Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Chimeric EWSR1-FLI1 regulates the Ewing sarcoma susceptibility gene EGR2 via a GGAA microsatellite

Abstract

Deciphering the ways in which somatic mutations and germline susceptibility variants cooperate to promote cancer is challenging. Ewing sarcoma is characterized by fusions between EWSR1 and members of the ETS gene family, usually EWSR1-FLI1, leading to the generation of oncogenic transcription factors that bind DNA at GGAA motifs1,2,3. A recent genome-wide association study4 identified susceptibility variants near EGR2. Here we found that EGR2 knockdown inhibited proliferation, clonogenicity and spheroidal growth in vitro and induced regression of Ewing sarcoma xenografts. Targeted germline deep sequencing of the EGR2 locus in affected subjects and controls identified 291 Ewing-associated SNPs. At rs79965208, the A risk allele connected adjacent GGAA repeats by converting an interspaced GGAT motif into a GGAA motif, thereby increasing the number of consecutive GGAA motifs and thus the EWSR1-FLI1–dependent enhancer activity of this sequence, with epigenetic characteristics of an active regulatory element. EWSR1-FLI1 preferentially bound to the A risk allele, which increased global and allele-specific EGR2 expression. Collectively, our findings establish cooperation between a dominant oncogene and a susceptibility variant that regulates a major driver of Ewing sarcomagenesis.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: EGR2 overexpression is mediated by EWSR1-FLI1.
Figure 2: EGR2 is critical for the growth and tumorigenicity of Ewing sarcoma.
Figure 3: Fine-mapping and epigenetic profiling revealed candidate EGR2 regulatory elements.
Figure 4: Germline variation at mSat2 modulates EWSR1-FLI1–dependent EGR2 expression.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Referenced accessions

Gene Expression Omnibus

References

  1. Delattre, O. et al. Gene fusion with an ETS DNA-binding domain caused by chromosome translocation in human tumours. Nature 359, 162–165 (1992).

    Article  CAS  PubMed  Google Scholar 

  2. Gangwal, K. et al. Microsatellites as EWS/FLI response elements in Ewing's sarcoma. Proc. Natl. Acad. Sci. USA 105, 10149–10154 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Guillon, N. et al. The oncogenic EWS-FLI1 protein binds in vivo GGAA microsatellite sequences with potential transcriptional activation function. PLoS One 4, e4932 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Postel-Vinay, S. et al. Common variants near TARDBP and EGR2 are associated with susceptibility to Ewing sarcoma. Nat. Genet. 44, 323–327 (2012).

    Article  CAS  PubMed  Google Scholar 

  5. von Levetzow, C. et al. Modeling initiation of Ewing sarcoma in human neural crest cells. PLoS One 6, e19305 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Tirode, F. et al. Mesenchymal stem cell features of Ewing tumors. Cancer Cell 11, 421–429 (2007).

    Article  CAS  PubMed  Google Scholar 

  7. Delattre, O. et al. The Ewing family of tumors—a subgroup of small-round-cell tumors defined by specific chimeric transcripts. N. Engl. J. Med. 331, 294–299 (1994).

    Article  CAS  PubMed  Google Scholar 

  8. Patel, M. et al. Tumor-specific retargeting of an oncogenic transcription factor chimera results in dysregulation of chromatin and transcription. Genome Res. 22, 259–270 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Brohl, A.S. et al. The genomic landscape of the Ewing sarcoma family of tumors reveals recurrent STAG2 mutation. PLoS Genet. 10, e1004475 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Crompton, B.D. et al. The genomic landscape of pediatric Ewing sarcoma. Cancer Discov. 4, 1326–1341 (2014).

    Article  CAS  PubMed  Google Scholar 

  11. Tirode, F. et al. Genomic landscape of Ewing sarcoma defines an aggressive subtype with co-association of STAG2 and TP53 mutations. Cancer Discov. 4, 1342–1353 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Worch, J. et al. Racial differences in the incidence of mesenchymal tumors associated with EWSR1 translocation. Cancer Epidemiol. Biomarkers Prev. 20, 449–453 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Chung, C.C. & Chanock, S.J. Current status of genome-wide association studies in cancer. Hum. Genet. 130, 59–78 (2011).

