Skip to main content
International Journal of Clinical and Experimental Pathology logoLink to International Journal of Clinical and Experimental Pathology
. 2010 Mar 19;3(4):338–347.

Ewing Sarcoma, an enigmatic malignancy of likely progenitor cell origin, driven by transcription factor oncogenic fusions

Paul Jedlicka 1
PMCID: PMC2872742  PMID: 20490326

Abstract

Since its first description by James Ewing in 1921, Ewing Sarcoma has been a cryptic malignancy. A poorly differentiated tumor of uncertain histogenesis and aggressive biologic behavior, it is the second most common malignancy of bone and soft tissue affecting adolescents and young adults. Some two decades ago, the understanding of Ewing Sarcoma biology took a leap forward with the identification of recurrent EWS/Ets fusions, which drive onco-genesis in this disease. A further leap forward occurred over the last half decade with the application of gene silencing, global expression profiling and primary cell culture technologies to the study of this disease. Resulting work has revealed EWS/Ets fusions to be surprisingly versatile regulators of gene expression, and has narrowed the search for the elusive cell of origin. Improved understanding of EWS/Ets biology and relevant oncogenic pathways has in turn led to the development of targeted therapies, including, recently, small molecules targeting key complexes involving the oncogenic fusion itself. In many respects still remaining an enigma, Ewing Sarcoma is an important model for cancers originating in progenitor-type cells or manifesting progenitor-type cell features, and cancers containing recurrent oncogenic fusions, the latter a surprisingly expanding number.

Keywords: Ewing Sarcoma, fusion oncogene, transcription factor, progenitor cell

Introduction

Since its original description by James Ewing in 1921, Ewing Sarcoma remains an enigmatic malignancy. Ewing Sarcoma is a biologically aggressive, poorly differentiated tumor of bone and soft tissue, and less commonly viscera [1-3]. The current definition encompasses the historical entities of Ewing Sarcoma, Peripheral Primitive Neuroectodermal Tumor, Peripheral Neuroepithelioma and Askin Tumor, all of which share the same EWS/Ets oncogenic fusions and similar biologic behavior. Ewing Sarcoma represents the second most common bone and soft tissue malignancy in adolescents and young adults, peaking in incidence at about 5 per million in this age group. It is slightly more common in males, more common in Caucasians and rarely arises in individuals of African ancestry. As with other bone and soft tissue malignancies, Ewing Sarcoma presents with pain and/or a mass, and a destructive/infiltrative lesion on imaging. The most common bony sites are the long bones of the extremities, pelvis, chest wall and spine. Lesions of long bones typically involve the diaphysis.

Biologically, Ewing Sarcoma is a classic example of a malignancy driven by a fusion oncogene [3-8]. Oncogenic fusions in Ewing Sarcoma arise from specific chromosomal translocations that yield in-frame fusion of the amino terminus of the EWS gene on chromosome 22 and the carboxyl terminus, including the DNA-binding domain, of an Ets gene. EWS is a member of the TET family of proteins, which are ubiquitously expressed in all cells and appear to be involved in transcription and/or RNA processing. Ets genes, which tend to manifest more restricted expression, are tissue-specific transcription factors. The Ets gene is Fli1 in 85% of cases, Erg in 10% of cases, and Etv1, Etv4 or FEV in the remaining 5% of cases. In-frame fusion of EWS to an Ets factor in Ewing Sarcoma yields a highly expressed, non-physiologic, potent transcription factor, which activates an oncogenic program in the cells of origin of the tumor. Clinically, Ew-ing Sarcoma is an aggressive malignancy with a relatively poor long-term outcome [3]. Multiagent chemotherapy has greatly improved the survival of Ewing Sarcoma patients, and remains the mainstay of therapy. Unfortunately, overall long-term survival with the best conventional therapy remains around 50%. Moreover, one of four patients present with overt metas-tatic disease, and 75% of these patients do not survive 5 years. Five-year survival of patients with recurrent disease is even worse, at less than 10%.

With advances in technology, namely siRNA-mediated gene silencing, global gene expression profiling, and primary cell isolation, culture and manipulation, understanding of the molecular pathways of pathogenesis in Ewing Sarcoma has greatly advanced in the last decade. This increased understanding has, in turn, led to advances, not yet fully realized but highly promising, in therapy of this aggressive neoplasm. Ewing Sarcoma has been the subject of other excellent recent reviews [3, 5, 6, 8]. The goal of this review is to provide an update of the patho-biology of this disease, with a focus on exciting recent findings related to the mechanisms of regulation of oncogenic pathways by EWS/Ets fusions, histogenesis of this cryptic malignancy, and possibilities for biologically targeted therapy.

Oncogenesis by EWS/Ets fusions I: mechanisms of (dys) regulation of gene expression

Studies to date reveal EWS/Ets fusions to be multifunctional gain-of-function transcriptionai regulators, capable of exerting activating and repressive effects on gene expression, both important to oncogenesis in Ewing Sarcoma (Figure 1). The ability of EWS/Ets fusions to activate gene expression has been evident from the earliest studies of the function of these onco-proteins, which showed that the EWS component confers a more potent transcription activating domain than present in the parent Ets gene [9]. Moreover, other studies demonstrated a correlation between the potency of transcriptionai activation and clinical prognosis for the most commonly observed EWS/Ets fusions [10]. A number of genes have been shown to be directly activated by EWS/Ets fusions in the context of Ewing Sarcoma cells or mesenchymal progenitor cells, the candidate cell of origin, including IGF-1 and the transcriptional coregula-tor NR0B1 [11, 12]. The importance of activation of gene expression to Ewing Sarcoma onco-genesis by EWS/Ets fusions has been shown experimentally: the introduction of KRAB-Fli1, which contains a potent repressor domain in place of the EWS component, into Ewing Sarcoma cells, inhibits both oncogenic transformation in vitro and tumorigenicity in vivo [13]. Interestingly, recent ChlP-chip studies indicate that, in addition to utilizing similar DNA-binding sites to their parent Ets factors, EWS/Ets fusions make additional use of GGAA repeat-containing microsatellites to activate target gene promoters [12, 14]. This mechanism of activation requires, at minimum, four to five such repeats and appears to be specific to EWS/Fli1 upregulated genes [12].

