β-Catenin has a key role in Wnt signaling via effects on T-cell factor (TCF)-mediated transcription. Mutational defects in β-catenin regulation are seen in many cancers, leading to elevated β-catenin levels, enhanced binding of β-catenin to TCFs, and increased expression of TCF-regulated genes. Factors cooperating with β-catenin in transcription of TCF-regulated genes are not well defined. TIP49, an ATPase previously implicated as a cofactor for oncogenic transformation by c-Myc, has been shown to bind to β-catenin. We found that expression of an ATPase-deficient mutant form of TIP49 (TIP49D302N) substantially inhibited β-catenin-mediated neoplastic transformation of immortalized rat epithelial cells and anchorage-independent growth of human colon cancer cells with deregulated β-catenin. The TIP49D302N mutant inhibited β-catenin-mediated activation of TCF-dependent cellular genes. Similar inhibition of the expression of β-catenin/TCF-dependent genes was seen with small interfering RNA approaches against endogenous TIP49. TIP49 was found in complexes with chromatin remodeling and histone-modifying factors and cofactors, including the TIP60 histone acetylase-associated proteins transactivation/transformation-domain associated protein (TRRAP) and BAF53. Using chromatin immunoprecipitation methods, the TIP49, TIP60, and TRRAP proteins were found to interact with sequences in the regulatory region of the gene for ITF-2, a TCF-dependent cellular gene. The ability of TIP49D302N to inhibit ITF-2 gene expression was linked to decreased acetylation of histones in the vicinity of the TCF-binding sites in the ITF-2 promoter region. We suggest that TIP49 is an important cofactor in β-catenin/TCF gene regulation in normal and neoplastic cells, likely functioning in chromatin remodeling.

The Wnt signaling pathway regulates many important processes, such as polarity of cell division, cell proliferation, and cell fate determination (1, 2). Deregulation of Wnt signaling has been implicated in the pathogenesis of many different human cancers (reviewed in Refs. 3 and 4). β-Catenin has a central role in the canonical Wnt signaling pathway, and tight control of β-catenin protein levels and its subcellular localization play a crucial role in regulating the signaling activity of β-catenin. In the absence of Wnt stimulation, glycogen synthase kinase 3β (GSK3β) appears to phosphorylate several serine and/or threonine residues in β-catenin’s NH2-terminal region, leading to recognition of phosphorylated β-catenin by the β-TrCP1 ubiquitin ligase, polyubiquitination, and subsequent degradation of cytosolic β-catenin by the 26S proteasome (5, 6, 7, 8, 9, 10). Complex formation of GSK3β with β-catenin, adenomatous polyposis coli, and Axin (or its homolog, known as Conductin, Axil, or Axin2) facilitates β-catenin phosphorylation (11). Binding of a Wnt protein to its cognate Frizzled receptor complex at the cell surface leads to inhibition of the kinase activity of GSK3β, thus allowing unphosphorylated β-catenin protein to escape degradation and to accumulate and enter the nucleus (12). There, β-catenin can bind to members of the T-cell factor (TCF) family to promote target gene transcription (13, 14). In cancers of various types, β-catenin degradation is often inhibited as a result of inactivating mutations in adenomatous polyposis coli or Axin1 or gain of function mutations affecting the key phosphorylation sites in the NH2 terminus of β-catenin (4). As a consequence, β-catenin is stabilized, and complex formation with TCF is enhanced, leading to activation of a variety of downstream genes, including c-Myc(15), Cyclin D1(16, 17), matrilysin/MMP7(18), and ITF-2(19).6

Several proteins have been implicated in modulating the activity of the β-catenin/TCF complex and/or expression of TCF-regulated genes. For instance, the Groucho/transducin-like enhancer of split family of proteins have been shown to bind to TCFs in the absence of β-catenin, and the Groucho/transducin-like enhancer of split proteins repress TCF-dependent gene expression (20). Groucho interacts with histones (21) and the histone deacetylase Rpd3 (22), presumably forming a chromatin structure incompatible with transcription. Other mechanisms of inhibition of β-catenin/TCF transcription may include acetylation of TCF by cAMP-responsive element binding protein (CREB)-binding protein (CBP) in Drosophila(23) and binding of β-catenin by the small polypeptide inhibitor ICAT (inhibitor of β-cat and TCF) (24) or the proteins Duplin (25) or Chibby (26).

At present, the mechanisms by which the β-catenin/TCF complex activates expression of key target genes are not well understood. Previous attempts to search for critical β-catenin cofactors identified TATA box-binding protein (TBP; Ref. 27) and p300/CBP (28, 29, 30). Both proteins may function to link β-catenin to the basal transcription machinery. In addition, p300/CBP may help to loosen local chromatin structure via its histone acetyltransferase (HAT) activity, allowing access of other cofactors to target gene promoters. However, the interactions with TBP and p300/CBP do not appear to fully explain how β-catenin stimulates transcription. For example, some transactivating elements at the β-catenin COOH-terminal region can stimulate gene expression without interacting with TBP or p300 (27, 31). In addition, although TBP and p300/CBP are ubiquitously present, nuclear localization of β-catenin and its interaction with TCFs are not always sufficient for robust activation of TCF transcriptional activity, as is the case in normal T lymphocytes (32). Finally, the activity of p300 is evident only for a subset of β-catenin target gene promoters (28). β-Catenin, therefore, may use different cofactors to fulfill its function on particular target genes, and these cofactors may include components of the chromatin remodeling complexes or the basal transcription machinery. As an example, recent studies by Barker et al.(33) showed that the chromatin remodeling factor Brg-1 can interact with β-catenin to promote activation of certain TCF-target genes. Additionally, genetic approaches in Drosophila have identified the protein pygopus as a cofactor of β-catenin/TCF-mediated transcription (34, 35, 36). The presence of a PHD domain in pygopus implicates pygopus in chromatin remodeling (37, 38), but the exact mechanism underlying its function is not clear.

We have chosen to explore further the interaction of β-catenin with TIP49 (TBP-interacting protein 49; Ref. 39). The TIP49 protein [also known as TIP49a (40), Pontin52 (41), NMP238 (42), RUVBL1 (43), Rvbl (44), TAP54α (45), and TIH1p (46)] and TIP49’s close relative, TIP48 (39) [also known as TIP49b (47), Reptin52 (48), RVBL2 (43), Rvb2 (44), TAP54b (45), and TIH2p (49)], are members of a highly conserved protein family with sequence similarity to bacterial RuvB, an ATP-dependent DNA helicase that catalyzes branch migration in the Holiday junction. The TIP49 and TIP48 proteins have been suggested to possess single-strand DNA-stimulated ATPase activities and DNA helicase activities of opposite polarity (47, 50), consistent with a possible role for the proteins in chromatin metabolism and transcriptional regulation. Indeed, TIP49 and TIP48 have been found in the yeast INO80 chromatin remodeling complex, which is involved in transcription and DNA repair (44), as well as in the human TIP60 (human immunodeficiency virus-1 Tat interacting protein 60) HAT complex, which plays a role in DNA repair and apoptosis, possibly through histone modifications (45). Additionally, Jonsson et al.(51) showed that Rvb1 and Rvb2, the yeast homologs of human TIP49 and TIP48, are associated with a complex that shows ATP-dependent chromatin remodeling activity in vitro and that mutation of either protein affects transcription of over 5% of yeast genes. A more direct role of TIP49 and TIP48 in transcription is suggested by recent data on their ability to modestly affect β-catenin/TCF transcription of a model reporter gene (48), but the significance and role of TIP49 and TIP48 in β-catenin-mediated regulation of endogenous cellular genes have not been explored. Of some potential interest, in prior studies, TIP49 was found to interact with c-Myc and to affect c-Myc-mediated oncogenic transformation (39). However, no evidence was offered that TIP49 affects c-Myc-mediated transcription of reporter genes (39).

Here, we report data implicating TIP49 as an essential cofactor in β-catenin-mediated neoplastic transformation. Consistent with this observation, stable expression of an ATPase-deficient TIP49 mutant blocked expression of endogenous β-catenin/TCF target genes, and small interfering RNAs (siRNAs) directed against endogenous TIP49 had similar effects on β-catenin/TCF target gene activation. Inhibition of β-catenin/TCF target gene expression was linked to inhibition of histone acetylation of β-catenin target gene sequences, suggesting that TIP49 mediates effects on β-catenin/TCF target transcription through chromatin remodeling mechanisms.

Plasmids.

The wild-type allele of TIP49 was cloned by PCR using random hexamer-primed cDNA from 293 cells as template. The wild-type TIP49 cDNA was then used as a template in a PCR-based approach to generate TIP49D302N. Both wild-type TIP49 and TIP49D302N PCR-generated cDNAs contain a FLAG epitope tag at the COOH terminus. Both PCR products were subcloned into either the pcDNA3 vector (Invitrogen, San Diego, CA) or the retroviral vector pPGS-CMV-CITE-neo (52). All plasmid sequences were verified by automated DNA sequencing. The pBabe-S33Yβ-catenin-puro expression vector encoding the S33Yβ-catenin mutant protein was generated by cloning the S33Yβ-catenin cDNA into the retroviral vector pBabe-puro (53). The plasmid carrying a mutant K-ras cDNA (a valine to cysteine mutation at codon 12) in the retroviral vector pBMN and the constructs with the FLAG-tagged S33Yβ-catenin cDNA cloned into the pcDNA3 vector or the pBMN vector have been described previously (52). The reporter constructs pTOPFLASH and pFOPFLASH (kindly provided by B. Vogelstein) contain either three copies of an optimal TCF-binding motif or three copies of a mutant motif, respectively (54). Plasmid pCH110 (Pharmacia, Piscataway, NJ) contains a functional LacZ gene downstream of a cytomegalovirus early-region promoter-enhancer element and was used as control for transfection efficiency in reporter assays.

Cell Lines and Tissue Culture.