    Article  PubMed  Google Scholar 

  14. Dominy, J.E. Jr. et al. Discovery and characterization of a second mammalian thiol dioxygenase, cysteamine dioxygenase. J. Biol. Chem. 282, 25189–25198 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Chandra, A., Lan, S., Zhu, J., Siclari, V.A. & Qin, L. Epidermal growth factor receptor (EGFR) signaling promotes proliferation and survival in osteoprogenitors by increasing early growth response 2 (EGR2) expression. J. Biol. Chem. 288, 20488–20498 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Topilko, P. et al. Krox-20 controls myelination in the peripheral nervous system. Nature 371, 796–799 (1994).

    Article  CAS  PubMed  Google Scholar 

  17. Mackintosh, C., Madoz-Gúrpide, J., Ordóñez, J.L., Osuna, D. & Herrero-Martín, D. The molecular pathogenesis of Ewing's sarcoma. Cancer Biol. Ther. 9, 655–667 (2010).

    Article  CAS  PubMed  Google Scholar 

  18. Gao, C. et al. HEFT: eQTL analysis of many thousands of expressed genes while simultaneously controlling for hidden factors. Bioinformatics 30, 369–376 (2014).

    Article  CAS  PubMed  Google Scholar 

  19. Radtke, I. et al. Genomic analysis reveals few genetic alterations in pediatric acute myeloid leukemia. Proc. Natl. Acad. Sci. USA 106, 12944–12949 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Moffatt, M.F. et al. Genetic variants regulating ORMDL3 expression contribute to the risk of childhood asthma. Nature 448, 470–473 (2007).

    Article  CAS  PubMed  Google Scholar 

  21. Northcott, P.A. et al. Subgroup-specific structural variation across 1,000 medulloblastoma genomes. Nature 488, 49–56 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wang, K. et al. Integrative genomics identifies LMO1 as a neuroblastoma oncogene. Nature 469, 216–220 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. GTEx Consortium. The Genotype-Tissue Expression (GTEx) project. Nat. Genet. 45, 580–585 (2013).

  24. Labalette, C. et al. Hindbrain patterning requires fine-tuning of early krox20 transcription by Sprouty 4. Development 138, 317–326 (2011).

    Article  CAS  PubMed  Google Scholar 

  25. Weisinger, K., Kayam, G., Missulawin-Drillman, T. & Sela-Donenfeld, D. Analysis of expression and function of FGF-MAPK signaling components in the hindbrain reveals a central role for FGF3 in the regulation of Krox20, mediated by Pea3. Dev. Biol. 344, 881–895 (2010).

    Article  CAS  PubMed  Google Scholar 

  26. ENCODE Project Consortium. et al. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

  27. Riggi, N. et al. EWS-FLI1 utilizes divergent chromatin remodeling mechanisms to directly activate or repress enhancer elements in Ewing sarcoma. Cancer Cell 26, 668–681 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ernst, J. et al. Mapping and analysis of chromatin state dynamics in nine human cell types. Nature 473, 43–49 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Maurano, M.T. et al. Systematic localization of common disease-associated variation in regulatory DNA. Science 337, 1190–1195 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Ghislain, J. et al. Characterisation of cis-acting sequences reveals a biphasic, axon-dependent regulation of Krox20 during Schwann cell development. Development 129, 155–166 (2002).

    Article  CAS  PubMed  Google Scholar 

  31. 1000 Genomes Project Consortium. et al. An integrated map of genetic variation from 1,092 human genomes. Nature 491, 56–65 (2012).

  32. Robison, L.L. et al. The Childhood Cancer Survivor Study: a National Cancer Institute–supported resource for outcome and intervention research. J. Clin. Oncol. 27, 2308–2318 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Edwards, S.L., Beesley, J., French, J.D. & Dunning, A.M. Beyond GWASs: illuminating the dark road from association to function. Am. J. Hum. Genet. 93, 779–797 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Faye, L.L., Machiela, M.J., Kraft, P., Bull, S.B. & Sun, L. Re-ranking sequencing variants in the post-GWAS era for accurate causal variant identification. PLoS Genet. 9, e1003609 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Surdez, D. et al. Targeting the EWSR1–FLI1 oncogene-induced protein kinase PKC-β abolishes Ewing sarcoma growth. Cancer Res. 72, 4494–4503 (2012).