Figure 1.

Figure 1

EWS/Ets gene regulatory pathways. EWS/Fli1 and other EWS/Ets fusions both activate and repress gene expression. EWS/Fli1 activates some target genes directly, through Ets DNA-binding sites (EBS) or through GGAA mi-crosatellites [(GGAA)n]. Some target genes are coreguiated (activated or repressed) with the transcriptionai coregula-tor NR0B1, itself an EWS/Fli1-induced gene. EWS/Fli1 is also able to repress some target genes directly, through Ets DNA-binding sites (EBS). Additional mechanisms of target gene repression are induction of the transcriptionai repres-sor Nkx2.2, and induction of the histone methyltransferase EZH2. EWS/Ets fusions may also regulate target gene expression through non-transcriptional mechanisms, including RNA processing.

The extent and importance of transcriptional repression by EWS/Ets oncogenic fusions have only been appreciated more recently. Indication that EWS/Ets fusions could repress gene expression had existed from studies of some target genes, notably the Transforming Growth Factor beta Receptor II (TGFβRII) [15]. However, the extent of repression of gene expression only became evident from global gene expression profiling studies of Ewing Sarcoma cells with stably silenced EWS/Ets [16, 17]. As evident from studies of TGFPRII, EWS/Ets fusions are capable of transcriptional repression in some contexts. However, gene expression profiling studies have also revealed entirely novel and unexpected additional mechanisms of transcriptional repression operant in Ewing Sarcoma. The homeoprotein Nkx2.2 is one of the most strongly induced genes by EWS/Fli1 in Ewing Sarcoma cells [17]. Functional analyses of Nkx2.2 reveal that it is a transcriptional repressor responsible for down-regulation of a subset of genes by EWS/Ets [18]. Since Nkx2.2 is required for EWS/Ets-mediated transformation [17], this EWS/Ets-Nkx2.2 pathway establishes the biological importance of transcriptional repression in Ewing Sarcoma oncogenesis. The orphan nuclear receptor NR0B1, another highly induced gene by EWS/Fli1, is also required for optimal transformation by EWS/Ets [19], and is a known transcriptional corepressor [20]. Interestingly, ChlP-chip, ChIP and biochemical experiments indicate that EWS/Fli1 and NR0B1 interact on target gene promoters [21]. Thus, NR0B1 is another candidate for target gene repression by EWS/Ets. Analyses of NR0B1 function indicate that it is able to both repress and activate gene expression; moreover, on one studied promoter, EWS/Fli1 appears to inhibit activation by NR0B1 [21]. The EWS/Ets-NROBl pathway thus highlights the likely importance of combinatorial protein-protein interactions and overall promoter context in the end-effect of EWS/Ets fusions on target gene expression. An additional, potentially far-reaching, mechanism of repression of gene expression by EWS/Ets fusions involves a histone methyl transferase [22]. The histone methyltransferase enhancer of Zeste, Drosophila, Homolog 2 (EZH2) silences gene expression by lysine methylation, as part of the Polycomb Repressor Complex 2 (PRC2). EZH2 is an upregulated target, apparently direct, of EWS/Fli1, and is required for optimal oncogenic transformation, tumor growth and metastasis. Interestingly, siRNA silencing and expression profiling experiments indicate that EZH2 inhibits endothelial, epithelial and neural differentiation in Ewing Sarcoma cells. Like the Nkx2.2 studies discussed above, therefore, these studies again underscore the functional importance of repression of gene expression in Ewing Sarcoma pathogenesis.

Early experiments defining the oncogenic potential of EWS/Ets oncogenic fusions in NIH3T3 cells suggested the existence of a DNA-binding-independent transforming activity [23, 24], in addition to the DNA-binding-dependent activity discussed above. This was quite intriguing, given the known functions of the EWS gene in RNA processing [25]. Interestingly, several studies have shown that EWS/Ets fusions can affect RNA splicing [26-29]. Moreover, EWS/Fli1 and EWS interact in the cell [30, 31]. This leaves open the interesting possibility that oncogenic transformation by EWS/Fli1 in part involves mechanisms related to RNA processing. Such a mechanism is clearly not sufficient for oncogenic transformation, as indicated by the inability of a DNA-binding impaired EWS/Fli1 allele to rescue transformation in Ewing Sarcoma cells with stably silenced endogenous EWS/Fli1 or to induce tumors in primary bone derived cells, in contrast to “wild type” EWS/Fli1 [17, 32]. However, such a mechanism may be needed for optimal transformation and/or for other, currently less well studied, parameters of Ewing Sarcoma tumor biology.

Oncogenesis by EWS/Ets fusions II: cell of origin and mechanisms of oncogenic transformation

The cell of origin of Ewing Sarcoma has long been a subject of debate, especially among pathologists! A number of studies have recently converged on mesenchymal progenitor cells as a likely cell of origin. On the one hand, Riggi et al I showed that stable expression of an EWS/Ets oncogenic fusion in mouse mesenchymal progenitor cells was sufficient to give rise to poorly differentiated tumors resembling Ewing Sarcoma [33]. The authors introduced an EWS/Fli1 fusion into a primary culture of mouse mesenchymal progenitor cells (mMPCs) using a retroviral expression vector and obtained Ewing Sarcoma-like tumors upon injection of the manipulated cells into SCID mice. Importantly, the tu-morigenic mMPCs apparently maintained intact p53 and pl6INK4A/pl9ARF function (as seen in most Ewing Sarcomas [34, 35]), thus suggesting that, in the appropriate cell type, EWS/Ets expression may be sufficient to initiate and support tumorigenesis. In a similar study, Castillero-Trejo et al introduced the EWS/Fli1 fusion into primary bone-derived cells and also obtained Ewing Sarcoma-like tumors [32]. The efficiency of tumor formation increased with passage number and, in this study, some tumors acquired p53 mutations. Thus, mouse progenitor-type cells from bone and/or bone marrow are permissive to tumor formation upon expression of EWS/Fli1 (Table 1).

Table 1.