Phoenix, RK3E, 293, LoVo, SW480, and HCT116 cells were grown in DMEM supplemented with 10% fetal bovine serum. β-Catenin-transformed cell lines RK3E/S33Y-B, RK3E/S33Y-C, and RK3E/ΔN47 have been described previously by Kolligs et al.(52). The polyclonal RK3E-derived cell lines RK3E/Neo, RK3E/TIP49wt, and RK3E/TIP49D302N were obtained after infection with the respective pPGS-CMV-CITE-neo vector-based retrovirus and subsequent selection of the bulk cell population in G418 at a concentration of 1 mg/ml. After 1 week, the G418 concentration was reduced to 250 μg/ml, and the expression of the FLAG-tagged TIP49 protein was confirmed by Western blot analysis. Focus formation assays in these stable RK3E/Neo, RK3E/TIP49wt, and RK3E/TIP49D302N lines were carried out as described previously (52). Polyclonal G418-resistant β-catenin-transformed RK3E, SW480, HCT116, or LoVo cell lines stably expressing wild-type TIP49, the TIP49D302N mutant, or only the neomycin resistance gene were generated in the same manner as that for RK3E-derived cells. Assays of colony formation in soft agar were performed as described previously (52). To assess the effect of TIP49D302N on histone acetylation of β-catenin target gene promoters, RK3E/Neo, RK3E/TIP49wt,and RK3E/TIP49D302N cells were infected with retroviral supernatant from amphotrophic Phoenix cells transfected with pBabe-S33Yβ-catenin-puro. After 1 week of drug selection of the transduced cells in 1.0 μg/ml puromycin (Sigma, St. Louis, MO), cells were subjected to analysis with chromatin immunoprecipitation (ChIP) assays.

Luciferase Reporter Gene Assays.

Transient transfection and reporter gene assays were carried out essentially as described previously (52). In brief, RK3E cells were transfected with 0.2 μg of pcDNA3/S33Yβ-catenin, 1.2 μg of the pcDNA3 expression vectors encoding either wild-type TIP49 or TIP49D302N, 0.4 μg of pTOPFLASH or pFOPFLASH, and 0.2 μg of pCH110. The total mass of DNA for each transfection was kept constant by adding empty pcDNA3 vector. Two days after transfection, luciferase activities were measured in a luminometer and normalized to β-galactosidase expression.

Northern Blot Analysis.

Total RNA was extracted from cells using Trizol reagent (Invitrogen). Ten μg of total RNA were separated on 1.2% formaldehyde-agarose gels and transferred to Zeta-Probe GT-membranes (Bio-Rad, Hercules, CA) by capillary action. Human ITF-2 and Axin 2 and rat ITF-2, Axil, and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) gene fragments were amplified by PCR and labeled with [32P]dCTP by random priming. Northern blot hybridization to 32P-labeled probes was carried out by standard methods. Signals were detected by exposure to BioMax-MS film (Kodak, Rochester, NY) at −80°C with an intensifying screen. Just after the autoradiograph was obtained for each gene, the membrane was placed in a phosphorimaging cassette (Molecular Dynamics, Sunnyvale, CA). The cassettes were then exposed in a PhosphorImager (Molecular Dynamics), and subsequent quantitative analysis was performed using Imagequant software (Molecular Dynamics).

Coimmunoprecipitation and Western Blot Assays.

Western blot analysis was performed as described previously (52). For coimmunoprecipitation, LoVo cells stably expressing wild-type TIP49, the TIP49D302N mutant, or only the neomycin resistance gene were lysed in lysis buffer [10 mm Tris (pH 7.05), 50 mm NaCl, and 1% Triton X-100] supplemented with proteinase inhibitors (complete proteinase inhibitors; Roche Applied Science, Indianapolis, IN). To determine the level of protein expression, whole-cell lysates were analyzed by Western blotting with anti-β-catenin monoclonal antibody (Transduction Laboratories, Lexington, KY) or anti-FLAG M2 antibody (Sigma) for detection of TIP49. Immunoprecipitations (IPs) were carried out in coimmunoprecipitation buffer [50 mm Tris (pH 7.5), 100 mm NaCl, 0.1% Triton X-100, 15 mm EGTA, and proteinase inhibitors] with anti-FLAG polyclonal antibody (Sigma) followed by incubation with protein A-agarose (Pierce, Rocklord, IL). Precipitates were washed four times with coimmunoprecipitation buffer and subjected to Western blot analysis with anti-β-catenin monoclonal antibody and anti-FLAG M2 antibody. For immunoprecipitation with HCT116 and SW480 lines stably expressing FLAG-tagged TIP49, anti-FLAG M2 affinity gel (Sigma) was used, and the assay was performed according to the manufacturer’s suggestions. Polyclonal antibody against BAF53 was kindly provided by Gerald Crabtree. For assessment of interaction of β-catenin with transactivation/transformation domain-associated protein (TRRAP), 293 cells were lysed in lysis buffer [20 mm Tris (pH 7.5), 140 mm NaCl, 1% Triton X-100, and 10% glycerol] supplemented with complete proteinase inhibitors and immunoprecipitated with rabbit polyclonal antibody against TRRAP (Santa Cruz Biotechnology, Santa Cruz, CA) or polyclonal anti-polyhistidine antibody (Santa Cruz Biotechnology) as control. To assess protein expression in RK3E cells, whole-cell extracts were prepared with radioimmunoprecipitation assay lysis buffer (Tris-buffered saline, 0.5% deoxcholate, 0.1% SDS, and 1% NP40) supplemented with complete proteinase inhibitors. Lysates were then analyzed by Western blotting as described previously (52) with anti-β-catenin monoclonal antibody and anti-FLAG M2 antibody. Expression of β-actin was used to control for loading of the lanes and detected with anti-actin polyclonal antibody (Sigma).

siRNA Transfection.

The TIP49 siRNA and the control scrambled siRNA duplexes with two thymidine overhangs were purchased from Dharmacon Research (Lafayette, CO). The target sequence for the TIP49 siRNA is 5′-AATGGCTGGAAGAGCTGTCTT-3′. For transfection, 293 cells were cotransfected with 1.68 μg of siRNA duplxes, together with 4 μg of pcDNA3/S33Yβ-catenin or 4 μg of pcDNA3/S33Yβ-catenin plus 4 μg of pcDNA3/TIP49D302N (in 60-mm culture dish) by Lipofectamine 2000 (Invitrogen). The total mass of plasmid DNA for each transfection was kept constant by adding empty pcDNA3 vector. Two days after transfection, cells were lysed and subjected to Western blot using an anti-TIP49 rabbit polyclonal antibody generated against a TIP49 peptide corresponding to amino acids 437–456 (Alpha Diagnostic, San Antonio, TX) for endogenous TIP49 expression and an anti-FLAG M2 antibody for detection of S33Yβ-catenin expression. For detection of the transfected FLAG-tagged TIP49D302N, the anti-FLAG M2 antibody was used. Expression of β-actin was used for loading control. In parallel, total RNA from the transfected cells was prepared and subjected to Northern blot analysis using human ITF-2, AXIN2, TIP49, and GAPDH cDNA probes.

ChIP Assay.

The ChIP assay was performed according to the manufacturer’s recommendation (Upstate Biotechnology, Lake Placid, NY). In brief, 1 × 106 to 4 × 106 cells were treated with 1% formaldehyde to cross-link proteins to the DNA. The chromatin was then prepared, sonicated, and immunoprecipitated with rabbit polyclonal antibodies against acetylated histone H4 (Upstate Biotechnology), TIP49, TRRAP (Santa Cruz Biotechnology), and TIP60 (Upstate Biotechnology). The recovered DNA was analyzed by PCR using a pair of primers that flank the TCF-binding site at −565 in the rat ITF-2 promoter or the primers flanking the TCF-binding site at −1874 in the human ITF-2 promoter. As negative controls, distal sequences were amplified with primers corresponding to the region 1 kb upstream of the TCF-binding site or the downstream coding region in the rat ITF-2 gene and primers located at the downstream coding region in human ITF-2. The sequences for these primers are as follows: (a) the TCF site in rat ITF-2, 5′-GTAAGGAGGTGGTCAATACTCTAG-3′ (forward primer) and 5′-GTAATCCTGAGCTTGAAGAGTGG-3′ (reverse primer); (b) upstream for rat ITF-2, 5′-CGACTCGATCATAGGCTACCC-3′ (forward primer) and 5′-CTGAGAAAGCACAGTCAGGTTC-3′ (reverse primer); (c) downstream for rat ITF-2, 5′-GAGCCGAATTGAAGACCGTTTAG-3′ (forward primer) and 5′-ATGAGTGAATGTCTGTTGGCTGAG-3′ (reverse primer); (d) the TCF site in human ITF-2, 5′-GGAGAGGACCCACATCCCTC-3′ (forward primer) and 5′-GAGCAGGCGACCATAGAGTGG-3′ (reverse primer); and (e) downstream for human ITF-2, 5′-AAGCCGAATTGAAGATCGTTTAG-3′ (forward primer) and 5′-ATGAGTGAATGTCTGTTGGCTGAG-3′ (reverse primer). The input sample contains 0.1% of the total input chromatin as PCR template.

Role of TIP49 in Transient Assays of TCF-Dependent Reporter Gene Activity.

Perhaps in part because it may be a component of several different nuclear protein complexes, TIP49 has been implicated in diverse pathways including neoplastic transformation (39), DNA repair, and apoptosis (44, 45). Functions in chromatin metabolism and transcription have been suggested for TIP49, but its specific role in these processes is rather obscure. TIP49 is homologous to the bacterial RuvB protein, and it possesses intrinsic ATPase activity and contains the Walker A and Walker B motifs (Fig. 1,A), which are found in proteins that bind and hydrolyze ATP (55, 56). A single missense mutation in the Walker B box (DEVH→NEVH) of bacterial RuvB is sufficient to its abolish ATPase activities, creating a dominant negative allele (57). A previous study by Wood et al.(39) showed that an analogous mutation in the Walker B box of TIP49 (D302→N; Fig. 1,A) was able to interfere with c-Myc-mediated transformation. However, it was also reported in the Wood et al.(39) study that the TIP49D302N mutant did not show any effect on the ability of c-Myc to activate transcription of c-Myc-dependent reporter genes. In light of the fact that TIP49 interacts with β-catenin and has been suggested to positively modulate β-catenin/TCF transcription (41, 48), we first tested whether the TIP49D302N mutant could affect activation of the model TCF reporter construct pTOPFLASH, which contains three optimal TCF-binding sites upstream of a minimal promoter element and the luciferase gene (54). Ectopic expression of the S33Y oncogenic form of β-catenin (S33Yβ-catenin; Ref. 52) strongly activated TCF-dependent transcription in the E1A-immortalized RK3E rat epithelial cell line (Fig. 1,B). However, whereas expression of wild-type TIP49 did not enhance the ability of β-catenin to activate TCF reporter gene activity, expression of the TIP49D302N mutant had a modest ability to inhibit the effect of β-catenin on TCF-dependent reporter gene activity (Fig. 1 B).

