSGPAA Phosphorylation of Dvl3 by CK1ε: MS/MS analysis 16 625 2/6 ↔ n o RERAPSERSGPAASEHS 633 0/1 ↑ n o RSGPAASEHSHRSHHSLAS C2 636 0/5 ↑ n ASEHSHRSHHSLASSLR C2 639 0/6 ↑ n ~ EHSHRSHHSLASSLRSH C3 642 1/3 ↖↗ n o SHHSLASSLRSHHTH C3 643 1/5 ↖↗ n X SHHSLASSLRSHHTH C3 689 2/5 ↖↗ Dvl2 PPGRDLASVPPELTA 697 1/1 ↔ Y o SVPPELTASRQSFRMAMG 700 1/1 ↔ Y X ELTASRQSFRMAMGN 1 Position of aa; 2 Number of identification in Dvl/Dvl3 +CK1ε; (found in n= experiments, n=7)3 Presence of phosphorylation; ↑ only in induced; ↖↗ more in induced; ↔ constitutively phosphorylated; ↓ less in induced; 4 Y- conserved in all isoforms (species considered: M. musculus, H. sapiens, X. laevis, X. tropicalis) as S/T/Y; y- conserved in all Dvl isoforms either as S/T/Y or polar aa (D,E,Q,N); Dvl1/Dvl2 conserved also in this isoform; n- not conserved; 5;6;7;8; X high strindency prediction; o medium stringeny prediction; ~ low stringency prediction;9 phosphorylated peptide sequence; Phosphorylation in Bold10 C1/C2/C3/C4 – mutated in cluster Předmět: JBC/2014/547521 ‐‐ Manuscript Decision Od: jkyriakis@asbmb.org Datum: 24.1.2014 16:56 Komu: "Vítězslav Bryja"  MS ID#: JBC/2014/547521 MS TITLE: Functional analysis of Dishevelled‐3 phosphorylation identifies distinct mechanisms driven by Casein Kinase 1&[epsilon] and Frizzled5 Dear Dr. Bryja: Your manuscript entitled "Functional analysis of Dishevelled‐3 phosphorylation identifies distinct mechanisms driven by Casein Kinase 1&[epsilon] and Frizzled5" (JBC/2014/547521) has been reviewed by the Editorial Board.   Unfortunately, the decision was made to decline the manuscript in its current form.  The reviewers did, however, recommend consideration of a suitably and substantially revised version that addresses the criticisms raised in the comments shown below. Note that no guarantee of acceptance of the revised manuscript is implied. Some new experimentation is needed in any revision, and a revision that does not include new experiments will not be considered.  Chiefly, only a gain‐of‐function approach was taken to identify CK1‐related phosphorylation sites. These events have to be evaluated in the absence of CK1 activity by using a loss‐of‐function approach, such as RNAi.  In addition, the in vivo physiological/functional significance of the phosphorylation (or de‐phosphorylation) of some of the sites, such as S280/S311, needs to be demonstrated.  Lastly, several of the figures need to be reconfigured for greater clarity. Upon receipt of a revised manuscript, it will be sent for further editorial evaluation.  The revised manuscript should be submitted via the JBC Electronic Submission site at http://submit.jbc.org as a single PDF file. Any supplemental files to be published with the manuscript must conform to our Editorial Policies. Supplemental information must be limited to such things as videos, 3‐D structures/images, extended chemical syntheses, extensive NMR data, molecular dynamics, kinetic modeling data, and other large data sets — e.g., those obtained with microarray analyses or mass spectrometry studies. (For more information, please see http://www.jbc.org/site/misc/ifora.xhtml#supplemental) The manuscript should be accompanied by a cover letter in which you provide a point‐by‐point listing of the changes made. The cover letter should not be in the PDF file containing the manuscript—it must be entered at the online submission site.  It is important for major findings of your study to be intelligible to all of our readers, including those who are not specialists in the field.  Please review your title and summary to make sure they convey your essential points succinctly and clearly and do not contain undefined abbreviations or specialized terms. If you resubmit, please carefully follow the Instructions to Authors at the JBC Online Submission Site for preparation of text and graphic files at http://www.jbc.org/site/misc/itoa.xhtml to resubmit your manuscript on line. Please note that the revised manuscript must be submitted within 120 days of the decision date or a new manuscript number will be assigned.  Please include the manuscript number of your original submission in your cover letter. Thank you for the opportunity to consider your work. Sincerely, John M. Kyriakis Associate Editor Reviewer 1 COMMENTS FOR THE AUTHOR: The authors addressed which Dvl residues were phosphorylated by CK epsilon. Using LC‐MS/MS, 16 sites  were identified to be candidate phosphorylation sites of Dvl induced by CK epsilon. They then carried out  mutation experiments. Mutations in PDZ domain prevented activation induced by Wnt/beta‐catenin.  Mutations of clustered S/T in the C‐terminus prevented CK1‐induced mobility shift, whereas the  downstream signal was not affected. However, phosphorylation triggered by Fzd5 seems to be qualitatively  JBC/2014/547521 ‐‐ Manuscript Decision mailbox:///C:/Users/Vita Bryja/AppData/Roaming/Thunderbird... 1 z 2 4.3.2014 8:47 different. To explore further the functional implication, they examined subcellular localization. Although  Dvl signal was found to be punctate, co‐expression of CK epsilon lead to even pattern, and Fzd expression  resulted in membrane pattern. Mutations of clustered S/T in the C‐terminus Dvl prevented this subcellular  localization. From these data, the authors suggest that CK epsilon acts on Dvl that is independent of Fzd5.  The data are clear and supportive of the conclusions. However, some of the results should be described in  more detail and several data need to be shown clearly. Specific comments are as follows:  1. Figure 1A and B: Molecular weight for A should be labeled, and y‐axis of graph B should also be labeled. 2. Figure 3A, B, and C: y‐axis of the graphs should be labeled. Appropriate description about the graphs is  also lacking. 3. Figure 4E: y‐axis of the graphs should be labeled. Appropriate description about the graphs is lacking. 4. Figure 4B: I could not find retarded migration of the band of Dvl‐C1‐SA. 5. Figure 4F: Expression level of FLAG‐hDvl3 seems to be low when co‐transfected withV5‐Fzd5. Why? Reviewer 2 COMMENTS FOR THE AUTHOR: Bernatik et al. has analyzed phosphorylation events of human Dvl3 induced by CK1 kinase. Since there are ~50 phosphorylation sites identified, the authors focused on some, like S280 and S311, and carried out further in vitro functional analysis. Molecular details learned from this study are certainly very useful for a better understanding of Wnt signaling and its regulation. However, there are some major concerns about this work. 1) Only a gain‐of‐function approach was taken to identify CK1‐related phosphorylation sites. These events have to be evaluated in the absence of CK1 activity by using a loss‐of‐function approach, such as RNAi. 2) In vivo physiological/functional significance of the phosphorylation (or de‐phosphorylation) of some of the sites, such as S280/S311, need to be demonstrated. Sent on: January 24, 2014 JBC/2014/547521 ‐‐ Manuscript Decision mailbox:///C:/Users/Vita Bryja/AppData/Roaming/Thunderbird... 2 z 2 4.3.2014 8:47