    Article  CAS  PubMed  Google Scholar 

  36. International HapMap 3 Consortium. et al. Integrating common and rare genetic variation in diverse human populations. Nature 467, 52–58 (2010).

  37. Robinson, J.T. et al. Integrative genomics viewer. Nat. Biotechnol. 29, 24–26 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Carrillo, J. et al. Cholecystokinin down-regulation by RNA interference impairs Ewing tumor growth. Clin. Cancer Res. 13, 2429–2440 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Conrad, C., Gottgens, B., Kinston, S., Ellwart, J. & Huss, R. GATA transcription in a small rhodamine 123(low)CD34(+) subpopulation of a peripheral blood-derived CD34(–)CD105(+) mesenchymal cell line. Exp. Hematol. 30, 887–895 (2002).

    Article  CAS  PubMed  Google Scholar 

  40. Thalmeier, K. et al. Establishment of two permanent human bone marrow stromal cell lines with long-term post irradiation feeder capacity. Blood 83, 1799–1807 (1994).

    Article  CAS  PubMed  Google Scholar 

  41. Wiederschain, D. et al. Single-vector inducible lentiviral RNAi system for oncology target validation. Cell Cycle 8, 498–504 (2009).

    Article  CAS  PubMed  Google Scholar 

  42. Dai, M. et al. Evolving gene/transcript definitions significantly alter the interpretation of GeneChip data. Nucleic Acids Res. 33, e175 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Culhane, A.C., Thioulouse, J., Perrière, G. & Higgins, D.G. MADE4: an R package for multivariate analysis of gene expression data. Bioinformatics 21, 2789–2790 (2005).

    Article  CAS  PubMed  Google Scholar 

  44. Melot, T. et al. Production and characterization of mouse monoclonal antibodies to wild-type and oncogenic FLI-1 proteins. Hybridoma 16, 457–464 (1997).

    Article  CAS  PubMed  Google Scholar 

  45. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Franken, N.A.P., Rodermond, H.M., Stap, J., Haveman, J. & van Bree, C. Clonogenic assay of cells in vitro. Nat. Protoc. 1, 2315–2319 (2006).

    Article  CAS  PubMed  Google Scholar 

  47. Boeva, V. et al. De novo motif identification improves the accuracy of predicting transcription factor binding sites in ChIP-Seq data analysis. Nucleic Acids Res. 38, e126 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Yeager, M. et al. Genome-wide association study of prostate cancer identifies a second risk locus at 8q24. Nat. Genet. 39, 645–649 (2007).

    Article  CAS  PubMed  Google Scholar 

  49. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. McKenna, A. et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Pruim, R.J. et al. LocusZoom: regional visualization of genome-wide association scan results. Bioinformatics 26, 2336–2337 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Barrett, J.C., Fry, B., Maller, J. & Daly, M.J. Haploview: analysis and visualization of LD and haplotype maps. Bioinformatics 21, 263–265 (2005).

    Article  CAS  PubMed  Google Scholar 

  53. Purcell, S. et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Gabriel, S.B. et al. The structure of haplotype blocks in the human genome. Science 296, 2225–2229 (2002).

    Article  CAS  PubMed  Google Scholar 

  55. Clavel-Chapelon, F. et al. E3N, a French cohort study on cancer risk factors. E3N Group. Etude Epidémiologique auprès de femmes de l'Education Nationale. Eur. J. Cancer Prev. 6, 473–478 (1997).

    Article  CAS  PubMed  Google Scholar 

  56. Wang, Z. et al. Improved imputation of common and uncommon SNPs with a new reference set. Nat. Genet. 44, 6–7 (2012).

    Article  CAS  Google Scholar 

  57. Howie, B.N., Donnelly, P. & Marchini, J. A flexible and accurate genotype imputation method for the next generation of genome-wide association studies. PLoS Genet. 5, e1000529 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Boeva, V., Lermine, A., Barette, C., Guillouf, C. & Barillot, E. Nebula—a web-server for advanced ChIP-seq data analysis. Bioinformatics 28, 2517–2519 (2012).