Mouse xenograft models of Ewing Sarcoma using primary, unmodified (non-transformed and non-immortalized) cells

Author Cells/source Growth medium Cell characterization EWS fusion Time of introduction In vitro assays Mice Route time post infection tumors p53/Ink/Arf
Riggi et al, 2005 mMPC (C57BL/6 BM) custom Flow cytometry; differentiation assay Fli1, “t2” ? None SCID sc 5 days 100% at 6 wks Intact
Riggi et al, 2008 hMPC (femoral head BM) IMDM/10% FCS/PDGF Flow cytometry; differentiation assay Fli1, “t2” PD #8 Cell growth:no change “immunocompromised” renal capsule ? None ?
Castillero-Trejo et al, 2005 mPBDC (BALB/c BM) DMEM/10% FCS (Flow cytometry of tumor cells) Fli1, t1 P0 Soft agar (EWS/Fli1-dependent growth) syngeneic sc/ip/iv ? <P15:rare; P17-18:40%; >P28: 100% p53 mutation in some tumors

Abbreviations used: m=mouse; h=human; MPC=mesenchymal progenitor cells; PBDC=primary bone-derived cells; BM=bone marrow; tl=type 1 EWS/Fli1 fusion; t2=type 2 EWS/Fli1 fusion (written in quotation marks because the SK-N-MC cell line, from which this fusion was reported to be derived from, harbors the type 1 fusion); PD=population doubling; P=cell passage number following isolation; sc=subcutaneous; ip=intraperitoneal; iv=intravenous.

Tirade et al, on the other hand, demonstrated that stable silencing of EWS/Fli1 in Ewing Sarcoma cells results in a gene expression signature that overlaps with mesenchymal progenitor cells [36]. Furthermore, the authors demonstrated that the resulting cells are able to differentiate along adipocytic, chondrocytic and osteocytic lineages under appropriate growth conditions, a defining property of mesenchymal progenitor cells. Interestingly, control cells (without EWS/Fli1 silencing) showed a similar ability to differentiate along the same lineages, albeit to a lesser degree. In a similar analysis, Potikyan et al compared the gene expression profile of Ewing Sarcoma cells with silenced EWS/Fli1 to the expression profiles of a panel of diverse cells lines, and obtained the closest match with human fetal lung fibroblasts, thus further supporting a mesenchymal origin [37]. These studies thus suggest that Ewing Sarcoma cells possess mesenchymal progenitor cell-type properties, which are enhanced upon loss of expression of the EWS/Ets oncogenic fusion. This idea is further supported by the observation that expression of EWS/Ets fusions in pluri-potent marrow stromal cells inhibits their ability to differentiate along osteogenic and adipogenic lineages [38]. Interestingly, the mesenchymal progenitor properties of Ewing Sarcoma cells appear to be particularly prominent in CD133+ cancer stem cells isolated from primary tumors, compared to their CD133- counterparts [39].

Several other studies, however, suggest that tumor initiation in Ewing Sarcoma may be more complex than single step transformation of mesenchymal progenitor cells by EWS/Ets fusions. First, expression of EWS/Fli1 in primary human mesenchymal progenitor cells yields a Ewing Sarcoma-like expression signature, but no tumors in a mouse xenograft model [40]. Since this analysis was performed by the same group that did the analogous mouse experiments, the difference in results is not likely to be due to different technique. Rather, it may be due to differences in the precise cells used, species differences, or exact fusion expression levels. Second, in experiments also performed by this same group, expression of similar levels of EWS/Erg, the second most common Ewing Sarcoma oncogenic fusion, is not well tolerated by mouse mesenchymal progenitor cells, in contrast to EWS/Fli1 [11]. Further, in a mouse genetic model, induced expression of EWS/Fli1 in the bone marrow compartment, including mesenchymal progenitor cells, leads to the rapid development of leukemia, but not Ewing Sarcoma-like tumors [41]. In a different mouse genetic model, targeted expression of EWS/Fli1 in developing embryonic limb bud mesenchyme, does give rise to tumors with some resemblance to Ewing Sarcoma, but only in combination with inactivation of p53 [42]. Possible explanations for the discrepancy between the xenograft and genetic models include differences in the precise nature and/or abundance of the isolated/ targeted cells, differences in EWS/Fli1 expression levels, and possible longer latency of Ewing Sarcoma compared to leukemia. Lastly, even the EWS/Fli1-driven mouse mesenchymal progenitor cell-derived tumors showed deviations from Ewing Sarcoma [33]. Namely, at the molecular level, the mMPC-EWS/Fli1 tumors strongly activated the expression of a number of genes of potential pathogenic importance including IGF-1, Myc, Id2 and MMP3. However, of these genes, only IGF-1 is identified as an EWS/ Fli1-regulated target in fusion silencing systems in Ewing Sarcoma cell lines [16, 17, 43]. Thus, while all of the above studies represent progress toward an understanding of Ewing Sarcoma histogenesis, the issue of the precise human cell of origin and the question of a requirement for second “hits” remain unresolved.

Oncogenesis by EWS/Ets fusions III: EWS/Ets-regulated oncogenic pathways and targeted therapy

Given the pivotal importance of EWS/Ets fusions in Ewing Sarcoma pathogenesis, these oncogenes have been referred to as the “perfect target without a therapeutic target” [44]. One possible way to target EWS/Ets fusions is inhibition of expression. Such as approach, using small interfering RNA (siRNA) technology, has been successful in inhibiting the growth of Ewing Sarcoma cells in a mouse xenograft metastatic tumor model [45]. Another possible approach is targeting of the fusion proteins themselves. However, this approach has been hampered by difficulties with biochemical purification and analysis of the fusion proteins, due to the presence of unstructured regions and poor solubility [44]. A related approach, and one that has borne the most fruit thus far, is targeting of critical protein-protein interactions of EWS/Ets fusions.