TIP49D302N Inhibits Neoplastic Transformation by Mutant β-Catenin.

Our previous work has shown that cancer-derived mutant forms of β-catenin can readily induce morphologically transformed foci when introduced into the RK3E cell line, an adenovirus E1A-immortalized rat epithelial line (52). Although c-Myc has been suggested to be a key downstream target of Wnt/β-catenin/TCF action in cancer cells (15), in the RK3E system, it appears that stabilized oncogenic forms of β-catenin mediate transformation through c-Myc-independent mechanisms (52). Based on the interaction of TIP49 with β-catenin, we sought to determine whether TIP49 plays a role in β-catenin-induced transformation of RK3E. Polyclonal G418-resistant RK3E lines with stable expression of either wild-type TIP49 (TIP49wt) or TIP49D302N were generated, and focus formation assays with these cells were carried out using replication-defective retroviruses encoding the S33Y oncogenic mutant form of β-catenin. Consistent with prior studies (52), S33Yβ-catenin induced 500–600 foci in the control RK3E/Neo cells in 4 weeks (Fig. 2). Although S33Yβ-catenin generated roughly equivalent numbers of foci in the RK3E/TIP49wt and control RK3E/Neo lines, the number of foci generated in the RK3E/TIP49D302N line was 70–80% less than that in the control line (Fig. 2, A and B). Expression of the TIP49wt and TIP49D302N proteins in the RK3E lines is shown in Fig. 2 C.

To assess the specificity of the effect of TIP49D302N on β-catenin-induced neoplastic transformation versus cell growth and neoplastic transformation in general, the growth and transformation susceptibility of the various RK3E lines were studied further. No clear-cut effects on in vitro proliferation rates were observed when the RK3E lines expressing wild-type or D302N forms of TIP49 were compared with one another or with RK3E/Neo control cells (data not shown). In addition, the ability of mutant K-ras to induce anchorage-independent growth in the control RK3E/Neo and RK3E/TIP49D302N lines was assessed. As shown in Fig. 3, mutant K-ras showed similar transforming activity in the RK3E/TIP49D302N and RK3E/Neo lines, indicating that the requirement for TIP49 in β-catenin-mediated transformation of RK3E cells does not simply reflect a role for TIP49 in oncogenic transformation in general.

To assess the role of TIP49 in human cancer cells with defective β-catenin regulation, we generated stable cell lines of LoVo (adenomatous polyposis coli mutant) human colon cancer cells expressing TIP49wt or TIP49D302N, as well as control G418-resistant cells. A marked reduction in colony formation in soft agar was seen in LoVo cells expressing TIP49D302N relative to the other two lines (Fig. 4). Taken together, the results in RK3E and LoVo cells imply that TIP49 plays a role in β-catenin-mediated neoplastic transformation and that its ATPase activity is likely required for its function.

Expression of TIP49D302N Interferes with Expression of Endogenous β-Catenin/TCF Target Genes.

The oncogenic activity of β-catenin appears to be closely related to its transcriptional activity (52). Our finding that TIP49D302N strongly inhibited β-catenin-mediated neoplastic transformation contrasted sharply with the modest inhibitory effects of TIP49D302N in the transient reporter gene assay with the pTOPFLASH model reporter gene construct. Considering the possible function of TIP49 in chromatin remodeling, we hypothesized that the effect of TIP49 on β-catenin/TCF transcription might not be readily apparent with DNA templates not organized into higher order chromatin complexes. Rather, the function of TIP49 in β-catenin/TCF transcription might be more obvious in studies of endogenous β-catenin/TCF-regulated genes. To test this idea, we performed Northern blot analyses to assess expression of two endogenous β-catenin/TCF target genes, namely, ITF-2(19) and Axil(58, 59, 60, 61). As shown in Fig. 5,A, Northern blot analysis of the time course of ITF-2 and Axil gene expression after infection of the control RK3E/Neo cells with the retrovirus encoding the S33Yβ-catenin mutant protein or the control retroviral vector revealed that both ITF-2 and Axil were induced by β-catenin at 2 days, with much stronger induction at day 6, largely paralleling the time course of exogenous S33Yβ-catenin protein accumulation in the cells after retrovirus infection (Fig. 5,B). Although no obvious difference in induction of ITF-2 and Axil by β-catenin was observed at day 2 for RK3E cells stably expressing TIP49D302N versus the control RK3E/neo cells, the ability of the stabilized S33Yβ-catenin protein to mediate further elevation of ITF-2 and Axil expression at day 6 was substantially blunted in RK3E/TIP49D302N cells (Fig. 5, A and C). The minimal induction of ITF-2 and Axil by β-catenin seen in RK3E/TIP49D302N cells may be due to the incomplete disruption of normal TIP49 function by the TIP49D302N mutant.

To further address the role of TIP49 in β-catenin/TCF transcription, we also pursued Northern blot studies of ITF-2 and Axil expression in β-catenin-transformed lines stably expressing the TIP49D302N mutant. Consistent with our previous results (19, 58), ITF-2 and Axil were highly expressed in the three β-catenin-transformed RK3E lines compared with the parental RK3E cells (Fig. 5, D and F). Stable ectopic expression of the TIP49D302N protein in all three β-catenin-transformed RK3E lines reduced ITF-2 and Axil expression, with ITF-2 expression decreased by 30–40% and Axil expression decreased by 40–50%, when compared with the levels of expression seen in the control RK3E line (Neo) or the line stably expressing wild-type TIP49 (Fig. 5, D and F). Expression of TIP49D302N did not affect β-catenin expression (Fig. 5 E).

To strengthen the case that endogenous TIP49 plays a role in regulating the transcription of chromosomal β-catenin/TCF target genes, we used siRNA approaches to reduce endogenous TIP49 levels in human 293 cells and then used Northern blot approaches to study effects on expression of the ITF-2 and AXIN2 genes. As shown in Fig. 6,A, endogenous TIP49 levels were moderately reduced in 293 cells transfected with the TIP49-specific siRNAs (Fig. 6,A, Lanes 2 and 4), but not the control scrambled siRNAs (Lanes 1 and 3). As expected, in the cotransfection experiments, the siRNAs had no effect on ectopic S33Yβ-catenin expression (Fig. 6,A, Lanes 3 and 4). In cells transduced with the TIP49-specific siRNAs, modest inhibition of basal expression of ITF-2 and AXIN2 was seen, and more significant inhibition of S33Yβ-catenin-induced ITF-2 and AXIN2 expression was observed (Fig. 6, B and C). In fact, the inhibitory effects on β-catenin/TCF target gene expression seen with the TIP49 siRNAs were a bit stronger than those seen with the TIP49D302N mutant (Fig. 6, B and C). Based on these results and the finding that TIP49 is required for β-catenin-mediated transformation, we propose that TIP49 is a critical cofactor for neoplastic transformation by β-catenin via its role in regulating β-catenin/TCF transcription of endogenous chromosomal target genes.

Interaction of TIP49 With Chromatin-Associated Proteins and TCF Sites.

As noted above, TIP49 has previously been found to be complexed in the nucleus with β-catenin and several other proteins, including c-Myc, TRRAP, a c-Myc-interacting protein that is part of several distinct HAT complexes, and BAF53, another c-Myc interacting cofactor that may have independent associations with chromatin remodeling factors (39, 62). Other work has indicated that TIP49 may be present in chromatin modifying and remodeling complexes, such as the yeast INO80 remodeling complex (44) and the human TIP60 HAT complex (45). To address the means by which TIP49 exerts its effects on β-catenin/TCF target gene expression, we sought to show that both wild-type TIP49 and the TIP49D302N proteins did, in fact, interact with β-catenin in cells. As shown in Fig. 7,A, IP of the FLAG-tagged TIP49 and TIP49D302N proteins from LoVo cells stably expressing the FLAG-tagged proteins yielded similar amounts of immunoprecipitated β-catenin protein. To determine whether BAF53 was complexed with TIP49 and β-catenin in cells, we generated HCT116 and SW480 colon cancer cell lines expressing a FLAG-epitope-tagged form of wild-type TIP49 as well as control drug-resistant lines. After IP of the lysates with anti-FLAG antibodies, we analyzed the immunoprecipitates for the presence of β-catenin and BAF53 by immunoblotting. As shown in Fig. 7,B, endogenous β-catenin and BAF53 proteins were specifically present in the anti-FLAG immunoprecipitates from HCT116 and SW480 cells expressing the FLAG-tagged TIP49 proteins. Efforts to establish using combined IP and immunoblot approaches that endogenous TIP49 associated with endogenous β-catenin and BAF53 proteins were hindered by the fact that the TIP49 and BAF53 proteins essentially comigrate with the strongly reactive immunoglobulin heavy chain band in combined IP-immunoblot experiments (data not shown). Nevertheless, we were able to show that endogenous TRRAP protein is complexed with β-catenin in 293 cells in which endogenous β-catenin was transiently stabilized by treatment with the GSK3β inhibitor lithium chloride (Fig. 7 C).

Based on the evidence that TIP49 binds to β-catenin in protein complexes that likely include the TRRAP and BAF53 proteins and our prior results implicating TIP49 in the regulation of expression of β-catenin/TCF target genes, we sought to determine whether we could find evidence that TIP49 and associated proteins were specifically tethered to TCF sites in regulatory regions of β-catenin/TCF-regulated genes. We carried out ChIP studies assessing the ability of antibodies against TIP49, TRRAP, TCF-4, and the TIP60 HAT protein to immunoprecipitate sequences containing a previously implicated TCF site in the human ITF-2 promoter region (Fig. 7,D). As shown in Fig. 7,E, in ChIP studies of formaldehyde cross-linked chromatin preparations isolated from 293 cells in which endogenous β-catenin was stabilized by treatment with the GSK3β inhibitor lithium chloride, the control IgG antibody yielded no specific PCR product for the region containing the TCF site. In contrast, DNA was specifically recovered in the ChIPs with antibodies against TIP49 as well as with antibodies against TCF-4, TIP60, and TRRAP. Further evidence of the specificity of the ChIP of TIP49 and the TIP49-associated proteins with the TCF site in the ITF-2 promoter region was obtained by showing that DNA sequences located at the downstream coding region in the ITF-2 gene were not specifically recovered in ChIPs (Fig. 7 E).

Expression of TIP49D302N Inhibits Histone Acetylation in the Proximal ITF-2 Promoter Region.

The findings above suggest that TIP49 and associated proteins, such as TRRAP and the TIP60 HAT factor, may facilitate β-catenin/TCF-mediated transcription via effects on chromatin structure and activity. As such, we sought to determine, using ChIP assays, whether the TIP49D302N mutant protein could affect chromatin modification when recruited by β-catenin to TCF target gene promoters, such as the TCF site in the proximal promoter region of the rat ITF-2 gene. Chromatin was prepared from RK3E/Neo, RK3E/TIP49, and RK3E/TIP49D302N cells, all of which had been stably transduced to express S33Yβ-catenin protein ectopically. Evidence was obtained that TIP49 specifically interacted in ChIP assays with the TCF site in the proximal promoter region of the rat ITF-2 gene, but not with nonspecific upstream or downstream sequences at the ITF-2 locus (Fig. 8,B). In ChIPs using antibodies against acetylated histone H4, we found that histone acetylation of the ITF-2 promoter region covering the TCF-binding site at −565 was greatly reduced in RK3E/TIP49D302N cells, when compared with the acetylation state of the ITF-2 promoter region in control RK3E cells and cells stably expressing wild-type TIP49 (Fig. 8,C). Expression of the TIP49D302N protein also seemed to moderately reduce histone acetylation in other regions of the ITF-2 gene (Fig. 8 C, upstream and downstream). The results suggest that TIP49, when recruited by β-catenin to promoters of TCF-target genes, affects histone acetylation of nucleosomes nearby to where it is recruited. Notably, histone acetylation is a modification that is generally correlated with increased transcriptional activity (63, 64, 65). TIP49 may exert effects on histone acetylation through its ATP-dependent chromatin remodeling activity and/or, perhaps more likely, by its ability to recruit other cofactors that directly alter the acetylation state of histones.

The findings presented here implicate TIP49 as an important cofactor in activation of β-catenin/TCF-regulated genes and β-catenin-mediated neoplastic transformation. We found that expression of a presumptive dominant negative allele of TIP49 (TIP49D302N) inhibited β-catenin-induced neoplastic transformation of RK3E cells and suppressed the anchorage-independent growth properties of a human colon carcinoma line with defective β-catenin regulation. Additionally, the TIP49D302N mutant protein inhibited activation of endogenous β-catenin/TCF target genes. Using siRNA-based methods to inhibit endogenous TIP49 function, we obtained further support for the role of TIP49 in regulating β-catenin/TCF target gene expression. Using IP-Western blot approaches, TIP49 was found in complexes with chromatin remodeling and histone modifying factors and cofactors, including the TIP60 histone acetylase-associated proteins TRRAP and BAF53. In ChIP assays, the TIP49, TIP60, and TRRAP proteins were found to interact with sequences in the regulatory region of the gene for ITF-2, a TCF-dependent cellular gene. Moreover, the ability of TIP49D302N to inhibit ITF-2 gene expression was linked to decreased acetylation of histones in the vicinity of the TCF-binding sites in the ITF-2 promoter region. Based on these data, we suggest that TIP49 regulates β-catenin-mediated TCF-target gene expression via effects on chromatin structure and activity.

Although TIP49 was initially identified in large part because of its binding to various nuclear factors, the functional significance of the interaction between TIP49 and many of the presumptive TIP49 partner proteins has not been well established. For instance, the results of Wood et al.(39) suggested that TIP49 played a crucial role in c-Myc-mediated oncogenic transformation. However, in contrast to the data on the function of TIP49 in activation of β-catenin/TCF-regulated target genes, no direct role for TIP49 in transcriptional activation of c-Myc reporter genes or endogenous c-Myc-regulated target genes has yet been provided in the literature. At this point, it is not clear whether the inability to demonstrate a role for TIP49 in the regulation of c-Myc-dependent gene expression may be largely due to the transient reporter gene assays performed in the initial studies (39). As we have found, the effects of TIP49 on β-catenin/TCF transcription are very modest in transient reporter assays, where model reporter gene constructs are not likely to be present in the higher order chromatin structures characteristic of endogenous cellular genes. Nevertheless, in light of the prior observations on the role of TIP49 in c-Myc-mediated neoplastic transformation and the fact that the c-MYC gene has been identified as a downstream target of the Wnt/β-catenin/TCF pathway in human colorectal cancer (15), it could be hypothesized that TIP49 functions in β-catenin-mediated transformation by regulating c-Myc expression and/or activity. We have largely excluded this possibility. First, the TIP49D302N mutant protein had no effect on c-Myc expression in RK3E cells (data not shown). Second, our previous studies have indicated that β-catenin-induced neoplastic transformation of RK3E cells is independent of c-Myc expression and activity (52). Moreover, in contrast to the rather unresolved situation regarding the effects of TIP49 on the function of c-MYC in gene regulation, the ability of β-catenin/TCF to activate transcription of endogenous target genes, such as Axil/AXIN2 and ITF-2, appears to be critically dependent on TIP49 function.

In our work, we have used a TIP49D302N mutant protein to assess the functional role of TIP49 in β-catenin-mediated transformation and transcriptional activation. As described above, TIP49D302N contains a missense mutation in the Walker B motif (DEVH→NEVH), which is expected to abolish the ATPase activity of TIP49, as has been shown for the bacterial RuvB protein (57). The mutant TIP49 protein can still form a complex with β-catenin and TIP48 (39), but it cannot bind and hydrolyze ATP. Thus, TIP49D302N likely functions in a dominant negative manner, and this same mutant form of TIP49 is the one used to demonstrate a role for TIP49 in c-Myc-mediated neoplastic transformation (39). The involvement of ATPase activity in the function of RuvB-like proteins is also supported by recent study showing that the liebeskummer mutation in the zebrafish homologue of Reptin (also known as TIP48) increases Reptin/TIP48 ATPase activity and enhances its repressive effect on β-catenin-mediated transcription. The mutant allele has been shown to promote cardiac hyperplasia during zebrafish development (66), and, in the same study, it was shown that reduction in TIP49 levels during zebrafish development essentially phenocopies the effects of the liebeskummer mutation in Reptin. More definitive insights into why the TIP48/reptin protein appears to antagonize β-catenin/TCF transcription and the TIP49 protein appears to stimulate β-catenin/TCF transcription await the results of additional studies.

Interestingly, the TIP49D302N mutant in our study had only very modest inhibitory effects on β-catenin-induced activation of a transiently transfected TCF-responsive reporter gene. Perhaps consistent with this result, as noted above, TIP49D302N also showed no effect on c-Myc-mediated transcription of several different c-Myc response reporter constructs (39). The difference we observed in the magnitude of the effect of TIP49D302N on transcription of a transiently introduced reporter template versus chromatinized templates might be attributable to the requirement of TIP49 DNA remodeling activity for transcription. As an ATP-dependent DNA helicase, TIP49 has been identified in several chromatin remodeling complexes. Transiently introduced reporter gene templates may not be assembled into a phased array of nucleosomes. As such, significant alterations in chromatin structure may not be required to allow access of trans-acting factors to the promoter region of the reporter gene template. It should be noted that significant differences in the transcriptional activity of episomal versus chromosomal templates have been observed previously in the case of glucocorticoid receptor-mediated transcription. In particular, in contrast to stable chromatin templates, a transiently introduced mouse mammary tumor virus reporter is activated by glucocorticoid receptor in a manner that is independent of the hBrg-1/BAF chromatin remodeling complex (67).

In light of these results with glucocorticoid receptor-mediated transcription and the results presented here, we would propose that in the absence of Wnt signaling or other β-catenin stabilizing factors, TCF target genes are in a “closed” chromatin structure that is imposed by Groucho corepressor proteins and their associated histone deacetylases. To activate TCF-regulated target genes, β-catenin may act to relax chromatin structure by recruiting chromatin modifying and remodeling cofactors. In support of this notion, previous studies have shown that β-catenin interacts with the chromatin remodeling factor Brg-1 (33) and HAT p300/CBP (28, 29, 30), and β-catenin appears to require their activities for activation of certain β-catenin/TCF target genes. Indeed, recent studies with the Cyclin D2 gene, a presumptive Wnt/β-cateninTCF pathway downstream gene in pituitary development, showed that specific coactivators, including p300/CBP, TIP60, and TRRAP, were recruited to the Cyclin D2 promoter in a temporally specific manner (68). Also of interest with regard to mechanisms β-catenin may use to activate gene expression are findings on the recently discovered protein called pygopus. Pygopus is a PHD domain-containing protein that has been shown to complex with β- catenin and regulate activation of β-catenin/TCF target genes (34, 35, 36). Because many PHD domain-containing proteins also have other domains whose function is linked to chromatin (37, 38), the presence of a PHD domain in pygopus might predict a role for pygopus in chromatin-related functions. Undoubtedly, future studies will illuminate the complex interplay of interactions between β-catenin, TCFs, TIP49, and the various other chromatin remodeling proteins and associated cofactors that have been implicated in regulation of β-catenin/TCF target gene transcription. Although perhaps an obvious point, the studies and their likely results have major implications for our understanding of the role of Wnt signaling in normal and cancer cells.

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.

Grant support: NIH Grant CA85463.

Note: The present address for N. Lee is DNA Software Inc., Ann Arbor, Michigan 48104.

Requests for reprints: Eric R. Fearon, Division of Molecular Medicine and Genetics, University of Michigan Medical Center, 4301 MSRB III, 1150 West Medical Center Drive, Ann Arbor, Michigan 48109. Phone: (734) 764-1549; Fax: (734) 647-7979; E-mail: [email protected]

6

See http://www.stanford.edu/∼rnusse/pathways/targets.html for a more complete listing of candidate Wnt/β-catenin/TCF target genes.

Fig. 1.

Effect of TIP49 on β-catenin-mediated activation of T-cell factor (TCF) transcription in transient assays with the TOPFLASH reporter. A, schematic diagram of wild-type and mutant (D302N) TIP49. Walker A and Walker B motifs are indicated, as well as the substitution of asparagine (N) for aspartic acid (D) in the Walker B motif in the TIP49D302N mutant. B, RK3E cells were transiently transfected with pcDNA3 expression vectors for the S33Yβ-catenin mutant (S33Y), wild-type TIP49, and TIP49D302N in the different combinations indicated. The ratio of luciferase activities from the cotransfected TCF-responsive reporter (pTOPFLASH) and the control luciferase reporter gene construct (pFOPFLASH) was determined 48 h after transfection. In all assays, the mass of DNA transfected was kept constant by cotransfection with empty vector (vec). The means and SDs of three independent experiments are shown.

Fig. 1.

Effect of TIP49 on β-catenin-mediated activation of T-cell factor (TCF) transcription in transient assays with the TOPFLASH reporter. A, schematic diagram of wild-type and mutant (D302N) TIP49. Walker A and Walker B motifs are indicated, as well as the substitution of asparagine (N) for aspartic acid (D) in the Walker B motif in the TIP49D302N mutant. B, RK3E cells were transiently transfected with pcDNA3 expression vectors for the S33Yβ-catenin mutant (S33Y), wild-type TIP49, and TIP49D302N in the different combinations indicated. The ratio of luciferase activities from the cotransfected TCF-responsive reporter (pTOPFLASH) and the control luciferase reporter gene construct (pFOPFLASH) was determined 48 h after transfection. In all assays, the mass of DNA transfected was kept constant by cotransfection with empty vector (vec). The means and SDs of three independent experiments are shown.

Close modal
Fig. 2.

RK3E focus formation by S33Yβ-catenin is inhibited by TIP49D302N. A, G418-resistant polyclonal RK3E lines stably expressing wild-type TIP49 (TIP49wt), TIP49D302N, or a control empty vector (Neo) were infected with retroviruses encoding S33Yβ-catenin (S33Y) or the control retroviral vector (vec). Four weeks after infection, the plates were fixed with glutaraldehyde and stained with methylene blue, and the foci were photographed. B, macroscopic foci formed by the cells described in A were counted, and the mean numbers of three independent experiments ± SD are shown. C, Western blot analysis of TIP49 protein expression in the polyclonal RK3E lines stably expressing TIP49wt and TIP49D302N. Enhanced chemiluminescence-Western blot analysis with an anti-FLAG antibody was carried out on whole-cell lysates to detect expression of the exogenous FLAG epitope-tagged TIP49 proteins. Equal loading and transfer of proteins were confirmed by stripping the blots and reprobing with an anti-β-actin antibody.

Fig. 2.

RK3E focus formation by S33Yβ-catenin is inhibited by TIP49D302N. A, G418-resistant polyclonal RK3E lines stably expressing wild-type TIP49 (TIP49wt), TIP49D302N, or a control empty vector (Neo) were infected with retroviruses encoding S33Yβ-catenin (S33Y) or the control retroviral vector (vec). Four weeks after infection, the plates were fixed with glutaraldehyde and stained with methylene blue, and the foci were photographed. B, macroscopic foci formed by the cells described in A were counted, and the mean numbers of three independent experiments ± SD are shown. C, Western blot analysis of TIP49 protein expression in the polyclonal RK3E lines stably expressing TIP49wt and TIP49D302N. Enhanced chemiluminescence-Western blot analysis with an anti-FLAG antibody was carried out on whole-cell lysates to detect expression of the exogenous FLAG epitope-tagged TIP49 proteins. Equal loading and transfer of proteins were confirmed by stripping the blots and reprobing with an anti-β-actin antibody.

Close modal
Fig. 3.

K-ras-mediated colony formation in soft agar is not inhibited by TIP49D302N. A, a control polyclonal G418-resistant RK3E line (Neo) and a G418-resistant polyclonal RK3E line stably expressing TIP49D302N were infected with retroviruses encoding mutant K-ras, and colony formation in soft agar was carried out by plating 104 cells of each line in 0.3% agar medium over agar underlayers. After incubation for 4 weeks, the cells were stained with methylene blue, and the colonies were photographed. B, the colonies formed as described in A were counted, and the mean numbers of three independent experiments ± SD are shown.

Fig. 3.

K-ras-mediated colony formation in soft agar is not inhibited by TIP49D302N. A, a control polyclonal G418-resistant RK3E line (Neo) and a G418-resistant polyclonal RK3E line stably expressing TIP49D302N were infected with retroviruses encoding mutant K-ras, and colony formation in soft agar was carried out by plating 104 cells of each line in 0.3% agar medium over agar underlayers. After incubation for 4 weeks, the cells were stained with methylene blue, and the colonies were photographed. B, the colonies formed as described in A were counted, and the mean numbers of three independent experiments ± SD are shown.

Close modal
Fig. 4.

Colony formation in soft agar by the human colorectal cancer cell line LoVo is inhibited by expression of TIP49D302N. A, colony formation assays in LoVo cells stably expressing either wild-type TIP49 (TIP49wt) or TIP49D302N mutant and the G418-resistant LoVo control (Neo) were carried out as described in the Fig. 3 legend. B, colonies formed as described in A were counted, and the mean numbers of three independent experiments ± SD are shown.

Fig. 4.

Colony formation in soft agar by the human colorectal cancer cell line LoVo is inhibited by expression of TIP49D302N. A, colony formation assays in LoVo cells stably expressing either wild-type TIP49 (TIP49wt) or TIP49D302N mutant and the G418-resistant LoVo control (Neo) were carried out as described in the Fig. 3 legend. B, colonies formed as described in A were counted, and the mean numbers of three independent experiments ± SD are shown.

Close modal
Fig. 5.

The TIP49D302N mutant inhibits expression of endogenous β-catenin/T-cell factor (TCF) target genes in RK3E cells. A, expression of TIP49D302N inhibits ITF-2 and Axil induction by S33Yβ-catenin. A control G418-resistant RK3E cell line (RK3E/Neo) and the G418-resistant RK3E line stably expressing the TIP49D302N mutant (RK3E/TIPD302N) were infected with the retrovirus encoding S33Yβ-catenin (S33Y) or the control retroviral empty vector (vec). Cells were harvested at the indicated time points, and total RNA was isolated. Northern blot analysis was carried out sequentially with rat ITF-2, Axil, and GAPDH cDNA probes. B, expression of the exogenous FLAG epitope-tagged S33Yβ-catenin (S33Y) and FLAG epitope-tagged TIP49 in RK3E/Neo (Neo) and RK3E/TIPD302N (D302N) cells after infection with the retrovirus encoding S33Yβ-catenin. Enhanced chemiluminescence-Western blot analysis with an anti-FLAG antibody (for S33Yβ-catenin and TIP49) was carried out on whole-cell lysates prepared at various time points after infection. Equal loading and transfer of proteins were confirmed by stripping the blots and reprobing with an anti-actin antibody. C, quantitative analysis of the intensities of the gel bands from the Northern blot shown in A. The signal strength detected for RK3E/Neo cells at day 6 postinfection was arbitrarily set as 1 unit. D, Northern blot analyses of ITF-2 and Axil expression were carried out on total RNA from parental RK3E cells, β-catenin-transformed RK3E lines (RK3E/ΔN47, RK3E/S33Y-B, and RK3E/S33Y-C) also stably expressing either wild-type TIP49 (TIP49) or the TIP49D302N mutant (D302N), and the control G418-resistant cell line (Neo). After obtaining the autoradiograph for ITF-2 and Axil, the blot was stripped and rehybridized to a GAPDH cDNA probe to control for loading and transfer of RNA. E, expression of the FLAG epitope-tagged S33Yβ-catenin (S33Y) and TIP49 in the various β-catenin-transformed RK3E lines used for the Northern blot above. Enhanced chemiluminescence-Western blot analysis was carried out as described in B. F, PhosphorImager analysis of Northern blots shown in D. The signal strength detected for each control G418-resistant β-catenin-transformed RK3E line (Neo) was arbitrarily set as 1 unit.

Fig. 5.

The TIP49D302N mutant inhibits expression of endogenous β-catenin/T-cell factor (TCF) target genes in RK3E cells. A, expression of TIP49D302N inhibits ITF-2 and Axil induction by S33Yβ-catenin. A control G418-resistant RK3E cell line (RK3E/Neo) and the G418-resistant RK3E line stably expressing the TIP49D302N mutant (RK3E/TIPD302N) were infected with the retrovirus encoding S33Yβ-catenin (S33Y) or the control retroviral empty vector (vec). Cells were harvested at the indicated time points, and total RNA was isolated. Northern blot analysis was carried out sequentially with rat ITF-2, Axil, and GAPDH cDNA probes. B, expression of the exogenous FLAG epitope-tagged S33Yβ-catenin (S33Y) and FLAG epitope-tagged TIP49 in RK3E/Neo (Neo) and RK3E/TIPD302N (D302N) cells after infection with the retrovirus encoding S33Yβ-catenin. Enhanced chemiluminescence-Western blot analysis with an anti-FLAG antibody (for S33Yβ-catenin and TIP49) was carried out on whole-cell lysates prepared at various time points after infection. Equal loading and transfer of proteins were confirmed by stripping the blots and reprobing with an anti-actin antibody. C, quantitative analysis of the intensities of the gel bands from the Northern blot shown in A. The signal strength detected for RK3E/Neo cells at day 6 postinfection was arbitrarily set as 1 unit. D, Northern blot analyses of ITF-2 and Axil expression were carried out on total RNA from parental RK3E cells, β-catenin-transformed RK3E lines (RK3E/ΔN47, RK3E/S33Y-B, and RK3E/S33Y-C) also stably expressing either wild-type TIP49 (TIP49) or the TIP49D302N mutant (D302N), and the control G418-resistant cell line (Neo). After obtaining the autoradiograph for ITF-2 and Axil, the blot was stripped and rehybridized to a GAPDH cDNA probe to control for loading and transfer of RNA. E, expression of the FLAG epitope-tagged S33Yβ-catenin (S33Y) and TIP49 in the various β-catenin-transformed RK3E lines used for the Northern blot above. Enhanced chemiluminescence-Western blot analysis was carried out as described in B. F, PhosphorImager analysis of Northern blots shown in D. The signal strength detected for each control G418-resistant β-catenin-transformed RK3E line (Neo) was arbitrarily set as 1 unit.

Close modal
Fig. 6.

TIP49 siRNA reduces TIP49 protein levels and the transcription of endogenous β-catenin/T-cell factor (TCF) target genes. In A, Lanes 1–4, as indicated, 293 cells were transiently transfected with TIP49 siRNA (Si) or the control scrambled siRNA (Sc), together with pcDNA3 expression vectors for the S33Yβ-catenin mutant (S33Y) or the empty vector (vec). In Lane 5, 293 cells were transfected with S33Yβ-catenin and TIP49D302N (D302N) mutant, together with the scrambled siRNA. Whole cell lysates, prepared 48 h after transfection, were subjected to Western blot with a polyclonal antibody against TIP49 (for endogenous TIP49) or anti-FLAG M2 antibody (for expression of TIP49D302N and S33Yβ-catenin). Blots were stripped and incubated with a rabbit polyclonal antibody against β-actin to verify equal loading of the samples. B, 293 cells were transiently transfected as described in A. Cells were harvested 48 h after transfection, and total RNA was subjected to Northern blot analysis sequentially with human ITF-2, Axin2, TIP49, and GAPDH cDNA probes. C, PhosphorImager analysis of Northern blots shown in B. The signal strength detected for cells transfected with the empty vector plus the control siRNA was arbitrarily set as 1 unit.

Fig. 6.

TIP49 siRNA reduces TIP49 protein levels and the transcription of endogenous β-catenin/T-cell factor (TCF) target genes. In A, Lanes 1–4, as indicated, 293 cells were transiently transfected with TIP49 siRNA (Si) or the control scrambled siRNA (Sc), together with pcDNA3 expression vectors for the S33Yβ-catenin mutant (S33Y) or the empty vector (vec). In Lane 5, 293 cells were transfected with S33Yβ-catenin and TIP49D302N (D302N) mutant, together with the scrambled siRNA. Whole cell lysates, prepared 48 h after transfection, were subjected to Western blot with a polyclonal antibody against TIP49 (for endogenous TIP49) or anti-FLAG M2 antibody (for expression of TIP49D302N and S33Yβ-catenin). Blots were stripped and incubated with a rabbit polyclonal antibody against β-actin to verify equal loading of the samples. B, 293 cells were transiently transfected as described in A. Cells were harvested 48 h after transfection, and total RNA was subjected to Northern blot analysis sequentially with human ITF-2, Axin2, TIP49, and GAPDH cDNA probes. C, PhosphorImager analysis of Northern blots shown in B. The signal strength detected for cells transfected with the empty vector plus the control siRNA was arbitrarily set as 1 unit.

Close modal
Fig. 7.

Interaction of TIP49 with chromatin-associated proteins and T-cell factor (TCF) sites. A, the TIP49D302N mutant interacts with β-catenin to the same extent as wild-type TIP49. Lysates of LoVo cells stably expressing empty vector (vec), FLAG-tagged wild-type TIP49, or TIP49D302N mutant (D302N) were prepared and immunoprecipitated with anti-FLAG polyclonal antibody (Lanes 4–6). Precipitated proteins were separated through SDS-PAGE and Western blotted for β-catenin using anti-β-catenin antibody (top panel) or for TIP49 using anti-FLAG monoclonal antibody (bottom panel). Lanes 1–3, Western blotting of the whole-cell lysates from the LoVo stable lines used for immunoprecipitation (IP) in Lanes 4–6. B, TIP49 interacts with BAF53 and β-catenin. Human colorectal cancer cells HCT116 and SW480 stably expressing FLAG-tagged TIP49 and the corresponding G418-resistant control cells (Neo) were lysed and subjected to IP with anti-FLAG M2 affinity gel (Lanes 3, 4, 7, and 8). Precipitated proteins were Western blotted for β-catenin using anti-β-catenin antibody (top panel) or for BAF53 using anti-BAF53 polyclonal antibody (bottom panel). Lanes 1, 2, 5, and 6 represent Western blotting of whole-cell lysates from the corresponding HCT116 and SW480 cells used for IP. C, β-catenin interacts with TRRAP. 293 cells stably expressing FLAG-tagged TIP49 were treated with 40 mm LiCl for 1 h. Whole cell lysates were collected and subjected to IP with a rabbit polyclonal antibody against TRRAP (Lane 3) or a polyclonal anti-polyhistidine antibody (α-His) as negative control (Lane 2). Precipitated proteins were resolved by SDS-PAGE and immunoblotted with anti-β-catenin antibody. D, schematic diagram of human ITF-2 promoter region, indicating a previously identified TCF-binding site at –1874 (□) and the positions of forward (F) and reverse (R) PCR primers (represented as arrows) used to detect the ITF-2 promoter region in the ChIP assay shown in E. Note that the transcription start site for human ITF-2 has not been defined and that the consensus TCF-binding site is indicated relative to most 5′ nucleotide present in the human ITF-2B cDNA sequence in the GenBank accession. E, occupancy of the human ITF-2 promoter by TIP49 and associated proteins. Chromatin collected from 293 cells, after treatment with 40 mm LiCl for 1 h, was subjected to IP with control IgG or anti-TCF-4 monoclonal antibody or polyclonal antibodies to TIP49, TIP60, and TRRAP. Recovered DNA was used as template in PCR amplification using the pair of primers (as described in D) that flank the TCF-binding site at −1874. As negative controls, distal sequences were detected with primers corresponding to a downstream coding region of ITF-2. Input DNA used for PCR corresponded to 0.1% of total chromatin DNA for immunoprecipitation.

Fig. 7.

Interaction of TIP49 with chromatin-associated proteins and T-cell factor (TCF) sites. A, the TIP49D302N mutant interacts with β-catenin to the same extent as wild-type TIP49. Lysates of LoVo cells stably expressing empty vector (vec), FLAG-tagged wild-type TIP49, or TIP49D302N mutant (D302N) were prepared and immunoprecipitated with anti-FLAG polyclonal antibody (Lanes 4–6). Precipitated proteins were separated through SDS-PAGE and Western blotted for β-catenin using anti-β-catenin antibody (top panel) or for TIP49 using anti-FLAG monoclonal antibody (bottom panel). Lanes 1–3, Western blotting of the whole-cell lysates from the LoVo stable lines used for immunoprecipitation (IP) in Lanes 4–6. B, TIP49 interacts with BAF53 and β-catenin. Human colorectal cancer cells HCT116 and SW480 stably expressing FLAG-tagged TIP49 and the corresponding G418-resistant control cells (Neo) were lysed and subjected to IP with anti-FLAG M2 affinity gel (Lanes 3, 4, 7, and 8). Precipitated proteins were Western blotted for β-catenin using anti-β-catenin antibody (top panel) or for BAF53 using anti-BAF53 polyclonal antibody (bottom panel). Lanes 1, 2, 5, and 6 represent Western blotting of whole-cell lysates from the corresponding HCT116 and SW480 cells used for IP. C, β-catenin interacts with TRRAP. 293 cells stably expressing FLAG-tagged TIP49 were treated with 40 mm LiCl for 1 h. Whole cell lysates were collected and subjected to IP with a rabbit polyclonal antibody against TRRAP (Lane 3) or a polyclonal anti-polyhistidine antibody (α-His) as negative control (Lane 2). Precipitated proteins were resolved by SDS-PAGE and immunoblotted with anti-β-catenin antibody. D, schematic diagram of human ITF-2 promoter region, indicating a previously identified TCF-binding site at –1874 (□) and the positions of forward (F) and reverse (R) PCR primers (represented as arrows) used to detect the ITF-2 promoter region in the ChIP assay shown in E. Note that the transcription start site for human ITF-2 has not been defined and that the consensus TCF-binding site is indicated relative to most 5′ nucleotide present in the human ITF-2B cDNA sequence in the GenBank accession. E, occupancy of the human ITF-2 promoter by TIP49 and associated proteins. Chromatin collected from 293 cells, after treatment with 40 mm LiCl for 1 h, was subjected to IP with control IgG or anti-TCF-4 monoclonal antibody or polyclonal antibodies to TIP49, TIP60, and TRRAP. Recovered DNA was used as template in PCR amplification using the pair of primers (as described in D) that flank the TCF-binding site at −1874. As negative controls, distal sequences were detected with primers corresponding to a downstream coding region of ITF-2. Input DNA used for PCR corresponded to 0.1% of total chromatin DNA for immunoprecipitation.

Close modal
Fig. 8.

Effects of TIP49 expression on histone acetylation in the ITF-2 promoter region. A, schematic of the rat ITF-2 promoter region. The open box represents a previously identified TCF-binding site at −565. The arrows indicate the positions of forward (F) and reverse (R) PCR primers used to detect the area of TCF-binding site in ChIP assay. B, TIP49 is present at a consensus TCF-binding site in the rat ITF-2 promoter. Chromatin was collected from the RK3E/TIP49 cells stably expressing the S33Yβ-catenin mutant and subjected to immunoprecipitation with polyclonal antibodies to TIP49. PCR was performed using the pair of primers that flank the TCF-binding site at −565 in the rat ITF-2 promoter or the primers flanking the region 1 kb upstream of the TCF-binding site or a downstream coding region in ITF-2 as negative controls. Input DNA (0.1% of total chromatin DNA used for immunoprecipitation) and minus (−) antibody controls were also included. C, expression of TIP49D302N interferes with acetylation of histones in the vicinity of the ITF-2 TCF-binding site. RK3E/Neo (Neo), RK3E/TIP49 (TIP49), and RK3E/TIP49D302N (D302N) cells were infected with the retrovirus encoding the S33Yβ-catenin mutant. After a week of drug selection of the infected cells in puromycin, soluble chromatin was prepared from these cells and immunoprecipitated with antibodies against acetylated histone H4 (αAcH4). PCR was carried out as described in B.

Fig. 8.

Effects of TIP49 expression on histone acetylation in the ITF-2 promoter region. A, schematic of the rat ITF-2 promoter region. The open box represents a previously identified TCF-binding site at −565. The arrows indicate the positions of forward (F) and reverse (R) PCR primers used to detect the area of TCF-binding site in ChIP assay. B, TIP49 is present at a consensus TCF-binding site in the rat ITF-2 promoter. Chromatin was collected from the RK3E/TIP49 cells stably expressing the S33Yβ-catenin mutant and subjected to immunoprecipitation with polyclonal antibodies to TIP49. PCR was performed using the pair of primers that flank the TCF-binding site at −565 in the rat ITF-2 promoter or the primers flanking the region 1 kb upstream of the TCF-binding site or a downstream coding region in ITF-2 as negative controls. Input DNA (0.1% of total chromatin DNA used for immunoprecipitation) and minus (−) antibody controls were also included. C, expression of TIP49D302N interferes with acetylation of histones in the vicinity of the ITF-2 TCF-binding site. RK3E/Neo (Neo), RK3E/TIP49 (TIP49), and RK3E/TIP49D302N (D302N) cells were infected with the retrovirus encoding the S33Yβ-catenin mutant. After a week of drug selection of the infected cells in puromycin, soluble chromatin was prepared from these cells and immunoprecipitated with antibodies against acetylated histone H4 (αAcH4). PCR was carried out as described in B.

Close modal

We thank Drs. Bert Vogelstein, Gerald Crabtree, and Gary Nabel for generously providing reagents and Ira Winer and David Van Mater for providing comments on the manuscript.

1
Bienz M., Clevers H. Linking colorectal cancer to Wnt signaling.
Cell
,
103
:
311
-320,  
2000
.
2
Peifer M., Polakis P. Wnt signaling in oncogenesis and embryogenesis: a look outside the nucleus.
Science (Wash. DC)
,
287
:
1606
-1609,  
2000
.
3
Morin P. J. β-Catenin signaling and cancer.
Bioessays
,
21
:
1021
-1030,  
1999
.
4
Polakis P. Wnt signaling and cancer.
Genes Dev.
,
14
:
1837
-1851,  
2000
.
5
Aberle H., Bauer A., Stappert J., Kispert A., Kemler R. β-Catenin is a target for the ubiquitin-proteasome pathway.
EMBO J.
,
16
:
3797
-3804,  
1997
.
6
Behrens J., Jerchow B. A., Wurtele M., Grimm J., Asbrand C., Wirtz R., Kuhl M., Wedlich D., Birchmeier W. Functional interaction of an axin homolog, conductin, with β-catenin, APC, and GSK3β.
Science (Wash. DC)
,
280
:
596
-599,  
1998
.
7
Hart M. J., de los Santos R., Albert I. N., Rubinfeld B., Polakis P. Downregulation of β-catenin by human axin and its association with the APC tumor suppressor, β-catenin and GSK3β.
Curr. Biol.
,
8
:
573
-581,  
1998
.
8
Jiang J., Struhl G. Regulation of the Hedgehog and Wingless signalling pathways by the F- box/WD40-repeat protein Slimb.
Nature (Lond.)
,
391
:
493
-496,  
1998
.
9
Zeng L., Fagotto F., Zhang T., Hsu W., Vasicek T. J., Perry W. L., III, Lee J. J., Tilghman S. M., Gumbiner B. M., Costantini F. The mouse Fused locus encodes Axin, an inhibitor of the Wnt signaling pathway that regulates embryonic axis formation.
Cell
,
90
:
181
-192,  
1997
.
10
Kishida S., Yamamoto H., Hino S., Ikeda S., Kishida M., Kikuchi A. DIX domains of Dvl and axin are necessary for protein interactions and their ability to regulate β-catenin stability.
Mol. Cell. Biol.
,
19
:
4414
-4422,  
1999
.
11
Ikeda S., Kishida S., Yamamoto H., Murai H., Koyama S., Kikuchi A. Axin, a negative regulator of the Wnt signaling pathway, forms a complex with GSK-3β and β-catenin and promotes GSK-3β-dependent phosphorylation of β-catenin.
EMBO J.
,
17
:
1371
-1384,  
1998
.
12
Fagotto F., Gluck U., Gumbiner B. M. Nuclear localization signal-independent and importin/karyopherin-independent nuclear import of β-catenin.
Curr. Biol.
,
8
:
181
-190,  
1998
.
13
Behrens J., von Kries J. P., Kuhl M., Bruhn L., Wedlich D., Grosschedl R., Birchmeier W. Functional interaction of β-catenin with the transcription factor LEF-1.
Nature (Lond.)
,
382
:
638
-642,  
1996
.
14
Molenaar M., van de Wetering M., Oosterwegel M., Peterson-Maduro J., Godsave S., Korinek V., Roose J., Destree O., Clevers H. XTcf-3 transcription factor mediates β-catenin-induced axis formation in Xenopus embryos.
Cell
,
86
:
391
-399,  
1996
.
15
He T. C., Sparks A. B., Rago C., Hermeking H., Zawel L., da Costa L. T., Morin P. J., Vogelstein B., Kinzler K. W. Identification of c-MYC as a target of the APC pathway.
Science (Wash. DC)
,
281
:
1509
-1512,  
1998
.
16
Shtutman M., Zhurinsky J., Simcha I., Albanese C., D’Amico M., Pestell R., Ben-Ze’ev A. The cyclin D1 gene is a target of the β-catenin/LEF-1 pathway.
Proc. Natl. Acad. Sci. USA
,
96
:
5522
-5527,  
1999
.
17
Tetsu O., McCormick F. B-Catenin regulates expression of cyclin D1 in colon carcinoma cells.
Nature (Lond.)
,
398
:
422
-426,  
1999
.
18
Crawford H. C., Fingleton B. M., Rudolph-Owen L. A., Goss K. J., Rubinfeld B., Polakis P., Matrisian L. M. The metalloproteinase matrilysin is a target of β-catenin transactivation in intestinal tumors.
Oncogene
,
18
:
2883
-2891,  
1999
.
19
Kolligs F. T., Nieman M. N., Winer I., Hu G., Mater D. S. v., Feng Y., Wu R., Cho K. R., Fearon E. R. ITF-2, a downstream target of the Wnt/TCF pathway, is activated in human cancers with β-catenin defects and promotes neoplastic transformation.
Cancer Cells
,
1
:
145
-155,  
2002
.
20
Cavallo R. A., Cox R. T., Moline M. M., Roose J., Polevoy G. A., Clevers H., Peifer M., Bejsovec A. Drosophila Tcf and Groucho interact to repress Wingless signalling activity.
Nature (Lond.)
,
395
:
604
-608,  
1998
.
21
Palaparti A., Baratz A., Stifani S. The Groucho/transducin-like enhancer of split transcriptional repressors interact with the genetically defined amino-terminal silencing domain of histone H3.
J. Biol. Chem.
,
272
:
26604
-26610,  
1997
.
22
Chen G., Fernandez J., Mische S., Courey A. J. A functional interaction between the histone deacetylase Rpd3 and the corepressor groucho in Drosophila development.
Genes Dev.
,
13
:
2218
-2230,  
1999
.
23
Waltzer L., Bienz M. Drosophila CBP represses the transcription factor TCF to antagonize Wingless signalling.
Nature (Lond.)
,
395
:
521
-525,  
1998
.
24
Tago K., Nakamura T., Nishita M., Hyodo J., Nagai S., Murata Y., Adachi S., Ohwada S., Morishita Y., Shibuya H., Akiyama T. Inhibition of Wnt signaling by ICAT, a novel β-catenin-interacting protein.
Genes Dev.
,
14
:
1741
-1749,  
2000
.
25
Sakamoto I., Kishida S., Fukui A., Kishida M., Yamamoto H., Hino S., Michiue T., Takada S., Asashima M., Kikuchi A. A novel β-catenin-binding protein inhibits β-catenin-dependent Tcf activation and axis formation.
J. Biol. Chem.
,
275
:
32871
-32878,  
2000
.
26
Takemaru K., Yamaguchi S., Lee Y. S., Zhang Y., Carthew R. W., Moon R. T. Chibby, a nuclear β-catenin-associated antagonist of the Wnt/Wingless pathway.
Nature (Lond.)
,
422
:
905
-909,  
2003
.
27
Hecht A., Litterst C. M., Huber O., Kemler R. Functional characterization of multiple transactivating elements in β-catenin, some of which interact with the TATA-binding protein in vitro.
J. Biol. Chem.
,
274
:
18017
-18025,  
1999
.
28
Hecht A., Vleminckx K., Stemmler M. P., van Roy F., Kemler R. The p300/CBP acetyltransferases function as transcriptional coactivators of β-catenin in vertebrates.
EMBO J.
,
19
:
1839
-1850,  
2000
.
29
Sun Y., Kolligs F. T., Hottiger M. O., Mosavin R., Fearon E. R., Nabel G. J. Regulation of β-catenin transformation by the p300 transcriptional coactivator.
Proc. Natl. Acad. Sci. USA
,
97
:
12613
-12618,  
2000
.
30
Takemaru K. I., Moon R. T. The transcriptional coactivator CBP interacts with β-catenin to activate gene expression.
J. Cell Biol.
,
149
:
249
-254,  
2000
.
31
van de Wetering M., Cavallo R., Dooijes D., van Beest M., van Es J., Loureiro J., Ypma A., Hursh D., Jones T., Bejsovec A., Peifer M., Mortin M., Clevers H. Armadillo coactivates transcription driven by the product of the Drosophila segment polarity gene dTCF.
Cell
,
88
:
789
-799,  
1997
.
32
Prieve M. G., Waterman M. L. Nuclear localization and formation of β-catenin-lymphoid enhancer factor 1 complexes are not sufficient for activation of gene expression.
Mol. Cell. Biol.
,
19
:
4503
-4515,  
1999
.
33
Barker N., Hurlstone A., Musisi H., Miles A., Bienz M., Clevers H. The chromatin remodelling factor Brg-1 interacts with β-catenin to promote target gene activation.
EMBO J.
,
20
:
4935
-4943,  
2001
.
34
Kramps T., Peter O., Brunner E., Nellen D., Froesch B., Chatterjee S., Murone M., Zullig S., Basler K. Wnt/wingless signaling requires BCL9/legless-mediated recruitment of pygopus to the nuclear β-catenin-TCF complex.
Cell
,
109
:
47
-60,  
2002
.
35
Parker D. S., Jemison J., Cadigan K. M. Pygopus, a nuclear PHD-finger protein required for Wingless signaling in Drosophila.
Development (Camb.)
,
129
:
2565
-2576,  
2002
.
36
Thompson B., Townsley F., Rosin-Arbesfeld R., Musisi H., Bienz M. A new nuclear component of the Wnt signalling pathway.
Nat. Cell Biol.
,
4
:
367
-373,  
2002
.
37
Aasland R., Gibson T. J., Stewart A. F. The PHD finger: implications for chromatin-mediated transcriptional regulation.
Trends Biochem. Sci.
,
20
:
56
-59,  
1995
.
38
Jacobson S., Pillus L. Modifying chromatin and concepts of cancer.
Curr. Opin. Genet. Dev.
,
9
:
175
-184,  
1999
.
39
Wood M. A., McMahon S. B., Cole M. D. An ATPase/helicase complex is an essential cofactor for oncogenic transformation by c-Myc.
Mol. Cell
,
5
:
321
-330,  
2000
.
40
Kanemaki M., Makino Y., Yoshida T., Kishimoto T., Koga A., Yamamoto K., Yamamoto M., Moncollin V., Egly J. M., Muramatsu M., Tamura T. Molecular cloning of a rat 49-kDa TBP-interacting protein (TIP49) that is highly homologous to the bacterial RuvB.
Biochem. Biophys. Res. Commun.
,
235
:
64
-68,  
1997
.
41
Bauer A., Huber O., Kemler R. Pontin52, an interaction partner of β-catenin, binds to the TATA box binding protein.
Proc. Natl. Acad. Sci. USA
,
95
:
14787
-14792,  
1998
.
42
Holzmann K., Gerner C., Korosec T., Poltl A., Grimm R., Sauermann G. Identification and characterization of the ubiquitously occurring nuclear matrix protein NMP 238.
Biochem. Biophys. Res. Commun.
,
252
:
39
-45,  
1998
.
43
Qiu X. B., Lin Y. L., Thome K. C., Pian P., Schlegel B. P., Weremowicz S., Parvin J. D., Dutta A. An eukaryotic RuvB-like protein (RUVBL1) essential for growth.
J. Biol. Chem.
,
273
:
27786
-27793,  
1998
.
44
Shen X., Mizuguchi G., Hamiche A., Wu C. A chromatin remodelling complex involved in transcription and DNA processing.
Nature (Lond.)
,
406
:
541
-544,  
2000
.
45
Ikura T., Ogryzko V. V., Grigoriev M., Groisman R., Wang J., Horikoshi M., Scully R., Qin J., Nakatani Y. Involvement of the TIP60 histone acetylase complex in DNA repair and apoptosis.
Cell
,
102
:
463
-473,  
2000
.
46
Kikuchi N., Gohshi T., Kawahire S., Tachibana T., Yoneda Y., Isobe T., Lim C. R., Kohno K., Ichimura T., Omata S., Horigome T. Molecular shape and ATP binding activity of rat p50, a putative mammalian homologue of RuvB DNA helicase.
J. Biochem. (Tokyo)
,
125
:
487
-494,  
1999
.
47
Kanemaki M., Kurokawa Y., Matsu-ura T., Makino Y., Masani A., Okazaki K., Morishita T., Tamura T. A. TIP49b, a new RuvB-like DNA helicase, is included in a complex together with another RuvB-like DNA helicase, TIP49a.
J. Biol. Chem.
,
274
:
22437
-22444,  
1999
.
48
Bauer A., Chauvet S., Huber O., Usseglio F., Rothbacher U., Aragnol D., Kemler R., Pradel J. Pontin52 and reptin52 function as antagonistic regulators of β-catenin signalling activity.
EMBO J.
,
19
:
6121
-6130,  
2000
.
49
Gohshi T., Shimada M., Kawahire S., Imai N., Ichimura T., Omata S., Horigome T. Molecular cloning of mouse p47, a second group mammalian RuvB DNA helicase-like protein: homology with those from human and Saccharomyces cerevisiae.
J. Biochem. (Tokyo)
,
125
:
939
-946,  
1999
.
50
Makino Y., Kanemaki M., Kurokawa Y., Koji T., Tamura T. A rat RuvB-like protein, TIP49a, is a germ cell-enriched novel DNA helicase.
J. Biol. Chem.
,
274
:
15329
-15335,  
1999
.
51
Jonsson Z. O., Dhar S. K., Narlikar G. J., Auty R., Wagle N., Pellman D., Pratt R. E., Kingston R., Dutta A. Rvb1p and Rvb2p are essential components of a chromatin remodeling complex that regulates transcription of over 5% of yeast genes.
J. Biol. Chem.
,
276
:
16279
-16288,  
2001
.
52
Kolligs F. T., Hu G., Dang C. V., Fearon E. R. Neoplastic transformation of RK3E by mutant β-catenin requires deregulation of Tcf/Lef transcription but not activation of c-myc expression.
Mol. Cell. Biol.
,
19
:
5696
-5706,  
1999
.
53
Littlewood T. D., Hancock D. C., Danielian P. S., Parker M. G., Evan G. I. A modified oestrogen receptor ligand-binding domain as an improved switch for the regulation of heterologous proteins.
Nucleic Acids Res.
,
23
:
1686
-1690,  
1995
.
54
Korinek V., Barker N., Morin P. J., van Wichen D., de Weger R., Kinzler K. W., Vogelstein B., Clevers H. Constitutive transcriptional activation by a β-catenin-Tcf complex in APC−/− colon carcinoma.
Science (Wash. DC)
,
275
:
1784
-1787,  
1997
.
55
Gorbalenya A. E., Koonin E. V., Donchenko A. P., Blinov V. M. Two related superfamilies of putative helicases involved in replication, recombination, repair and expression of DNA and RNA genomes.
Nucleic Acids Res.
,
17
:
4713
-4730,  
1989
.
56
Schmid S. R., Linder P. D-E-A-D protein family of putative RNA helicases.
Mol. Microbiol.
,
6
:
283
-291,  
1992
.
57
Mezard C., Davies A. A., Stasiak A., West S. C. Biochemical properties of RuvBD113N: a mutation in helicase motif II of the RuvB hexamer affects DNA binding and ATPase activities.
J. Mol. Biol.
,
271
:
704
-717,  
1997
.
58
Leung J. Y., Kolligs F. T., Wu R., Zhai Y., Kuick R., Hanash S., Cho K. R., Fearon E. R. Activation of AXIN2 expression by β-catenin/TCF: A feedback repressor pathway regulating Wnt signaling.
J. Biol. Chem.
,
277
:
21657
-21665,  
2002
.
59
Lustig B., Jerchow B., Sachs M., Weiler S., Pietsch T., Karsten U., van de Wetering M., Clevers H., Schlag P. M., Birchmeier W., Behrens J. Negative feedback loop of Wnt signaling through upregulation of conductin/axin2 in colorectal and liver tumors.
Mol. Cell. Biol.
,
22
:
1184
-1193,  
2002
.
60
Jho E. H., Zhang T., Domon C., Joo C. K., Freund J. N., Costantini F. Wnt/β-catenin/Tcf signaling induces the transcription of Axin2, a negative regulator of the signaling pathway.
Mol. Cell. Biol.
,
22
:
1172
-1183,  
2002
.
61
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., Williams L. T. Elevated expression of axin2 and hnkd mRNA provides evidence that Wnt/β-catenin signaling is activated in human colon tumors.
Proc. Natl. Acad. Sci. USA
,
98
:
14973
-14978,  
2001
.
62
Park J., Wood M. A., Cole M. D. BAF53 forms distinct nuclear complexes and functions as a critical c-Myc-interacting nuclear cofactor for oncogenic transformation.
Mol. Cell. Biol.
,
22
:
1307
-1316,  
2002
.
63
Mizzen C. A., Allis C. D. Linking histone acetylation to transcriptional regulation.
Cell Mol. Life Sci.
,
54
:
6
-20,  
1998
.
64
Kornberg R. D., Lorch Y. Chromatin-modifying and -remodeling complexes.
Curr. Opin. Genet. Dev.
,
9
:
148
-151,  
1999
.
65
Kingston R. E., Narlikar G. J. ATP-dependent remodeling and acetylation as regulators of chromatin fluidity.
Genes Dev.
,
13
:
2339
-2352,  
1999
.
66
Rottbauer W., Saurin A. J., Lickert H., Shen X., Burns C. G., Wo Z. G., Kemler R., Kingston R., Wu C., Fishman M. Reptin and pontin antagonistically regulate heart growth in zebrafish embryos.
Cell
,
111
:
661
-672,  
2002
.
67
Fryer C. J., Archer T. K. Chromatin remodelling by the glucocorticoid receptor requires the BRG1 complex.
Nature (Lond.)
,
393
:
88
-91,  
1998
.
68
Kioussi C., Briata P., Baek S. H., Rose D. W., Hamblet N. S., Herman T., Ohgi K. A., Lin C., Gleiberman A., Wang J., Brault V., Ruiz-Lozano P., Nguyen H. D., Kemler R., Glass C. K., Wynshaw-Boris A., Rosenfeld M. G. Identification of a Wnt/Dvl/β-catenin→Pitx2 pathway mediating cell-type-specific proliferation during development.
Cell
,
111
:
673
-685,  
2002
.