    Article  CAS  PubMed  Google Scholar 

  59. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Meyer, L.R. et al. The UCSC Genome Browser database: extensions and updates 2013. Nucleic Acids Res. 41, D64–D69 (2013).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

T.G.P.G. is supported by a grant from the German Research Foundation (DFG GR3728/2-1), by the Daimler and Benz Foundation in cooperation with the Reinhard Frank Foundation, by LMU Munich's Institutional Strategy LMUexcellent within the framework of the German Excellence Initiative and by the Mehr LEBEN für krebskranke Kinder–Bettina-Bräu-Stiftung. D.S. is supported by the Institut Curie–SIRIC (Site de Recherche Intégrée en Cancérologie) program. F.C.-A. is supported by a grant from the Asociación Pablo Ugarte and Instituto de Salud Carlos III RTICC RD12/0036/0027. D.G.C. is supported by a grant from the InfoSarcomes Association. The Childhood Cancer Survivor Study is supported by the National Cancer Institute (CA55727, G.T. Armstrong), with funding for genotyping from the Intramural Research Program of the National Institutes of Health, National Cancer Institute. This work was supported by grants from the Institut Curie; INSERM; the Agence Nationale de la Recherche (Investissements d'Avenir) (ANR-10-EQPX-03); the Canceropôle Ile-de-France (ANR10-INBS-09-08); the Ligue Nationale Contre le Cancer (Equipe labellisée); the Institut National du Cancer (PLBIO14-237); the European PROVABES (ERA-649 NET TRANSCAN JTC-2011), ASSET (FP7-HEALTH-2010-259348) and EEC (HEALTH-F2-2013-602856) projects; and the Société Française des Cancers de l'Enfant. The following associations supported this work: Courir pour Mathieu, Dans les pas du Géant, Olivier Chape, Les Bagouzamanon, Enfants et Santé, and les Amis de Claire. The Charnay laboratory was financed by INSERM, the CNRS, the Ministère de la Recherche et Technologie, and the Fondation pour la Recherche Médicale. It has received support under the program “Investissements d'Avenir” launched by the French Government and implemented by the ANR, with the references ANR-10-LABX-54 MEMOLIFE and ANR-11-IDEX-0001-02 PSL* Research University. We thank J. Maris for providing genotype information for the neuroblastoma data set, and we thank L. Liang and W. Cookson for providing access to genotype data on the LCL data set. We thank H. Kovar (Children's Cancer Research Institute Vienna, Vienna, Austria) for providing cell lines STA-ET-1, STA-ET-3 and STA-ET-8, and F. Redini (University of Nantes, Nantes, France) for providing cell line TC-32. Human MSC lines L87 and V54-2 were kindly provided by P. Nelson (University Hospital LMU, Munich, Germany). We also thank the following clinicians and pathologists for providing samples used in this work: I. Aerts, P. Anract, C. Bergeron, L. Boccon-Gibod, F. Boman, F. Bourdeaut, C. Bouvier, R. Bouvier, L. Brugiéres, E. Cassagnau, J. Champigneulle, C. Cordonnier, J. M. Coindre, N. Corradini, A. Coulomb-Lhermine, A. De Muret, G. De Pinieux, A.S. Defachelles, A. Deville, F. Dijoud, F. Doz, C. Dufour, K. Fernandez, N. Gaspard, L. Galmiche-Rolland, C. Glorion, A. Gomez-Brouchet, J.M. Guinebretiere, H. Jouan, C. Jeanne-Pasquier, B. Kantelip, F. Labrousse, V. Laithier, F. Larousserie, G. Leverger, C. Linassier, P. Mary, G. Margueritte, E. Mascard, A. Moreau, J. Michon, C. Michot, F. Millot, Y. Musizzano, M. Munzer, B. Narciso, O. Oberlin, D. Orbach, H. Pacquement, Y. Perel, B. Petit, M. Peuchmaur, J.Y. Pierga, C. Piguet, S. Piperno-Neumann, E. Plouvier, D. Ranchere-Vince, J. Rivel, C. Rouleau, H. Rubie, H. Sartelet, G. Schleiermacher, C. Schmitt, N. Sirvent, D. Sommelet, P. Terrier, R. Tichit, J. Vannier, J.M. Vignaud and V. Verkarre. We also thank D. Darmon for technical assistance, S. Grossetete-Lalami for bioinformatic assistance, and V.R. Buchholz and E. Butt for critical reading of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