A number of different studies have identified protein-protein interaction partners for EWS/Ets oncogenic fusions. EWS/Fli1 interacts with the transcription factor CBP, and a dominant-negative form of CBP sensitizes Ewing Sarcoma cells to apoptosis [46]. Promoter interactions with the AP-1 (fos/jun) transcription complex may be important for transcriptional regulation by EWS/Ets fusions, as a C-terminal deletion mutant of EWS/Fli1, defective in cooperative interactions with AP-1, is impaired in cellular transformation [47]. As discussed above, EWS/ Fli1 forms a complex with the transcriptional co-regulator NR0B1 [21], which is itself a transcriptional target of EWS/Fli1 and is essential for EWS/Fli1-mediated oncogenic transformation [19]. Similarly, EWS/Fli1 forms complexes with RNA Helicase A, and this interaction stimulates EWS/Fli1 promoter activation and cellular transformation [48]. Additional protein-protein interactions of EWS/Fli1, currently with unknown functional significance in Ewing Sarcoma, include the POU homeodomain protein Brn-3a [49], the BRCA-interacting protein BARD1 [50], hyperphosphorylated RNA Polymerase II (Pol II) [29], RNA Pol II subunit hsRPB7 [51], and serum response factor (SRF) [52]. The interaction between EWS/Fli1 and RNA Helicase A (RHA) is the first to have been successfully “drugged” thus far. Using surface plasmon resonance screening and recombinant EWS/Fli1, Erkizan et al identified a small molecule able to interfere with the EWS/Fli1-RHA interaction, induce apoptosis of Ewing Sarcoma cells and inhibit tumor growth in a mouse xenograft model [53].

EWS/Ets fusions also regulate growth factor pathways (Figure 2), which may lend themselves as potential drug targets, either alone or in combination with inhibitors of EWS/Ets activity. The Insulin-like Growth Factor 1 (IGF-1) signaling pathway is critical to Ewing Sarcoma on-cogenesis, via autocrine, and likely endocrine, pathways [54-56]. EWS/Fli1 activates IGF-1 expression [11, 43], while down regulating the expression of the IGF-binding protein IGFBP3 [16], thus presumably stimulating the autocrine pathway. Blockade of the IGF-1 receptor (IGF-IR) is inhibitory to Ewing Sarcoma growth in vitro and in in vivo mouse xenograft models [55-59], and clinical trials involving both small molecule and antibody inhibitors of IGF-IR in Ewing Sarcoma are ongoing (reviewed in [3]). Additional inhibition of pathway components downstream of the growth factor/ receptor interaction, including MAPK/MEK, PI3-Kand mTOR [60-63], may have added efficacy to IGF-IR blockade. Interestingly, by an undefined mechanism, the mTOR inhibitor Rapamycin reduces cellular levels of EWS/Fli1 [64], and may thus have “double efficacy” in Ewing Sarcoma therapy [65]. EWS/Ets fusions also regulate Wnt signaling in a complex manner. EWS/Ets fusions regulate members of the Dickkopf family, secreted modulators of Wnt signaling. Specifically, they induce DKK2, apparently directly, while repressing DKK1 [36, 66, 67]. DKK1 is suppressive to Ewing Sarcoma tumor growth, while DKK2 may be somewhat tumor promoting [66]. Interestingly, EWS/Fli1 also generally antagonizes the activation of β-catenin responsive genes, possibly by sequestration of TCF/Lef [67]. Additionally, Platelet-Derived Growth Factor C (PDGF-C) is an EWS/Fli1-upregulated gene in NIH3T3 cells ([68]; but, interestingly, not in an EWS/Fli1 silencing/rescue model [17]), and a dominant negative form of PDGF-C inhibits Ewing Sarcoma cell growth [69]. EWS/Fli1 also upregu-lates the effector of Hedgehog pathway signaling Glil, which is required for optimal cellular transformation [70]. Finally, restoration of TGFpRII expression levels, normally down-regulated by EWS/Ets, is inhibitory to Ewing Sarcoma tumor growth [71].

Figure 2.

Figure 2

Growth factor pathways regulated by EWS/Ets oncogenic fusions with demonstrated functions in Ewing Sarcoma oncogenesis. Pro-oncogenic (green) and tumor suppressive (red) targets of EWS/Ets fusions in growth factor pathways. Pro-oncogenic targets include PDGF-C, the mediator of Hedgehog-Smoothened signaling Gli1, IGF-1 and DKK2, a regulator of Wnt signaling. mTOR, downstream of IGF-1 signaling, may positively regulate EWS/Ets fusion levels. EWS/Fli1 antagonizes nuclear (3-catenin transcriptional activity, possibly by sequestering Lef. EWS/Ets-regulated tumor suppressive targets with demonstrated function in Ewing Sarcoma include TGFβRII and DKK1. The expression of DKK1 and DKK2 appears to be mutually antagonistic.

Other known oncogenesis-modifying pathways, currently not known to be regulated by EWS/Ets fusions, are also operant in Ewing Sarcoma, and may be amenable to drug targeting. These include Notch signaling [72] and the NF-kB pathway [70, 73]. Lastly, the identification of presumptive cooperating pathways in EWS/Ets-mediated cellular transformation may present additional opportunities for targeted therapy in Ewing Sarcoma.

Conclusions and perspectives

Over the past two decades, much has been learned about the biology of Ewing Sarcoma, including the pivotal role of EWS/Ets oncogenic fusions and the pathways they regulate, the importance of cellular context for their action and the likely cell of origin, and mechanisms of regulation of downstream gene expression. This information has opened the door to targeted therapies for this aggressive malignancy, including blockade of IGF signaling, and more recently, interference with key EWS/Ets protein-protein interactions. Information learned from the study of Ewing Sarcoma biology has been, and will continue to be, of relevance to other cancers, given the prevalence of activated IGF signaling in other malignancies and the surprising occurrence, in growing numbers of tumors, of oncogenic fusions between EWS family members and transcription factors. Similarly, application of growing knowledge of other highly conserved pathways in cancer is likely to suggest additional ways to attack Ewing Sarcoma therapeutically. Perhaps in some of our lifetimes, Ewing Sarcoma will become a historical curiosity rather than the taker of the lives of children and young adults in their prime.

Acknowledgments

This study was supported by University of Colorado School of Medicine and Cancer Center.

References

  • 1.de Alava E, Gerald WL. Molecular biology of the Ewing's sarcoma/primitive neuroectodermal tumor family. J Clin Oncol. 2000;18(1):204–213. doi: 10.1200/JCO.2000.18.1.204. [DOI] [PubMed] [Google Scholar]
  • 2.Li S, Siegal GP. Small cell tumors of bone. Adv AnatPathol. 17(1):1–11. doi: 10.1097/PAP.0b013e3181bb6b9c. [DOI] [PubMed] [Google Scholar]
  • 3.Ludwig JA. Ewing sarcoma: historical perspectives, current state-of-the-art, and opportunities for targeted therapy in the future. Curr Opin Oncol. 2008;20(4):412–418. doi: 10.1097/CCO.0b013e328303ba1d. [DOI] [PubMed] [Google Scholar]
  • 4.Arvand A, Denny CT. Biology of EWS/ETS fusions in Ewing's family tumors. Oncogene. 2001;20(40):5747–5754. doi: 10.1038/sj.onc.1204598. [DOI] [PubMed] [Google Scholar]
  • 5.Janknecht R. EWS-ETS oncoproteins: the linchpins of Ewing tumors. Gene. 2005;363:1–14. doi: 10.1016/j.gene.2005.08.007. [DOI] [PubMed] [Google Scholar]
  • 6.Kovar H. Context matters: the hen or egg problem in Ewing's sarcoma. Semin Cancer Biol. 2005;15(3):189–196. doi: 10.1016/j.semcancer.2005.01.004. [DOI] [PubMed] [Google Scholar]
  • 7.Ladanyi M. EWS-FLI1 and Ewing's sarcoma: recent molecular data and new insights. Cancer Biol Ther. 2002;1(4):330–336. [PubMed] [Google Scholar]
  • 8.Riggi N, Stamenkovic I. The Biology of Ewing sarcoma. Cancer Lett 2007. 254(1):1–10. doi: 10.1016/j.canlet.2006.12.009. [DOI] [PubMed] [Google Scholar]
  • 9.May WA, Lessnick SL, Braun BS, Klemsz M, Lewis BC, Lunsford LB, Hromas R, Denny CT. The Ewing's sarcoma EWS/FLI-1 fusion gene encodes a more potent transcriptional activator and is a more powerful transforming gene than FLI-1. Mol Cell Biol. 1993;13(12):7393–7398. doi: 10.1128/mcb.13.12.7393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Lin PP, Brody RI, Hamelin AC, Bradner JE, Healey JH, Ladanyi M. Differential transactivation by alternative EWS-FLI1 fusion proteins correlates with clinical heterogeneity in Ewing's sarcoma. Cancer Res. 1999;59(7):1428–1432. [PubMed] [Google Scholar]
  • 11.Cironi L, Riggi N, Provero P, Wolf N, Suva ML, Suva D, Kindler V, Stamenkovic I. IGF1 is a com mon target gene of Ewing's sarcoma fusion pro teins in mesenchymal progenitor cells. PLoS ONE. 2008;3(7):e2634. doi: 10.1371/journal.pone.0002634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Gangwal K, Sankar S, Hollenhorst PC, Kinsey M, Haroldsen SC, Shah AA, Boucher KM, Watkins WS, Jorde LB, et al. Microsatellites as EWS/FLI response elements in Ewing's sarcoma. Proc Natl Acad Sci USA. 2008;105(29):10149–10154. doi: 10.1073/pnas.0801073105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Chan D, Wilson TJ, Xu D, Cowdery HE, Sanij E, Hertzog PJ, Kola I. Transformation induced by Ewing's sarcoma associated EWS/FLI-1 is suppressed by KRAB/FLI-1. Br J Cancer. 2003;88(1):137–145. doi: 10.1038/sj.bjc.6600669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Guillon N, Tirade F, Boeva V, Zynovyev A, Barillot E, Delattre O. The oncogenic EWS-FLI1 protein binds in vivo GGAA microsatellite sequences with potential transcriptional activation function. PLoS One. 2009;4(3):e4932. doi: 10.1371/journal.pone.0004932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Im YH, Kim HT, Lee C, Poulin D, Welford S, Sorensen PH, Denny CT, Kim SJ. EWS-FLI1, EWS-ERG, and EWS-ETV1 oncoproteins of Ewing tumor family all suppress transcription of transforming growth factor beta type II receptor gene. Cancer Res. 2000;60(6):1536–1540. [PubMed] [Google Scholar]
  • 16.Prieur A, Tirade F, Cohen P, Delattre O. EWS/FLI-1 silencing and gene profiling of Ewing cells reveal downstream oncogenic pathways and a crucial role for repression of insulin-like growth factor binding protein 3. Mol Cell Biol. 2004;24(16):7275–7283. doi: 10.1128/MCB.24.16.7275-7283.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Smith R, Owen LA, Trem DJ, Wong JS, Whangbo JS, Golub TR, Lessnick SL. Expression profiling of EWS/FLI identifies NKX2.2 as a critical target gene in Ewing's sarcoma. Cancer Cell. 2006;9(5):405–416. doi: 10.1016/j.ccr.2006.04.004. [DOI] [PubMed] [Google Scholar]
  • 18.Owen LA, Kowalewski AA, Lessnick SL. EWS/FLI mediates transcriptional repression via NKX2.2 during oncogenic transformation in Ewing's sarcoma. PLoS ONE. 2008;3(4):e1965. doi: 10.1371/journal.pone.0001965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kinsey M, Smith R, Lessnick SL. NROB1 is required for the oncogenic phenotype mediated by EWS/FLI in Ewing's sarcoma. Mol Cancer Res. 2006;4(11):851–859. doi: 10.1158/1541-7786.MCR-06-0090. [DOI] [PubMed] [Google Scholar]
  • 20.Iyer AK, McCabe ER. Molecular mechanisms of DAX1 action. Mol Genet Metab. 2004;83(1-2):60–73. doi: 10.1016/j.ymgme.2004.07.018. [DOI] [PubMed] [Google Scholar]