T.G.P.G. coordinated and designed the study, performed all functional experiments, analyzed the sequencing data, performed bioinformatic analyses, wrote the paper, designed the figures and helped with grant applications. V. Bernard processed the sequencing data and performed bioinformatic analyses. P.G.-H. participated in the study design, cloned enhancer elements and contributed to data analysis. V.R. performed all sequencing experiments and helped analyze the bioinformatic data. D.S. contributed to the in vivo experiments, performed the ChIP-MiSeq experiments and provided experimental protocols. M.-M.A., F.C.-A., A.H.J. and S.Z. helped with functional experiments. O.M., F.T. and C.L. provided statistical advice and helped in the bioinformatic analyses. G. Perot assisted in generation of the shRNA constructs. M.-C.L.D., O.O. and P.M.-B. provided Ewing sarcoma samples and annotation. G. Pierron, S.R. and E.L. provided and prepared Ewing sarcoma samples. T.R.F. coordinated and supervised sequencing experiments. V. Boeva helped analyze ChIP-Seq data. J.A. provided the A673-TR-shEF1 and SK-N-MC–TR–shEF1 cell lines. A.S.V. performed principal-component analysis clustering of Ewing sarcoma subjects and controls. G.C.-T. and O.C. provided DNA from healthy controls. D.G.C. and S.J.C. provided genetic and statistical guidance. S.B., L.M.M., M.J.M. and S.J.C. provided data for the CCSS replication cohort and performed imputation analyses. P.C. provided advice on analyses concerning EGR2 and laboratory infrastructure. O.D. initiated, designed and supervised the study; provided biological and genetic guidance; analyzed the data; wrote the paper together with T.G.P.G.; and provided laboratory infrastructure and financial support. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Olivier Delattre.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Expression pattern of EGR2 and ADO in normal tissues relative to Ewing sarcoma.

The normal-body atlas consisted of 353 microarrays representing 63 individual tissue types (GSE3526). Gene expression levels are shown as the mean and s.e.m. of n ≥ 3 samples per tissue type. The normal-body atlas and the Ewing sarcoma (EwS) data (GSE34620) were normalized simultaneously by RMA using custom brainarray CDF (v18, ENTREZG). All microarray data were generated on Affymetrix HG-U133Plus2.0 arrays.

Supplementary Figure 2 eQTL analyses across tissue types identify Ewing sarcoma-specific correlations of EGR2 and ADO expression with the risk-allele (G) at rs1848797.

Data are shown as medians (horizontal bars) with ranges for the 25th−75th percentile (box) and 10th−90th percentile (whiskers). Outliers are depicted as dots. P values determined via linear regressions. EwS, Ewing sarcoma; AML, acute myeloid leukemia; LCL, lymphoblastoid cell lines.

Supplementary Figure 3 Analysis of EGR2 and ADO expression in Ewing sarcoma cell lines after knockdown of EWSR1-FLI1.

A673, SK-N-MC, EW7 and POE cells were transfected either with non-targeting control siRNA or siRNA targeting EWSR1-FLI1. Gene expression was assessed 48 h thereafter by qRT-PCR. FC, fold change. Data are shown as the mean and s.e.m.; n ≥ 5 independent experiments. Two-tailed unpaired Student’s t-test; ns, not significant; * P < 0.05, ** P < 0.01, *** P < 0.001.

Supplementary Figure 4 Knockdown of EGR2 reduces clonogenic growth, cell cycle progression, and viability of Ewing sarcoma cells.

(a) Analysis of EGR2 protein expression by immunoblot in Ewing sarcoma cell lines (loading control: β-actin). The neuroblastoma cell line SK-N-SH served as negative control. (b) Analysis of EGR2 protein expression by immunohistochemistry in tumors of xenografted Ewing sarcoma cell lines (A673, TC-71 and SK-ES1), the alveolar rhabdomyosarcoma cell line SJ-RH30, and the neuroblastoma cell line IMR-32, and comparison with EGR2 mRNA expression levels as determined by Affymetrix HG-U133Plus2.0 arrays (GSE36133). Scale bars = 200 µm. (c) qRT-PCR analysis of knockdown efficacy of siRNAs used to silence EGR2 or ADO (48 h after transfection). Data are shown as the mean and s.e.m.; n ≥ 4 independent experiments. (d) Analysis of clonogenic growth after seeding at low-density and serial re-transfection with siRNA every four days. Cells were fixed 9–14 days after seeding and individual colonies were colorized with crystal violet. (e) Analysis of cell cycle phases by PI staining 96 h after transfection with siRNA. EGR2 knockdown reduces the percentage of cells in S phase (P < 0.01 for A673, SK-N-MC, and POE; P = 0.12 for EW7), while increasing the percentage in sub G1 phase (all P < 0.0001). (f) Analysis of apoptosis by Annexin-V-staining 96 h after transfection with siRNA. Data in d-f are shown as the mean and s.e.m. of results obtained with two different siRNAs for EGR2 and three different siRNAs for ADO as displayed in c; n ≥ 3 independent experiments. Two-tailed unpaired Student’s t-test; ns, not significant; *** P < 0.001.