  • 21.Kinsey M, Smith R, Iyer AK, McCabe ER, Lessnick SL. EWS/FLI and its downstream target NROB1 interact directly to modulate transcription and oncogenesis in Ewing's sarcoma. Cancer Res. 2009;69(23):9047–9055. doi: 10.1158/0008-5472.CAN-09-1540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Richter GH, Plehm S, Fasan A, Rossler S, Unland R, Bennani-Baiti IM, Hotfilder M, Lowel D, von Luettichau I, et al. EZH2 is a mediator of EWS/FLI1 driven tumor growth and metastasis blocking endothelial and neuroectodermal differentiation. Proc Natl Acad Sci U SA. 2009;106(13):5324–5329. doi: 10.1073/pnas.0810759106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Jaishankar S, Zhang J, Roussel MF, Baker SJ. Transforming activity of EWS/FLI is not strictly dependent upon DNA-binding activity. Oncogene. 1999;18(40):5592–5597. doi: 10.1038/sj.onc.1202940. [DOI] [PubMed] [Google Scholar]
  • 24.Welford SM, Hebert SP, Deneen B, Arvand A, Denny CT. DNA binding domain-independent pathways are involved in EWS/Fli1-mediated oncogenesis. J Biol Chem. 2001;276(45):41977–41984. doi: 10.1074/jbc.M106757200. [DOI] [PubMed] [Google Scholar]
  • 25.Law WJ, Cann KL, Hicks GG. TLS, EWS and TAF15: a model for transcriptional integration of gene expression. Brief Funct Genomic Proteomic. 2006;5(1):8–14. doi: 10.1093/bfgp/ell015. [DOI] [PubMed] [Google Scholar]
  • 26.Chansky HA, Hu M, Hickstein DD, Yang L. Oncogenic TLS/ERG and EWS/Fli-1 fusion proteins inhibit RNA splicing mediated by YB-1 protein. Cancer Res. 2001;61(9):3586–3590. [PubMed] [Google Scholar]
  • 27.Knoop LL, Baker SJ. The splicing factor U1C represses EWS/FLI-mediated transactivation. J Biol Chem. 2000;275(32):24865–24871. doi: 10.1074/jbc.M001661200. [DOI] [PubMed] [Google Scholar]
  • 28.Knoop LL, Baker SJ. EWS/FLI alters 5'-splice site selection. J Biol Chem. 2001;276(25):22317–22322. doi: 10.1074/jbc.M008950200. [DOI] [PubMed] [Google Scholar]
  • 29.Yang L, Chansky HA, Hickstein DD. EWS.FIi-1 fusion protein interacts with hyperphosphory-lated RNA polymerase II and interferes with serine-arginine protein-mediated RNA splicing. J Biol Chem. 2000;275(48):37612–37618. doi: 10.1074/jbc.M005739200. [DOI] [PubMed] [Google Scholar]
  • 30.Embree U, Azuma M, Hickstein DD. Ewing sarcoma fusion protein EWSR1/FLI1 interacts with EWSR1 leading to mitotic defects in zebrafish embryos and human cell lines. Cancer Res. 2009;69(10):4363–4371. doi: 10.1158/0008-5472.CAN-08-3229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Spahn L, Siligan C, Bachmaier R, Schmid JA, Aryee DN, Kovar H. Homotypic and heterotypic interactions of EWS, FLI1 and their oncogenic fusion protein. Oncogene. 2003;22(44):6819–6829. doi: 10.1038/sj.onc.1206810. [DOI] [PubMed] [Google Scholar]
  • 32.Castillero-Trejo Y, Eliazer S, Xiang L, Richardson JA, Maria RL., Jr Expression of the EWS/FLI-1 oncogene in murine primary bone-derived cells Results in EWS/FLI-1-dependent, ewing sarcoma -like tumors. Cancer Res. 2005;65(19):8698–8705. doi: 10.1158/0008-5472.CAN-05-1704. [DOI] [PubMed] [Google Scholar]
  • 33.Riggi N, Cironi L, Provero P, Suva ML, Kaloulis K, Garcia-Echeverria C, Hoffmann F, Trumpp A, Stamenkovic I. Development of Ewing's sarcoma from primary bone marrow-derived mesenchymal progenitor cells. Cancer Res. 2005;65(24):11459–11468. doi: 10.1158/0008-5472.CAN-05-1696. [DOI] [PubMed] [Google Scholar]
  • 34.Lopez-Guerrero JA, Pellin A, Noguera R, Carda C, Llombart-Bosch A. Molecular analysis of the 9p21 locus and p53 genes in Ewing family tumors. Lab Invest. 2001;81(6):803–814. doi: 10.1038/labinvest.3780290. [DOI] [PubMed] [Google Scholar]
  • 35.Tsuchiya T, Sekine K, Hinohara S, Namiki T, Nobori T, Kaneko Y. Analysis of the pl6INK4, pl4ARF, pl5, TP53, and MDM2 genes and their prognostic implications in osteosarcoma and Ewing sarcoma. Cancer Genet Cytogenet. 2000;120(2):91–98. doi: 10.1016/s0165-4608(99)00255-1. [DOI] [PubMed] [Google Scholar]
  • 36.Tirode F, Laud-Duval K, Prieur A, Delorme B, Charbord P, Delattre O. Mesenchymal stem cell features of Ewing tumors. Cancer Cell. 2007;11(5):421–429. doi: 10.1016/j.ccr.2007.02.027. [DOI] [PubMed] [Google Scholar]
  • 37.Potikyan G, France KA, Carlson MR, Dong J, Nelson SF, Denny CT. Genetically defined EWS/FLI1 model system suggests mesenchymal origin of Ewing's family tumors. Lab Invest. 2008;88(12):1291–1302. doi: 10.1038/labinvest.2008.99. [DOI] [PubMed] [Google Scholar]
  • 38.Torchia EC, Jaishankar S, Baker SJ. Ewing tumor fusion proteins block the differentiation of pluri-potent marrow stromal cells. Cancer Res. 2003;63(13):3464–3468. [PubMed] [Google Scholar]
  • 39.Suva ML, Riggi N, Stehle JC, Baumer K, Tercier S, Joseph JM, Suva D, Clement V, Provero P, et al. Identification of cancer stem cells in Ewing's sarcoma. Cancer Res. 2009;69(5):1776–1781. doi: 10.1158/0008-5472.CAN-08-2242. [DOI] [PubMed] [Google Scholar]
  • 40.Riggi N, Suva ML, Suva D, Cironi L, Provero P, Tercier S, Joseph JM, Stehle JC, Baumer K, et al. EWS-FLI-1 expression triggers a Ewing's sarcoma initiation program in primary human mesenchymal stem cells. Cancer Res. 2008;68(7):2176–2185. doi: 10.1158/0008-5472.CAN-07-1761. [DOI] [PubMed] [Google Scholar]
  • 41.Torchia EC, Boyd K, Rehg JE, Qu C, Baker SJ. EWS/FLI-1 induces rapid onset of myeloid/erythroid leukemia in mice. Mol Cell Biol. 2007;27(22):7918–7934. doi: 10.1128/MCB.00099-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Lin PP, Pandey MK, Jin F, Xiong S, Deavers M, Parant JM, Lozano G. EWS-FLI1 induces developmental abnormalities and accelerates sarcoma formation in a transgenic mouse model. Cancer Res. 2008;68(21):8968–8975. doi: 10.1158/0008-5472.CAN-08-0573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Herrero-Martin D, Osuna D, Ordonez JL, Sevillano V, Martins AS, Mackintosh C, Campos M, Madoz-Gurpide J, Otero-Motta AP, et al. Stable interference of EWS-FLI1 in an Ewing sarcoma cell line impairs IGF-1/IGF-1R signalling and reveals TOPK as a new target. Br J Cancer. 2009;101(1):80–90. doi: 10.1038/sj.bjc.6605104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Uren A, Toretsky JA. Ewing's sarcoma oncoprotein EWS-FLI1: the perfect target without a therapeutic agent. Future Oncol. 2005;1(4):521–528. doi: 10.2217/14796694.1.4.521. [DOI] [PubMed] [Google Scholar]
  • 45.Hu-Lieskovan S, Heidel JD, Bartlett DW, Davis ME, Triche TJ. Sequence-specific knockdown of EWS-FLI1 by targeted, nonviral delivery of small interfering RNA inhibits tumor growth in a murine model of metastatic Ewing's sarcoma. Cancer Res. 2005;65(19):8984–8992. doi: 10.1158/0008-5472.CAN-05-0565. [DOI] [PubMed] [Google Scholar]
  • 46.Ramakrishnan R, Fujimura Y, Zou JP, Liu F, Lee L, Rao VN, Reddy ES. Role of protein-protein interactions in the antiapoptotic function of EWS-Fli-1. Oncogene. 2004;23(42):7087–7094. doi: 10.1038/sj.onc.1207927. [DOI] [PubMed] [Google Scholar]
  • 47.Kim S, Denny CT, Wisdom R. Cooperative DNA binding with AP-1 proteins is required for transformation by EWS-Ets fusion proteins. Mol Cell Biol. 2006;26(7):2467–2478. doi: 10.1128/MCB.26.7.2467-2478.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Toretsky JA, Erkizan V, Levenson A, Abaan OD, Parvin JD, Cripe TP, Rice AM, Lee SB, Uren A. Oncoprotein EWS-FLI1 activity is enhanced by RNA helicase A. Cancer Res. 2006;66(11):5574–5581. doi: 10.1158/0008-5472.CAN-05-3293. [DOI] [PubMed] [Google Scholar]
  • 49.Gascoyne DM, Thomas GR, Latchman DS. The effects of Brn-3a on neuronal differentiation and apoptosis are differentially modulated by EWS and its oncogenic derivative EWS/Fli-1. Oncogene. 2004;23(21):3830–3840. doi: 10.1038/sj.onc.1207497. [DOI] [PubMed] [Google Scholar]
  • 50.Spahn L, Petermann R, Siligan C, Schmid JA, Aryee DN, Kovar H. Interaction of the EWS NH2 terminus with BARD1 links the Ewing's sarcoma gene to a common tumor suppressor pathway. Cancer Res. 2002;62(16):4583–4587. [PubMed] [Google Scholar]
  • 51.Petermann R, Mossier BM, Aryee DN, Khazak V, Golemis EA, Kovar H. Oncogenic EWS-Fli1 interacts with hsRPB7, a subunit of human RNA poly-merase II. Onoogene. 1998;17(5):603–610. doi: 10.1038/sj.onc.1201964. [DOI] [PubMed] [Google Scholar]
  • 52.Watson DK, Robinson L, Hodge DR, Kola I, Papas TS, Seth A. FLU and EWS-FLI1 function as ternary complex factors and ELK1 and SAP1a function as ternary and quaternary complex factors on the Egrl promoter serum response elements. Oncogene. 1997;14(2):213–221. doi: 10.1038/sj.onc.1200839. [DOI] [PubMed] [Google Scholar]
  • 53.Erkizan HV, Kong Y, Merchant M, Schlottmann S, Barber-Rotenberg JS, Yuan L, Abaan OD, Chou TH, Dakshanamurthy S, et al. A small molecule blocking oncogenic protein EWS-FLI1 interaction with RNA helicase A inhibits growth of Ewing's sarcoma. Nat Med. 2009;15(7):750–756. doi: 10.1038/nm.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kim SY, Toretsky JA, Scher D, Helman U. The role of IGF-1R in pediatric malignancies. Oncologist. 2009;14(1):83–91. doi: 10.1634/theoncologist.2008-0189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Scotlandi K, Benini S, Sarti M, Serra M, Lollini PL, Maurici D, Picci P, Manara MC, Baldini N. Insulin-like growth factor I receptor-mediated circuit in Ewing's sarcoma/peripheral neuroecto-dermal tumor: a possible therapeutic target. Cancer Res. 1996;56(20):4570–4574. [PubMed] [Google Scholar]
  • 56.Yee D, Favoni RE, Lebovic GS, Lombana F, Powell DR, Reynolds CP, Rosen N. Insulin-like growth factor I expression by tumors of neuroectoder-mal origin with the t(11;22) chromosomal translocation. A potential autocrine growth factor. J Clin Invest. 1990;86(6):1806–1814. doi: 10.1172/JCI114910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Manara MC, Landuzzi L, Nanni P, Nicoletti G, Zambelli D, Lollini PL, Nanni C, Hofmann F, Gar-cia-Echeverria C, et al. Preclinical in vivo study of new insulin-like growth factor-I receptor–specific inhibitor in Ewing's sarcoma. Clin Cancer Res. 2007;13(4):1322–1330. doi: 10.1158/1078-0432.CCR-06-1518. [DOI] [PubMed] [Google Scholar]
  • 58.Martins AS, Mackintosh C, Martin DH, Campos M, Hernandez T, Ordonez JL, de Alava E. Insulin-like growth factor I receptor pathway inhibition by ADW742, alone or in combination with imatinib, doxorubicin, or vincristine, is a novel therapeutic approach in Ewing tumor. Clin Cancer Res. 2006;12(11):3532–3540. doi: 10.1158/1078-0432.CCR-05-1778. Pt 1. [DOI] [PubMed] [Google Scholar]
  • 59.Scotlandi K, Manara MC, Nicoletti G, Lollini PL, Lukas S, Benini S, Croci S, Perdichizzi S, Zambelli D, et al. Antitumor activity of the insulin-like growth factor-I receptor kinase inhibitor NVP-AEW541 in musculoskeletal tumors. Cancer Res. 2005;65(9):3868–3876. doi: 10.1158/0008-5472.CAN-04-3192. [DOI] [PubMed] [Google Scholar]
  • 60.Benini S, Manara MC, Cerisano V, Perdichizzi S, Strammiello R, Serra M, Picci P, Scotlandi K. Contribution of MEK/MAPK and PI3-K signaling pathway to the malignant behavior of Ewing's sarcoma cells: therapeutic prospects. Int J Cancer. 2004;108(3):358–366. doi: 10.1002/ijc.11576. [DOI] [PubMed] [Google Scholar]
  • 61.Silvany RE, Eliazer S, Wolff NC, Maria RL., Jr Interference with the constitutive activation of ERK1 and ERK2 impairs EWS/FLI-1-dependent transformation. Oncogene. 2000;19(39):4523–4530. doi: 10.1038/sj.onc.1203811. [DOI] [PubMed] [Google Scholar]
  • 62.Toretsky JA, Thakar M, Eskenazi AE, Frantz CN. Phosphoinositide 3-hydroxide kinase blockade enhances apoptosis in the Ewing's sarcoma family of tumors. Cancer Res. 1999;59(22):5745–5750. [PubMed] [Google Scholar]
  • 63.Zenali MJ, Zhang PL, Bendel AE, Brown RE. Morphoproteomic confirmation of constitutively activated mTOR, ERK, and NF-kappaB pathways in Ewing family of tumors. Ann Clin Lab Sci. 2009;39(2):160–166. [PubMed] [Google Scholar]
  • 64.Mateo-Lozano S, Tirado OM, Notario V. Rapamycin induces the fusion-type independent down-regulation of the EWS/FLI-1 proteins and inhibits Ewing's sarcoma cell proliferation. Oncogene. 2003;22(58):9282–9287. doi: 10.1038/sj.onc.1207081. [DOI] [PubMed] [Google Scholar]
  • 65.Mateo-Lozano S, Gokhale PC, Soldatenkov VA, Dritschilo A, Tirado OM, Notario V. Combined transcriptional and translational targeting of EWS/FLI-1 in Ewing's sarcoma. Clin Cancer Res. 2006;12(22):6781–6790. doi: 10.1158/1078-0432.CCR-06-0609. [DOI] [PubMed] [Google Scholar]
  • 66.Miyagawa Y, Okita H, Itagaki M, Toyoda M, Katagiri YU, Fujimoto J, Hata J, Umezawa A, Kiyo-kawa N. EWS/ETS regulates the expression of the Dickkopf family in Ewing family tumor cells. PLoS One. 2009;4(2):e4634. doi: 10.1371/journal.pone.0004634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Navarro D, Agra N, Pestana A, Alonso J, Gonzalez-Sancho JM. The EWS/FLI1 oncogenic protein inhibits expression of the Wnt inhibitor DICK-KOPF-1 gene and antagonizes ﹛beta﹜-catenin/ TCF-mediated transcription. Carcinogenesis. 2009 doi: 10.1093/carcin/bgp317. [DOI] [PubMed] [Google Scholar]
  • 68.Zwerner JP, May WA. PDGF-C is an EWS/FLI induced transforming growth factor in Ewing family tumors. Oncogene. 2001;20(5):626–633. doi: 10.1038/sj.onc.1204133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Zwerner JP, May WA. Dominant negative PDGF-C inhibits growth of Ewing family tumor cell lines. Oncogene. 2002;21(24):3847–3854. doi: 10.1038/sj.onc.1205486. [DOI] [PubMed] [Google Scholar]
  • 70.Zwerner JP, Joo J, Warner KL, Christensen L, Hu-Lieskovan S, Triche TJ, May WA. The EWS/FLI1 oncogenic transcription factor deregulates GUI. Oncogene. 2008;27(23):3282–3291. doi: 10.1038/sj.onc.1210991. [DOI] [PubMed] [Google Scholar]
  • 71.Hahm KB, Cho K, Lee C, Im YH, Chang J, Choi SG, Sorensen PH, Thiele CJ, Kim SJ. Repression of the gene encoding the TGF-beta type II receptor is a major target of the EWS-FLI1 oncoprotein. Nat Genet. 1999;23(2):222–227. doi: 10.1038/13854. [DOI] [PubMed] [Google Scholar]
  • 72.Ban J, Bennani-Baiti IM, Kauer M, Schaefer KL, Poremba C, Jug G, Schwentner R, Smrzka O, Muehlbacher K, et al. EWS-FLI1 suppresses NOTCH-activated p53 in Ewing's sarcoma. Cancer Res. 2008;68(17):7100–7109. doi: 10.1158/0008-5472.CAN-07-6145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Javelaud D, Poupon MF, Wietzerbin J, Besancon F. Inhibition of constitutive NF-kappa B activity suppresses tumorigenicity of Ewing sarcoma EW7 cells. IntJ Cancer. 2002;98(2):193–198. doi: 10.1002/ijc.10192. [DOI] [PubMed] [Google Scholar]

Articles from International Journal of Clinical and Experimental Pathology are provided here courtesy of e-Century Publishing Corporation

RESOURCES