Supplementary Figure 5 EGR2 is a downstream component of the FGF pathway in Ewing sarcoma.

(a) Scatter-dot plot and medians (horizontal bars) of EGFRs (gray) and FGFRs (red) expression levels in n = 117 primary Ewing sarcoma tumors (GSE34620). (b) Analysis of proliferation of Ewing sarcoma cell lines with a Resazurin assay. Cells were seeded in RPMI 1640 medium with 0.5–1% FCS and stimulated with the indicated growth factor (GF) concentrations for 72 h. Data are shown as the mean and s.e.m.; n ≥ 6 experiments. (c) Analysis of EGR2 induction by qRT-PCR in Ewing sarcoma cell lines after incubation with either EGF or bFGF (both 10 ng/ml). The EGFR expressing MDA-MB-231 breast cancer cell line served as a positive control for EGF action. Data are shown as the mean and s.e.m.; n ≥ 3 independent experiments. Two-tailed unpaired Student’s t-test, ** P < 0.01, *** P < 0.001.

Supplementary Figure 6 Targeted germline deep sequencing of principal-component analysis (PCA)-matched Ewing sarcoma cases and controls.

(a) PCA-clustering of the selected core population (dashed box) for sequencing. (b) Analysis of target-region coverage after mapping and base quality filtering of reads (MQ20 and BQ20). (c) Analysis of average nucleotide coverage (high-quality reads) of the chr10 target-region across all samples. The median nucleotide coverage is reported (217X).

Supplementary Figure 7 Schematic illustration of the workflow for next-generation sequencing and variant analysis.

AAR, alternative allele ratio.

Supplementary Figure 8 Genomic coordinates, evolutionary sequence conservation, aligned DNA sequences and homology of EGR2 enhancer elements.

Mammal Cons, 46 vertebrates basewise conservation from the UCSC genome browser60; GERP, Genomic Evolutionary Rate Profiling of 30 vertebrate species from the UCSC genome browser; MultiZ Align, multiple sequence alignment from the UCSC Genome browser; MSE, Myelinating Schwann cell Enhancer; BoneE, Bone Enhancer. Sequence alignment of murine and human DNA sequences was carried out using Clustal Ω (v1.2.0)45. Asterisks indicate homologous nucleotides.

Supplementary Figure 9 Genomic coordinates, epigenetic profile and reference sequence of the mSat1 locus.

GGAA repeats are underlined by arrows. The reported numbers of GGAA motifs correspond to the reference sequence (hg19).

Supplementary Figure 10 Validation of ChIP efficiency by qRT-PCR in the Ewing sarcoma cell line MHH-ES1 (A/T at rs79965208).

A described CCND1 EWSR1-FLI1 binding site was used as positive control47, and an intronic CCND1 locus (intron 2) served as negative control. Data are shown as the mean and s.d.

Supplementary Figure 11 Conditional association analysis suggests that rs79965208 within mSat2 is the major EGR2 regulatory variant at the chr10 susceptibility locus.

Supplementary information

Supplementary Figures

Supplementary Figures 1–11 (PDF 2330 kb)

Supplementary Data

(XLSX 253 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grünewald, T., Bernard, V., Gilardi-Hebenstreit, P. et al. Chimeric EWSR1-FLI1 regulates the Ewing sarcoma susceptibility gene EGR2 via a GGAA microsatellite. Nat Genet 47, 1073–1078 (2015). https://doi.org/10.1038/ng.3363

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3363

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer