Next Article in Journal
TIM-3 Is a Potential Immune Checkpoint Target in Cats with Mammary Carcinoma
Next Article in Special Issue
TRAF6 Promotes PRMT5 Activity in a Ubiquitination-Dependent Manner
Previous Article in Journal
Role of Etiology in Hepatocellular Carcinoma Patients Treated with Lenvatinib: A Counterfactual Event-Based Mediation Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of EWSR1-FLI1 Function by Post-Transcriptional and Post-Translational Modifications

1
Lineberger Comprehensive Cancer Center, The University of North Carolina at Chapel Hill, Chapel Hill, NC 27599, USA
2
Department of Biochemistry and Biophysics, The University of North Carolina at Chapel Hill, Chapel Hill, NC 27599, USA
3
Department of Genetics, The University of North Carolina at Chapel Hill, Chapel Hill, NC 27599, USA
4
Department of Pediatrics, The University of North Carolina at Chapel Hill, Chapel Hill, NC 27599, USA
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(2), 382; https://doi.org/10.3390/cancers15020382
Submission received: 5 December 2022 / Revised: 4 January 2023 / Accepted: 4 January 2023 / Published: 6 January 2023
(This article belongs to the Special Issue Protein Regulatory Mechanisms in Tumorigenesis)

Abstract

:

Simple Summary

Ewing sarcoma is a malignant pediatric bone cancer currently lacking targeted therapy. In the US there are ~200 patients diagnosed each year and relapse is associated with resistance to the standard-of-care chemotherapy. Thus, it remains an urgent unmet medical need to develop effective new cures for Ewing sarcoma. It is well-characterized that Ewing sarcoma is largely driven by unique gene fusions, with EWSR1-FLI1 being the most prevalent. In this review, we summarize up-to-date regulatory mechanisms for the onco-fusion protein EWSR1-FLI1 in Ewing sarcoma, including both post-transcriptional and post-translational modifications, to reveal knowledge gaps and propose potential new therapeutic directions.

Abstract

Ewing sarcoma is the second most common bone tumor in childhood and adolescence. Currently, first-line therapy includes multidrug chemotherapy with surgery and/or radiation. Although most patients initially respond to chemotherapy, recurrent tumors become treatment refractory. Pathologically, Ewing sarcoma consists of small round basophilic cells with prominent nuclei marked by expression of surface protein CD99. Genetically, Ewing sarcoma is driven by a fusion oncoprotein that results from one of a small number of chromosomal translocations composed of a FET gene and a gene encoding an ETS family transcription factor, with ~85% of tumors expressing the EWSR1::FLI1 fusion. EWSR1::FLI1 regulates transcription, splicing, genome instability and other cellular functions. Although a tumor-specific target, EWSR1::FLI1-targeted therapy has yet to be developed, largely due to insufficient understanding of EWSR1::FLI1 upstream and downstream signaling, and the challenges in targeting transcription factors with small molecules. In this review, we summarize the contemporary molecular understanding of Ewing sarcoma, and the post-transcriptional and post-translational regulatory mechanisms that control EWSR1::FLI1 function.

1. Introduction

Ewing sarcoma, first described in 1921 by James Ewing, encompasses three tumor types: “classic” Ewing sarcoma of bone, malignant small cell tumors of the chest wall (Askin’s tumor), and primitive neuroectodermal tumors of soft tissue origin (PNET) [1]. Although the definitive cell of origin remains unknown, Ewing sarcoma is thought to originate from mesenchymal progenitor cells in bone and soft tissue [2]. Ewing sarcoma, the second most common pediatric bone malignancy, constitutes 10% to 15% of all bone sarcomas and occurs most commonly in children and young adults [3]. The median age of diagnosis of this disease is 14–15 years [4,5]. Ewing sarcoma mostly occurs in the pelvis, mid-shaft bones and femur; however, it can occur in all bones and many soft tissues. Ewing sarcoma is considered as a systemic disease ab initio with the lung being the most common site of metastasis site. Usually, pain and swelling are observed at disease sites. Over the past century, our understanding of the development, progression and treatment of Ewing sarcoma has advanced greatly. Important discoveries include the identification of the molecular architecture of Ewing sarcoma, defining Ewing sarcoma pathology, and the development of successful therapies to treat Ewing sarcoma.
Ewing sarcoma is virtually uniformly associated with a gene fusion composed of the N-terminus of an RNA-binding protein, mostly commonly EWSR1 (ES breakpoint region 1) and the carboxyl terminal DNA-binding domain of an ETS (erythroblast transformation-specific) family transcription factor (Table 1): (1) ~85% cases exhibit a t(11;22)(q24;q12) translocation, which joins EWSR1 on chromosome 22 with FLI1 (friend leukemia integration 1) on chromosome 11, resulting in the EWSR1::FLI1 onco-fusion gene; (2) ~10% cases bear a t(21;12)(q22;q12) translocation, generating an EWSR1::ERG fusion gene; (3) and the remaining 5% cases contain other EWSR1-ETS fusion genes such as EWSR1::FEV, EWSR1::ETV1, EWSR1::ETV4 and others [6,7,8,9,10,11]. FUS and TAF15, other TAF family RNA-binding proteins, are also involved in the fusions with ETS proteins in a small percentage of Ewing sarcoma [12] (Table 1). Moreover, a reciprocal chromosomal translocation resulting in FLI1-EWSR1 fusion was also observed and reported in Ewing sarcoma cell lines and tumors [13]. The exact origin(s) for Ewing sarcoma remains unclear although mesenchymal stem cells [14], neural stem cells [15] and osteochondrogenic progenitors [16] have been proposed as candidates. To date there is no genetic murine model to mimic human Ewing sarcoma.
In this review, we summarize the up-to-date understanding of the pathology and molecular features for Ewing sarcoma, and various EWSR1::FLI1 regulatory mechanisms especially post-translational modifications, aiming to provide new insights for identifying novel drug targets to fight against this deadly pediatric cancer.

2. The Pathology of Ewing Sarcoma

Ewing sarcoma is not considered a familial cancer, although a genetic predisposition has been identified in population level studies [17]. Histologically, the tumors consist of generally uniform round cells with vesicular nuclei of finely dispersed chromatin and hyaline cytoplasm [18].More than 95% of tumors express the cell surface protein CD99 (also named MIC2), which has been used as a marker for Ewing sarcoma [19]. However, CD99 expression is not specific for Ewing sarcoma. CD99 is also expressed in certain normal tissues and other mesenchymal tumors. However, negative CD99 immunogenicity strongly argues against the diagnosis of Ewing sarcoma. FLI1 staining in nuclei is more specific for clinically defining Ewing sarcoma [20]. However, IHC is not sufficient for the diagnosis of Ewing’s sarcoma, and molecular translocation testing is required to exclude other round-cell sarcomas, per the guidelines from the 2020 WHO classification [21]. Other markers such as neuron-specific enolase (NSE), S-100 protein, CD57, neurofilaments, cytokeratin, desmin, caveolin-1, NK2 homeobox 2 (Nkx-2.2) or immunohistochemical markers such as B-cell CLL/lymphoma 11B (BCL11B) and Golgi glycoprotein 1 (GLG1) have been investigated in Ewing sarcoma diagnosis, especially in cases with negative CD99 detection [22,23,24,25,26,27,28]. In addition, alterations in DNA methylation have also been observed to distinguish Ewing sarcoma [29]. Together, currently a commonly accepted standard for the diagnosis of Ewing sarcoma includes a consistent histological morphology, staining for CD99, and evidence of EWSR1 rearrangement by fluorescence in situ hybridization, PCR or sequencing.

3. Therapies in Ewing Sarcoma

For nearly 40 years, the treatment for Ewing sarcoma has included systemic chemotherapy and local treatments, including radiotherapy and surgery. If surgery is not feasible or highly morbid, then radiation therapy may be an exclusive treatment. Radiotherapy can be combined with surgery when adequate surgical margins are anticipated to be difficult to achieve or margins are found to be positive following resection. Doxorubicin, vincristine, cyclophosphamide, etoposide and ifosfamide are the standard of care for patients in the US. Although the 5-year survival rate has been improved from less than 20% to greater than 70%, the recurrence rate remains high, with some patients relapsing many years from the end of treatment [30]. For relapsed patients, survival rates are less than 30%. Surgery (uncommon except for local recurrences), radiotherapy and chemotherapy (irinotecan and temozolomide) are used for the treatment with survival rates of up to 50% in selected patients [31,32]. Thus, there remains an urgent medical need to develop new treatments for Ewing sarcoma. Targeted treatments require a better understanding of the molecular mechanisms driving the development of Ewing sarcoma and chemo-resistance.
Although the identification of a tumor-specific oncogenic fusion on which Ewing sarcoma is dependent points to the perfect drug target, at present there is not an FDA-approved therapeutic. This is partially due to the challenges in directly targeting EWSR1::FLI1 [33,34]. YK-4-279 was identified as an interactor of RNA helicase A, which interacts with EWSR1-FLI1 [35], although a recent study suggests that TK216, the related molecule in clinical testing, targets microtubules [36]. Histone deacetylase inhibitors have also been shown to affect Ewing sarcoma cell proliferation possibly through EWSR1::FLI1-mediated transcription and chromatin regulation, although they have yet to demonstrate a clear signal in clinical testing [37,38,39,40,41,42,43,44,45]. Other therapies in clinical testing include an LSD1 inhibitor, that has been shown to reverse the EWSR1::FLI1 transcriptional program [46]. LSD1 (lysine specific demethylase 1) is a protein lysine demethylase first identified to demethylate H3K4me1/2 [47] and later shown to demethylate non-histone proteins [48]. EWSR1::FLI1 recruits LSD1 to NuRD (nucleosome remodeling and histone deacetylase) complexes to suppress transcription of genes including LOX and TGFBR2 [49]. Phase I trials have demonstrated efficacy of the combination of the PARP inhibitor talazoparib with temozolomide [50]. Other mechanisms for potential therapeutic targeting include competition of EWSR1::FLI1 with MRTFB on binding to gene promoters leading to the suppression of MRTFB-mediated transcription of TAZ (also known as transcriptional coactivator with PDZ-binding motif) [34]. As a downstream effector of the Hippo signaling pathway, TAZ has been shown to facilitate cell migration through multiple mechanisms including limiting cytoskeletal and focal adhesion maturation [51], altering metabolic programs [52] and others. Although EWSR1::FLI1 is indispensable to maintain Ewing sarcoma growth, increased expression of EWSR1::FLI1 is also not tolerated as it causes cell growth arrest and cell death [53]. Thus, EWSR1::FLI1 protein homeostasis is tightly controlled, and variability in EWSR1-FLI1 might enhance metastatic potential [54].

4. Oncogenic Mechanisms of EWSR1-FLI1 in Ewing Sarcoma

EWSR1::FLI1 is necessary to maintain Ewing sarcoma proliferation and survival and exerts an ability to transform human primary mesenchymal stem cells [55,56]. There are two common types of EWSR1::FLI1 fusions in patients: type 1 (fused by EWSR1 exon 7 with FLI1 exon 6) and type 2 (fused by EWSR1 exon 7 with FLI1 exon 5) fusions. EWSR1 contributes an unstructured domain that engages in phase separation and modulates transcriptional control and RNA splicing through interactions with the SWI/SNF complex and HNRNPs [57,58]. Both type 1 and type 2 EWSR1::FLI1 fusions retain the C-terminal DNA-binding domain of FLI1, thus serving as novel aberrant transcription activators. EWSR1::FLI1 binds to microsatellite regions that contain repeats of GGAA, the core of the ETS DNA recognition elements [59,60,61]. EWSR1::FLI1 binding promotes chromatin accessibility, neo-enhancer development and transcription regulation [60]. In addition, EWSR1::FLI1 also directly binds enhancer elements to modulate gene expression [62]. Overall, the chimeric EWSR1::FLI1 transcription factor promotes malignant transformation by regulating the transcription of a large number of downstream target genes [63].
Multiple studies have explored which EWSR1::FL1 transcriptional targets account for its transformation capacity and may serve as possible viable drug targets. Early studies recovered 99 putative transcription factors co-immunoprecipitated with EWSR1::FLI1-bound chromatin. MK-STYX (a MAPK phosphatase-like protein) identified by this approach was further validated [64] that in part mediates the oncogenic properties of EWSR1-FLI1. Moreover, a microarray analysis in Ewing sarcoma A673 cells revealed that depletion of endogenous EWSR1::FLI1 by retroviral siRNAs upregulated 320 genes and downregulated 1151 genes, among which NKX2.2 was reported as a critical EWSR1-FLI1 downstream target [65]. A meta-analysis further defined a “core” EWSR1-FLI1 transcriptional signature [66] by integrating transcriptional profiling data from distinct cell line models including NIH-3T3 [67,68], primary human fibroblasts [69], primary bone marrow-derived mesenchymal progenitor cells [55], mesenchymal stem cells [70], rhabdomyosarcoma cells [71], neuroblastomas [72], patient-derived Ewing sarcoma cell lines [73] and others. These EWSR1::FLI1 targets include transcription factors such as NKX2.2, GLI1, FOXM1, DAX-1, secreted proteins such as cholecystokinin and LOX, neural crest developmental genes such as MAPT [71], cell cycle regulators such as p21 [74], as well as kinases such as PIM3 [68], AURKA and AURKB [75,76,77,78]. However, EWSR1::FLI1 target identification outside of the context of cell-of-origin must be interpreted with caution. Recently, neo-transcripts have also been identified from silent genome regions uniquely activated by the EWSR1::FLI1 fusion, suggesting these neo-genes might be targetable for Ewing sarcoma treatment [79].
In addition to protein-coding genes, the long noncoding RNA EWSAT1 was found as an EWSR1::FLI1-induced product by RNA-seq analysis using primary pediatric human mesenchymal progenitor cells [80]. EWSR1::FLI1 repressed miR-708 expression to indirectly induce EYA3 transcription [81], and inhibited expression of the tumor suppressive miR-145 [82,83], and other miRNAs including miR-22, miR29a, miR-100, miR-125b, miR-221/222 and miR-271 [84] to modulate tumor growth.
Moreover, EWSR1::FLI1 not only directly modulates the transcription, but also controls transcript degradation [85] and alternative splicing [86] as additional regulatory mechanisms to govern RNA abundance.
In addition to directly modulating chromatin conformation and gene expression, EWSR1::FLI1 also induces genome instability that facilitates tumorigenesis. EWSR1::FLI1 promotes transcription to cause R-loops, which titrate BRCA1 away from sensing damaged DNA thus blocking BRCA1-mediated DNA damage repair [87]. Together, EWSR1::FLI1 utilizes at least three distinct mechanisms to promote Ewing sarcoma growth (Figure 1).
The EWSR1::FLI1 is located in granules in nuclei [88]. An unstructured domain in EWSR1 mediates the phase transition of EWSR1::FLI1 and enhancer activation [89]. Similarly, the low-complexity domain interactions among EWSR1::FLI1 are also reported (which are transient and dynamic) to promote transcriptional activity at a narrow optimal level [90]. Increasing concentrations of EWSR1-FLI1 low-complexity domain interactions promotes EWSR1::FLI1 phase transition in the nucleolus and suppresses the EWSR1::FLI1 transcriptional [91]. Thus, depending on the levels of EWSR1::FLI1 low-complexity domain interactions, phase transition can either promote or suppress transcription.

5. EWSR1-FLI1 Regulatory Mechanisms

5.1. Transcriptional Regulation

Transcriptional activation of the EWSR1::FLI1 fusion gene was accompanied by deposition of histone markers on the EWSR1 promoter [92]. Analysis of clinical cases by conventional cytogenetics, fluorescence in situ hybridization and nested PCR revealed that H3K4me3, H3K9ac and H3K27ac were significantly enriched in the EWSR1 promoter in Ewing sarcoma to facilitate transcription [93,94].The transcription factor SP1 directly binds to the EWSR1::FLI1 promoter to trigger transcription, a process induced by activated PI3K/Akt signaling [95]. In addition, hypoxia promotes EWSR1::FLI1 transcription in a HIF-1a-dependent manner [96].
The RNA-binding protein HNRNPH1 was shown to facilitate Ewing sarcoma cells to properly express EWSR1 exon 8 genomic breakpoint fusions [58,97]. The stability of EWSR1::FLI1 mRNA in approximately 10% of Ewing sarcomas is regulated by the carcinoembryonic RNA-binding protein LIN28B. Deletion of LIN28B led to decreased EWSR1::FLI1 expression, which affects the self-renewal and tumorigenicity of Ewing sarcoma cells [98]. Knocking down CRM1 (XPO1) significantly inhibited the expression of EWSR1::FLI1 fusion proteins at the post-transcriptional level with unknown mechanism(s) [99]. On the other hand, miR-145 was shown to suppress EWSR1::FLI1 transcription [82]. A small-molecule screen identified that histone deacetylase inhibitors decrease EWSR1::FLI1 levels, possibly contributing to the cytotoxic effect of these drugs on cells [40,45].

5.2. Translational Regulation

The stability of de novo synthesized EWSR1-FLI1 proteins can be reduced by treatment with lovastatin or tunicamycin, leading to reduced protein levels and decreased Ewing sarcoma cell growth [1,2,100]. In addition to suppressing EWSR1::FLI1 protein synthesis, tunicamycin also suppressed N-linked glycosylation (likely N-linked glycosylation of IGF1-R) that suppresses function and cell growth [101]. Notably, although EWSR1::FLI1 expression contributes to a proliferative phenotype, reduced levels of EWSR1::FLI1 proteins have been shown to decrease proliferation but induce a more motile phenotype [33,34].

5.3. Protein-Level Regulation

EWSR1::FLI1 is also regulated post-translationally (Figure 2). This includes regulation of the physical properties of EWSR1-FLI1 protein by various post-translational modifications that regulate protein function in an acute and spatial manner, as well as various binding proteins that either facilitate or suppress EWSR1::FLI1 function on chromatin.
Post-translational control: Multiple post-translational modifications of EWSR1::FLI1 have been described that regulate transcriptional activity and function in both tempo- and spatial manners (Figure 2). Specifically, EWSR1::FLI1 is phosphorylated at Thr79 in the N-terminal EWSR1 domain upon DNA damage or mitogen stimulations by ERK1/ERK2/JNK or p38-MAPKs, respectively. This modification presumably stimulates dimer formation and transcriptional activity [102]. O-GlcNAcylation of EWSR1::FLI1 was also reported to positively regulate oncogenic function in Ewing sarcoma [103]. In addition, acetylation of the C-terminal FLI1 region by PCAF increases its DNA-binding ability to potentiate transcriptional activity [104]. Ubiquitination of EWSR1::FLI1 is observed on the Lys380 residue that primes the protein for proteasomal degradation [105]. In addition, lysosome-dependent protein degradation is also reported [106]. Multiple E3 ubiquitin ligases have been observed to mediate EWSR1::FLI1 ubiquitination and degradation in Ewing sarcoma, including TRIM8 [53] and SPOP [107]. Deubiquitinases, including USP19 [108] and OTUD7A [107], stabilize EWSR1-FLI1 protein. Notably, CK1-mediated Ser486/487/488 (based on type-II variant) phosphorylation primed EWSR1::FLI1 for recognition and regulation by either SPOP or OTUD7A [107]. Recently, we identified 7Ai, a putative small-molecule OTUD7A inhibitor that suppresses EWSR1::FLI1 protein expression and subsequent Ewing sarcoma cell and tumor growth [107]. Notably, it is currently largely unclear if these modifications are also present in the native proteins.
Binding proteins: Control of EWSR1::FLI1 transcriptional activity is also achieved by multiple protein interactors. The RNA helicase A (RHA) binds EWSR1::FL1 to enhance its transcriptional activity [109]. PARP-1 also interacts with EWSR1::FLI1 to facilitate transcription [110]. hsRBP7, as a subunit of RNA polymerase holozyme II (Pol II) interacts with EWSR1::FLI1 through its EWSR1 portion [111]. EWSR1::FLI1 also complexes with EWSR1 (and with RNA Pol II) to exert its transcriptional activity. EWSR1::FLI1 forms a ternary complex with ELK1-SAP1a to bind SRF using the unique R-domain near the FLI1 DNA-binding region to upregulate ERG1 expression [112]. BARD1 as a putative tumor suppressor, interacts with the N-terminus of EWSR1::FLI1 [113]. The FOS-JUN dimer also interacts with EWSR1::FLI1 to bind AP1 sequences [114]. Additionally, EWSR1::FLI1 recruits the BAF complex to tumor-specific enhancers to promote activation of target genes [89]. Steroid-dependent translocation of EWSR1::FLI1 and glucocorticoid receptor into nuclei leads to EWSR1::FLI1 binding to the glucocorticoid receptor to enhance glucocorticoid receptor-mediated oncogenic transcription to facilitate Ewing sarcoma growth and migration [115]. Proteomics analysis also reveals that CIMPR (cation-independent mannose 6-phosphate receptor) as a EWSR1::FLI1 binding partner that regulates EWSR1::FLI1 degradation in a lysosome-dependent pathway [106]. Although these binding proteins exert important but distinct roles in Ewing sarcoma, whether they participate in different protein sub-complexes at various cellular compartments or under distinct pathophysiological functions remains to be further determined. Notably, to date there is no genetic mouse model to study Ewing sarcoma pathology, biology and the testing of therapeutics’ effects although zebrafish [116] and drosophila [117] models have been developed. There are limited number of established PDX murine models using both subcutaneous and tibial implantation.

6. Concluding Remarks and Future Perspectives

Chromosomal translocations between chr11 and chr22 occur specifically in Ewing sarcoma to generate the fusion oncogene EWSR1::FLI1 [118,119,120]. The exact mechanism resulting in this chromosomal translocation remains elusive. Although EWSR1::FLI1 presents a unique therapeutic target in Ewing sarcoma, to date no effective targeted therapies have been validated and approved to treat Ewing sarcoma. The first-line therapy in clinic relies on intensive chemotherapy combined with surgery and radiation. Presently, there is no standard of care for chemo-resistant, relapsed patients.
Firstly, more efforts in the molecular classification of Ewing sarcoma tumor subtypes would benefit the discovery of vulnerabilities and development of effective cures. Loss of TP53 and STAG2 is associated with a poor outcome [121]. Unlike breast cancer that is classified into basal-like, luminal and other types of tumors based on the genetic architecture, and prostate cancer is divided into AR+/AR or castration-resistant subtypes, treatment relevant molecular subtypes have yet to be identified. This effort may explain the heterogeneity observed in Ewing sarcoma [122] and may help further direct proper treatments/combination treatments for subtypes of Ewing sarcoma patients to improve treatment efficacy. However, the rarity of Ewing sarcoma limits the availability of a large cohort of patient samples that can be used to faithfully perform molecular subtyping. In addition, how to incorporate this information into clinical designs remains as another challenge. Notably, efforts in genotyping Ewing sarcomas [123,124,125,126] have begun to shed new light onto the Ewing sarcoma pathology.
Secondly, establishing genetic models that recapitulate Ewing sarcoma initiation and progression would help identify those key molecular events necessary for tumor development—although the unique positional relationship between GGAA-containing microsatellite enhancers and relevant genes unique to the human genome may preclude this approach. Although EWSR1::FLI1 is the driver for Ewing sarcoma, its expression in nontumor cells induces apoptosis and fails to promote tumor formation, bringing challenges in establishing genetic models and suggesting the relevance of additional genetic alternations and cellular context [127,128]. Given that Ewing sarcoma zebrafish models require the deletion of p53 or expression of anti-apoptotic BCL-2 family proteins [129], additional genetic manipulations may be necessary to facilitate EWSR1::FLI1-driven mouse models. Establishing a valid genetic murine model will greatly facilitate the understanding of key steps in Ewing sarcoma initiation and development, which in turn will facilitate new biomarker identification and drug target discovery.
Thirdly, developing additional patient-derived xenografts (PDX), as well as immune-competent murine models would benefit preclinical studies. To date, validation of effects of gene function or treatment effects in Ewing sarcoma largely relies on in vitro cell-line studies and xenografted mouse models in immunodeficient settings [130,131]. PDX mouse models mimic aspects of cancer development [132]. A limited number of Ewing sarcoma PDX models (https://www.pdxfinder.org/, accessed on 1 November 2022) have been established, and both flank transplantation and orthotopic transplantation (or -patient-derived orthotopic xenograft (PDOX)) have been successfully developed [131,133,134]. The use of PDX and PDOX models allows for optimized conditions for drug development and precise cancer therapy. In addition, establishing Ewing sarcoma cell lines derived from patients that can survive in immune-competent mice would greatly facilitate the studies using immune-checkpoint blockades and CAR-T cell therapies in Ewing sarcoma.
Lastly, understanding drug resistance mechanisms would provide new avenues for treatment modality. Efforts have been devoted to deciphering critical downstream transcriptional targets of EWSR1::FLI1, its binding partners, modifications and roles in regulating Ewing sarcoma proliferation and motility. Given Ewing sarcoma is a systemic disease, biomarkers that predict response to chemotherapy would be relevant. Although ~50% patients respond to chemotherapy initially, once relapsed, most patients are insensitive to chemotherapy. In addition, frequent intensive chemotherapy decreases patients’ quality of life. How to identify patients who are likely to do well with reduced intensity treatment first-line chemotherapy while preserving the current high efficacy, and how to augment therapy for patients with a high risk of relapse remains a critical yet unsatisfactory question. Thus, understanding the molecular mechanism(s) leading to chemo-resistance in relapsed patients may reveal possible novel drug targets and combination therapies to improve chemotherapy efficacy. Multiple treatment-resistance mechanisms will be identified, and it will be important to identify driver from passenger events among these mechanisms. More importantly, developing targeted therapies or precision medicine for each individual patient based on the unique genetic signatures from patients would be a pivotal direction to improve the treatment outcomes. This requires more in-depth investigations on both molecular mechanistic studies as well as pre-clinical and clinical examinations of newly proposed therapies. The development of PDX or PDOX models in immune-competent genetic murine models could facilitate this process.

Author Contributions

Information collection, L.Y.; writing—original draft preparation, L.Y.; writing—review and editing, I.J.D. and P.L.; figure/table construction, L.Y.; supervision, P.L.; funding acquisition, P.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Andrew McDonough B+ Foundation Award (P.L.), the UNC University Cancer Research Fund (P.L.), and NIH P30CA016086 (I.J.D).

Acknowledgments

We sincerely apologize to all colleagues whose important work could not be cited in this review owing to space limitations, especially prominent and pioneer works in the Ewing sarcoma field. We thank other Liu Lab members for critical reading of the manuscript and helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grunewald, T.G.P.; Cidre-Aranaz, F.; Surdez, D.; Tomazou, E.M.; de Alava, E.; Kovar, H.; Sorensen, P.H.; Delattre, O.; Dirksen, U. Ewing sarcoma. Nat. Rev. Dis. Prim. 2018, 4, 5. [Google Scholar] [CrossRef]
  2. Zöllner, S.; Amatruda, J.; Bauer, S.; Collaud, S.; de Álava, E.; DuBois, S.; Hardes, J.; Hartmann, W.; Kovar, H.; Metzler, M.; et al. Ewing Sarcoma—Diagnosis, Treatment, Clinical Challenges and Future Perspectives. J. Clin. Med. 2021, 10, 1685. [Google Scholar] [CrossRef] [PubMed]
  3. Torre, L.A.; Bray, F.; Siegel, R.L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A. Global cancer statistics, 2012. CA Cancer J. Clin. 2015, 65, 87–108. [Google Scholar] [CrossRef] [Green Version]
  4. Cotterill, S.; Ahrens, S.; Paulussen, M.; Jürgens, H.; Voûte, P.; Gadner, H.; Craft, A. Prognostic Factors in Ewing’s Tumor of Bone: Analysis of 975 Patients From the European Intergroup Cooperative Ewing’s Sarcoma Study Group. J. Clin. Oncol. 2000, 18, 3108–3114. [Google Scholar] [CrossRef] [PubMed]
  5. Stahl, M.; Ranft, A.; Paulussen, M.; Bölling, T.; Vieth, V.; Bielack, S.; Görtitz, I.; Braun-Munzinger, G.; Hardes, J.; Jürgens, H.; et al. Risk of recurrence and survival after relapse in patients with Ewing sarcoma. Pediatr. Blood Cancer 2011, 57, 549–553. [Google Scholar] [CrossRef] [PubMed]
  6. Desmaze, C.; Brizard, F.; Turc-Carel, C.; Melot, T.; Delattre, O.; Thomas, G.; Aurias, A. Multiple chromosomal mechanisms generate an EWS/FLI1 or an EWS/ERG fusion gene in Ewing tumors. Cancer Genet. Cytogenet. 1997, 97, 12–19. [Google Scholar] [CrossRef]
  7. Forscher, C.; Figlin, R.; Mita, M. Targeted therapy for sarcomas. Biol. Targets Ther. 2014, 8, 91–105. [Google Scholar] [CrossRef] [Green Version]
  8. Kauer, M.; Ban, J.; Kofler, R.; Walker, B.; Davis, S.; Meltzer, P.; Kovar, H. A molecular function map of Ewing’s sarcoma. PLoS ONE 2009, 4, e5415. [Google Scholar] [CrossRef] [Green Version]
  9. Osuna, D.; de Alava, E. Molecular pathology of sarcomas. Rev. Recent Clin. Trials 2009, 4, 12–26. [Google Scholar] [CrossRef] [PubMed]
  10. Delattre, O.; Zucman, J.; Plougastel, B.; Desmaze, C.; Melot, T.; Peter, M.; Kovar, H.; Joubert, I.; De Jong, P.; Rouleau, G.; et al. Gene fusion with an ETS DNA-binding domain caused by chromosome translocation in human tumours. Nature 1992, 359, 162–165. [Google Scholar] [CrossRef]
  11. Kovar, H.; Aryee, D.N.; Jug, G.; Henöckl, C.; Schemper, M.; Delattre, O.; Thomas, G.; Gadner, H. EWS/FLI-1 antagonists induce growth inhibition of Ewing tumor cells in vitro. Cell Growth Differ. 1996, 7, 429–437. [Google Scholar] [PubMed]
  12. Shing, D.C.; McMullan, D.J.; Roberts, P.; Smith, K.; Chin, S.F.; Nicholson, J.; Tillman, R.M.; Ramani, P.; Cullinane, C.; Coleman, N. FUS/ERG gene fusions in Ewing’s tumors. Cancer Res. 2003, 63, 4568–4576. [Google Scholar] [PubMed]
  13. Elzi, D.J.; Song, M.; Houghton, P.J.; Chen, Y.; Shiio, Y. The role of FLI-1-EWS, a fusion gene reciprocal to EWS-FLI-1, in Ewing sarcoma. Genes Cancer 2015, 6, 452–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Lin, P.P.; Wang, Y.; Lozano, G. Mesenchymal Stem Cells and the Origin of Ewing’s Sarcoma. Sarcoma 2011, 2011, 276463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. von Levetzow, C.; Jiang, X.; Gwye, Y.; von Levetzow, G.; Hung, L.; Cooper, A.; Hsu, J.H.; Lawlor, E.R. Modeling initiation of Ewing sarcoma in human neural crest cells. PLoS ONE 2011, 6, e19305. [Google Scholar] [CrossRef] [Green Version]
  16. Tanaka, M.; Yamazaki, Y.; Kanno, Y.; Igarashi, K.; Aisaki, K.-I.; Kanno, J.; Nakamura, T. Ewing’s sarcoma precursors are highly enriched in embryonic osteochondrogenic progenitors. J. Clin. Investig. 2014, 124, 3061–3074. [Google Scholar] [CrossRef] [Green Version]
  17. Grünewald, T.G.P.; Bernard, V.; Gilardi-Hebenstreit, P.; Raynal, V.; Surdez, D.; Aynaud, M.-M.; Mirabeau, O.; Cidre-Aranaz, F.; Tirode, F.; Zaidi, S.; et al. Chimeric EWSR1-FLI1 regulates the Ewing sarcoma susceptibility gene EGR2 via a GGAA microsatellite. Nat. Genet. 2015, 47, 1073–1078. [Google Scholar] [CrossRef] [Green Version]
  18. Ross, K.; Smyth, N.A.; Murawski, C.D.; Kennedy, J.G. The Biology of Ewing Sarcoma. ISRN Oncol. 2013, 2013, 759725. [Google Scholar] [CrossRef] [Green Version]
  19. Marcilla, D.; Machado, I.; Grünewald, T.G.P.; Llombart-Bosch, A.; de Álava, E. (Immuno)histological Analysis of Ewing Sarcoma. Methods Mol. Biol. 2021, 2226, 49–64. [Google Scholar] [CrossRef]
  20. Nilsson, G.; Wang, M.; Wejde, J.; Kreicbergs, A.; Larsson, O. Detection of EWS/FLI-1 by Immunostaining. An Adjunctive Tool in Diagnosis of Ewing’s Sarcoma and Primitive Neuroectodermal Tumour on Cytological Samples and Paraffin-Embedded Archival Material. Sarcoma 1999, 3, 25–32. [Google Scholar] [CrossRef]
  21. Sbaraglia, M.; Bellan, E.; Tos, A.P.D. The 2020 WHO Classification of Soft Tissue Tumours: News and perspectives. Pathologica 2020, 113, 70–84. [Google Scholar] [CrossRef] [PubMed]
  22. Kang, M.S.; Yoon, H.K.; Choi, J.B.; Eum, J.W. Extraskeletal Ewing’s sarcoma of the hard palate. J. Korean Med. Sci. 2005, 20, 687–690. [Google Scholar] [CrossRef] [Green Version]
  23. Collini, P.; Mezzelani, A.; Modena, P.; Dagrada, P.; Tamborini, E.; Luksch, R.; Gronchi, A.; Navarria, P.; Sozzi, G.; Pilotti, S. Evidence of Neural Differentiation in a Case of Post-therapy Primitive Neuroectodermal Tumor/Ewing Sarcoma of Bone. Am. J. Surg. Pathol. 2003, 27, 1161–1166. [Google Scholar] [CrossRef]
  24. Amann, G.; Zoubek, A.; Salzer-Kuntschik, M.; Windhager, R.; Kovar, H. Relation of neurological marker expression and EWS gene fusion types in MIC2/CD99-positive tumors of the Ewing family. Hum. Pathol. 1999, 30, 1058–1064. [Google Scholar] [CrossRef]
  25. Shi, X.; Zheng, Y.; Jiang, L.; Zhou, B.; Yang, W.; Li, L.; Ding, L.; Huang, M.; Gery, S.; Lin, D.-C.; et al. EWS-FLI1 regulates and cooperates with core regulatory circuitry in Ewing sarcoma. Nucleic Acids Res. 2020, 48, 11434–11451. [Google Scholar] [CrossRef]
  26. Boulay, G.; Volorio, A.; Iyer, S.; Broye, L.C.; Stamenkovic, I.; Riggi, N.; Rivera, M.N. Epigenome editing of microsatellite repeats defines tumor-specific enhancer functions and dependencies. Genes Dev. 2018, 32, 1008–1019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Boro, A.; Pretre, K.; Rechfeld, F.; Thalhammer, V.; Oesch, S.; Wachtel, M.; Schafer, B.W.; Niggli, F.K. Small-molecule screen identifies modulators of EWS/FLI1 target gene expression and cell survival in Ewing’s sarcoma. Int. J. Cancer 2012, 131, 2153–2164. [Google Scholar] [CrossRef]
  28. Jully, B.; Vijayalakshmi, R.; Gopal, G.; Sabitha, K.; Rajkumar, T. Junction region of EWS-FLI1 fusion protein has a dominant negative effect in Ewing’s Sarcoma in vitro. BMC Cancer 2012, 12, 513. [Google Scholar] [CrossRef] [Green Version]
  29. Park, H.-R.; Jung, W.-W.; Kim, H.-S.; Park, Y.-K. Microarray-based DNA methylation study of Ewing’s sarcoma of the bone. Oncol. Lett. 2014, 8, 1613–1617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Liebner, D.A. The indications and efficacy of conventional chemotherapy in primary and recurrent sarcoma. J. Surg. Oncol. 2015, 111, 622–631. [Google Scholar] [CrossRef]
  31. Strauss, S.J.; Frezza, A.M.; Abecassis, N.; Bajpai, J.; Bauer, S.; Biagini, R.; Bielack, S.; Blay, J.Y.; Bolle, S.; Bonvalot, S.; et al. Bone sarcomas: ESMO-EURACAN-GENTURIS-ERN PaedCan Clinical Practice Guideline for diagnosis, treatment and follow-up. Ann. Oncol. 2021, 32, 1520–1536. [Google Scholar] [CrossRef]
  32. Ferrari, A.; Dirksen, U.; Bielack, S. Sarcomas of Soft Tissue and Bone. Prog Tumor Res. 2016, 43, 128–141. [Google Scholar] [CrossRef]
  33. Franzetti, G.-A.; Laud-Duval, K.; van der Ent, W.; Brisac, A.; Irondelle, M.; Aubert, S.; Dirksen, U.; Bouvier, C.; De Pinieux, G.; Snaar-Jagalska, E.; et al. Cell-to-cell heterogeneity of EWSR1-FLI1 activity determines proliferation/migration choices in Ewing sarcoma cells. Oncogene 2017, 36, 3505–3514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Katschnig, A.M.; Kauer, M.O.; Schwentner, R.; Tomazou, E.M.; Mutz, C.N.; Linder, M.; Sibilia, M.; Alonso, J.; Aryee, D.N.T.; Kovar, H. EWS-FLI1 perturbs MRTFB/YAP-1/TEAD target gene regulation inhibiting cytoskeletal autoregulatory feedback in Ewing sarcoma. Oncogene 2017, 36, 5995–6005. [Google Scholar] [CrossRef] [Green Version]
  35. Erkizan, H.V.; Kong, Y.; Merchant, M.; Schlottmann, S.; Barber-Rotenberg, J.S.; Yuan, L.; Abaan, O.D.; Chou, T.H.; Dakshanamurthy, S.; Brown, M.L.; 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, 750–756. [Google Scholar] [CrossRef] [Green Version]
  36. Povedano, J.M.; Li, V.; Lake, K.E.; Bai, X.; Rallabandi, R.; Kim, J.; Xie, Y.; De Brabander, J.K.; McFadden, D.G. TK216 targets microtubules in Ewing sarcoma cells. Cell Chem. Biol. 2022, 29, 1325–1332.e4. [Google Scholar] [CrossRef]
  37. Bukowinski, A.; Chang, B.; Reid, J.M.; Liu, X.; Minard, C.G.; Trepel, J.B.; Lee, M.J.; Fox, E.; Weigel, B.J. A phase 1 study of entinostat in children and adolescents with recurrent or refractory solid tumors, including CNS tumors: Trial ADVL1513, Pediatric Early Phase-Clinical Trial Network (PEP-CTN). Pediatr. Blood Cancer 2021, 68, e28892. [Google Scholar] [CrossRef]
  38. Souza, B.K.; da Costa Lopez, P.L.; Menegotto, P.R.; Vieira, I.A.; Kersting, N.; Abujamra, A.L.; Brunetto, A.T.; Brunetto, A.L.; Gregianin, L.; de Farias, C.B.; et al. Targeting Histone Deacetylase Activity to Arrest Cell Growth and Promote Neural Differentiation in Ewing Sarcoma. Mol. Neurobiol. 2018, 55, 7242–7258. [Google Scholar] [CrossRef] [PubMed]
  39. Garcia-Dominguez, D.J.; Hontecillas-Prieto, L.; Rodriguez-Nunez, P.; Pascual-Pasto, G.; Vila-Ubach, M.; Garcia-Mejias, R.; Robles, M.J.; Tirado, O.M.; Mora, J.; Carcaboso, A.M.; et al. The combination of epigenetic drugs SAHA and HCI-2509 synergistically inhibits EWS-FLI1 and tumor growth in Ewing sarcoma. Oncotarget 2018, 9, 31397–31410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Pattenden, S.G.; Simon, J.M.; Wali, A.; Jayakody, C.N.; Troutman, J.; McFadden, A.W.; Wooten, J.; Wood, C.C.; Frye, S.V.; Janzen, W.P.; et al. High-throughput small molecule screen identifies inhibitors of aberrant chromatin accessibility. Proc. Natl. Acad. Sci. USA 2016, 113, 3018–3023. [Google Scholar] [CrossRef]
  41. Zorzi, A.P.; Bernstein, M.; Samson, Y.; Wall, D.A.; Desai, S.; Nicksy, D.; Wainman, N.; Eisenhauer, E.; Baruchel, S. A phase I study of histone deacetylase inhibitor, pracinostat (SB939), in pediatric patients with refractory solid tumors: IND203 a trial of the NCIC IND program/C17 pediatric phase I consortium. Pediatr. Blood Cancer 2013, 60, 1868–1874. [Google Scholar] [CrossRef]
  42. Witt, O.; Milde, T.; Deubzer, H.E.; Oehme, I.; Witt, R.; Kulozik, A.; Eisenmenger, A.; Abel, U.; Karapanagiotou-Schenkel, I. Phase I/II intra-patient dose escalation study of vorinostat in children with relapsed solid tumor, lymphoma or leukemia. Klin Padiatr. 2012, 224, 398–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Owen, L.A.; Kowalewski, A.A.; Lessnick, S.L. EWS/FLI mediates transcriptional repression via NKX2.2 during oncogenic transformation in Ewing’s sarcoma. PLoS ONE 2008, 3, e1965. [Google Scholar] [CrossRef] [Green Version]
  44. Sonnemann, J.; Dreyer, L.; Hartwig, M.; Palani, C.D.; Hong, L.T.T.; Klier, U.; Bröker, B.; Völker, U.; Beck, J.F. Histone deacetylase inhibitors induce cell death and enhance the apoptosis-inducing activity of TRAIL in Ewing’s sarcoma cells. J. Cancer Res. Clin. Oncol. 2007, 133, 847–858. [Google Scholar] [CrossRef]
  45. Sakimura, R.; Tanaka, K.; Nakatani, F.; Matsunobu, T.; Li, X.; Hanada, M.; Okada, T.; Nakamura, T.; Matsumoto, Y.; Iwamoto, Y. Antitumor effects of histone deacetylase inhibitor on Ewing’s family tumors. Int. J. Cancer 2005, 116, 784–792. [Google Scholar] [CrossRef] [PubMed]
  46. Pishas, K.I.; Drenberg, C.D.; Taslim, C.; Theisen, E.R.; Johnson, K.M.; Saund, R.S.; Pop, I.L.; Crompton, B.D.; Lawlor, E.R.; Tirode, F.; et al. Therapeutic Targeting of KDM1A/LSD1 in Ewing Sarcoma with SP-2509 Engages the Endoplasmic Reticulum Stress Response. Mol. Cancer Ther. 2018, 17, 1902–1916. [Google Scholar] [CrossRef] [Green Version]
  47. Shi, Y.; Lan, F.; Matson, C.; Mulligan, P.; Whetstine, J.R.; Cole, P.A.; Casero, R.A.; Shi, Y. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 2004, 119, 941–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Majello, B.; Gorini, F.; Saccà, C.D.; Amente, S. Expanding the Role of the Histone Lysine-Specific Demethylase LSD1 in Cancer. Cancers 2019, 11, 324. [Google Scholar] [CrossRef] [Green Version]
  49. Sankar, S.; Bell, R.; Stephens, B.; Zhuo, R.; Sharma, S.; Bearss, D.; Lessnick, S.L. Mechanism and relevance of EWS/FLI-mediated transcriptional repression in Ewing sarcoma. Oncogene 2013, 32, 5089–5100. [Google Scholar] [CrossRef] [Green Version]
  50. Federico, S.M.; Pappo, A.S.; Sahr, N.; Sykes, A.; Campagne, O.; Stewart, C.F.; Clay, M.R.; Bahrami, A.; McCarville, M.B.; Kaste, S.C.; et al. A phase I trial of talazoparib and irinotecan with and without temozolomide in children and young adults with recurrent or refractory solid malignancies. Eur. J. Cancer 2020, 137, 204–213. [Google Scholar] [CrossRef]
  51. Mason, D.E.; Collins, J.M.; Dawahare, J.H.; Nguyen, T.D.; Lin, Y.; Voytik-Harbin, S.L.; Zorlutuna, P.; Yoder, M.C.; Boerckel, J.D. YAP and TAZ limit cytoskeletal and focal adhesion maturation to enable persistent cell motility. J. Cell Biol. 2019, 218, 1369–1389. [Google Scholar] [CrossRef] [Green Version]
  52. Yamaguchi, H.; Taouk, G.M. A Potential Role of YAP/TAZ in the Interplay Between Metastasis and Metabolic Alterations. Front. Oncol. 2020, 10, 928. [Google Scholar] [CrossRef]
  53. Seong, B.K.A.; Dharia, N.V.; Lin, S.; Donovan, K.A.; Chong, S.; Robichaud, A.; Conway, A.; Hamze, A.; Ross, L.; Alexe, G.; et al. TRIM8 modulates the EWS/FLI oncoprotein to promote survival in Ewing sarcoma. Cancer Cell 2021, 39, 1262–1278.e7. [Google Scholar] [CrossRef] [PubMed]
  54. Bailey, K.; Cost, C.; Davis, I.; Glade-Bender, J.; Grohar, P.; Houghton, P.; Isakoff, M.; Stewart, E.; Laack, N.; Yustein, J.; et al. Emerging novel agents for patients with advanced Ewing sarcoma: A report from the Children’s Oncology Group (COG) New Agents for Ewing Sarcoma Task Force. F1000Research 2019, 8, 493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Riggi, N.; Cironi, L.; Provero, P.; Suva, M.L.; 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, 11459–11468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Riggi, N.; Suva, M.L.; Suva, D.; Cironi, L.; Provero, P.; Tercier, S.; Joseph, J.M.; Stehle, J.C.; Baumer, K.; Kindler, V.; et al. EWS-FLI-1 expression triggers a Ewing’s sarcoma initiation program in primary human mesenchymal stem cells. Cancer Res. 2008, 68, 2176–2185. [Google Scholar] [CrossRef] [Green Version]
  57. Lindén, M.; Vannas, C.; Österlund, T.; Andersson, L.; Osman, A.; Escobar, M.; Fagman, H.; Ståhlberg, A.; Åman, P. FET fusion oncoproteins interact with BRD4 and SWI/SNF chromatin remodelling complex subtypes in sarcoma. Mol. Oncol. 2022, 16, 2470–2495. [Google Scholar] [CrossRef]
  58. Neckles, C.; Boer, R.; Aboreden, N.; Walker, R.L.; Kim, B.-H.; Kim, S.; Schneekloth, J.S.; Caplen, N.J. HNRNPH1-dependent splicing of a fusion oncogene reveals a targetable RNA G-quadruplex interaction. RNA 2019, 25, 1731–1750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Gangwal, K.; Sankar, S.; Hollenhorst, P.C.; Kinsey, M.; Haroldsen, S.C.; Shah, A.A.; Boucher, K.M.; Watkins, W.S.; Jorde, L.B.; Graves, B.J.; et al. Microsatellites as EWS/FLI response elements in Ewing’s sarcoma. Proc. Natl. Acad. Sci. USA 2008, 105, 10149–10154. [Google Scholar] [CrossRef] [Green Version]
  60. Patel, M.; Simon, J.M.; Iglesia, M.D.; Wu, S.B.; McFadden, A.W.; Lieb, J.D.; Davis, I.J. Tumor-specific retargeting of an oncogenic transcription factor chimera results in dysregulation of chromatin and transcription. Genome Res. 2011, 22, 259–270. [Google Scholar] [CrossRef]
  61. Guillon, N.; Tirode, 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, e4932. [Google Scholar] [CrossRef] [Green Version]
  62. Riggi, N.; Knoechel, B.; Gillespie, S.M.; Rheinbay, E.; Boulay, G.; Suvà, M.L.; Rossetti, N.E.; Boonseng, W.E.; Oksuz, O.; Cook, E.B.; et al. EWS-FLI1 Utilizes Divergent Chromatin Remodeling Mechanisms to Directly Activate or Repress Enhancer Elements in Ewing Sarcoma. Cancer Cell 2014, 26, 668–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cidre-Aranaz, F.; Alonso, J. EWS/FLI1 Target Genes and Therapeutic Opportunities in Ewing Sarcoma. Front. Oncol. 2015, 5, 162. [Google Scholar] [CrossRef] [Green Version]
  64. Siligan, C.; Ban, J.; Bachmaier, R.; Spahn, L.; Kreppel, M.; Schaefer, K.L.; Poremba, C.; Aryee, D.N.; Kovar, H. EWS-FLI1 target genes recovered from Ewing’s sarcoma chromatin. Oncogene 2005, 24, 2512–2524. [Google Scholar] [CrossRef] [Green Version]
  65. Smith, R.; Owen, L.A.; Trem, D.J.; Wong, J.S.; Whangbo, J.S.; Golub, T.R.; Lessnick, S.L. Expression profiling of EWS/FLI identifies NKX2.2 as a critical target gene in Ewing’s sarcoma. Cancer Cell 2006, 9, 405–416. [Google Scholar] [CrossRef] [Green Version]
  66. Hancock, J.D.; Lessnick, S.L. A transcriptional profiling meta-analysis reveals a core EWS-FLI gene expression signature. Cell Cycle 2008, 7, 250–256. [Google Scholar] [CrossRef] [Green Version]
  67. Braunreiter, C.L.; Hancock, J.D.; Coffin, C.M.; Boucher, K.; Lessnick, S.L. Expression of EWS-ETS Fusions in NIH3T3 Cells Reveals Significant Differences to Ewing’s Sarcoma. Cell Cycle 2006, 5, 2753–2759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Deneen, B.; Welford, S.M.; Ho, T.; Hernandez, F.; Kurland, I.; Denny, C.T. PIM3 Proto-Oncogene Kinase Is a Common Transcriptional Target of Divergent EWS/ETS Oncoproteins. Mol. Cell Biol. 2003, 23, 3897–3908. [Google Scholar] [CrossRef] [Green Version]
  69. Lessnick, S.L.; Dacwag, C.S.; Golub, T.R. The Ewing’s sarcoma oncoprotein EWS/FLI induces a p53-dependent growth arrest in primary human fibroblasts. Cancer Cell 2002, 1, 393–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Tirode, F.; Laud-Duval, K.; Prieur, A.; Delorme, B.; Charbord, P.; Delattre, O. Mesenchymal Stem Cell Features of Ewing Tumors. Cancer Cell 2007, 11, 421–429. [Google Scholar] [CrossRef] [PubMed]
  71. Hu-Lieskovan, S.; Zhang, J.; Wu, L.; Shimada, H.; Schofield, D.E.; Triche, T.J. EWS-FLI1 fusion protein up-regulates critical genes in neural crest development and is responsible for the observed phenotype of Ewing’s family of tumors. Cancer Res. 2005, 65, 4633–4644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Rorie, C.J.; Thomas, V.D.; Chen, P.; Pierce, H.H.; O’Bryan, J.P.; Weissman, B.E. The Ews/Fli-1 Fusion Gene Switches the Differentiation Program of Neuroblastomas to Ewing Sarcoma/Peripheral Primitive Neuroectodermal Tumors. Cancer Res. 2004, 64, 1266–1277. [Google Scholar] [CrossRef] [Green Version]
  73. Kinsey, M.; Smith, R.; Lessnick, S.L. NR0B1 is required for the oncogenic phenotype mediated by EWS/FLI in Ewing’s sarcoma. Mol. Cancer Res. 2006, 4, 851–859. [Google Scholar] [CrossRef] [Green Version]
  74. Matsumoto, Y.; Tanaka, K.; Nakatani, F.; Matsunobu, T.; Matsuda, S.; Iwamoto, Y. Downregulation and forced expression of EWS-Fli1 fusion gene results in changes in the expression of G1regulatory genes. Br. J. Cancer 2001, 84, 768–775. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, S.; Hwang, E.E.; Guha, R.; O’Neill, A.F.; Melong, N.; Veinotte, C.J.; Saur, A.C.; Wuerthele, K.; Shen, M.; McKnight, C.; et al. High-throughput Chemical Screening Identifies Focal Adhesion Kinase and Aurora Kinase B Inhibition as a Synergistic Treatment Combination in Ewing Sarcoma. Clin. Cancer Res. 2019, 25, 4552–4566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Park, H.; Turkalo, T.K.; Nelson, K.; Folmsbee, S.S.; Robb, C.; Roper, B.; Azuma, M. Ewing sarcoma EWS protein regulates midzone formation by recruiting Aurora B kinase to the midzone. Cell Cycle 2014, 13, 2391–2399. [Google Scholar] [CrossRef] [Green Version]
  77. Winter, G.E.; Rix, U.; Lissat, A.; Stukalov, A.; Mullner, M.K.; Bennett, K.L.; Colinge, J.; Nijman, S.M.; Kubicek, S.; Kovar, H.; et al. An integrated chemical biology approach identifies specific vulnerability of Ewing’s sarcoma to combined inhibition of Aurora kinases A and B. Mol. Cancer Ther. 2011, 10, 1846–1856. [Google Scholar] [CrossRef] [Green Version]
  78. Wakahara, K.; Ohno, T.; Kimura, M.; Masuda, T.; Nozawa, S.; Dohjima, T.; Yamamoto, T.; Nagano, A.; Kawai, G.; Matsuhashi, A.; et al. EWS-Fli1 Up-Regulates Expression of the Aurora A and Aurora B Kinases. Mol. Cancer Res. 2008, 6, 1937–1945. [Google Scholar] [CrossRef] [Green Version]
  79. Vibert, J.; Saulnier, O.; Collin, C.; Petit, F.; Borgman, K.J.; Vigneau, J.; Gautier, M.; Zaidi, S.; Pierron, G.; Watson, S.; et al. Oncogenic chimeric transcription factors drive tumor-specific transcription, processing, and translation of silent genomic regions. Mol. Cell 2022, 82, 2458–2471.e9. [Google Scholar] [CrossRef]
  80. Marques Howarth, M.; Simpson, D.; Ngok, S.P.; Nieves, B.; Chen, R.; Siprashvili, Z.; Vaka, D.; Breese, M.R.; Crompton, B.D.; Alexe, G.; et al. Long noncoding RNA EWSAT1-mediated gene repression facilitates Ewing sarcoma oncogenesis. J. Clin. Investig. 2014, 124, 5275–5290. [Google Scholar] [CrossRef]
  81. Cook, P.J.; Ju, B.G.; Telese, F.; Wang, X.; Glass, C.K.; Rosenfeld, M.G. Tyrosine dephosphorylation of H2AX modulates apoptosis and survival decisions. Nature 2009, 458, 591–596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Ban, J.; Jug, G.; Mestdagh, P.; Schwentner, R.; Kauer, M.; Aryee, D.N.; Schaefer, K.L.; Nakatani, F.; Scotlandi, K.; Reiter, M.; et al. Hsa-mir-145 is the top EWS-FLI1-repressed microRNA involved in a positive feedback loop in Ewing’s sarcoma. Oncogene 2011, 30, 2173–2180. [Google Scholar] [CrossRef] [Green Version]
  83. Riggi, N.; Suvà, M.-L.; De Vito, C.; Provero, P.; Stehle, J.-C.; Baumer, K.; Cironi, L.; Janiszewska, M.; Petricevic, T.; Suvà, D.; et al. EWS-FLI-1 modulates miRNA145 and SOX2 expression to initiate mesenchymal stem cell reprogramming toward Ewing sarcoma cancer stem cells. Genes Dev. 2010, 24, 916–932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. McKinsey, E.L.; Parrish, J.K.; Irwin, A.E.; Niemeyer, B.F.; Kern, H.B.; Birks, D.K.; Jedlicka, P. A novel oncogenic mechanism in Ewing sarcoma involving IGF pathway targeting by EWS/Fli1-regulated microRNAs. Oncogene 2011, 30, 4910–4920. [Google Scholar] [CrossRef] [Green Version]
  85. France, K.A.; Anderson, J.L.; Park, A.; Denny, C.T. Oncogenic Fusion Protein EWS/FLI1 Down-regulates Gene Expression by Both Transcriptional and Posttranscriptional Mechanisms. J. Biol. Chem. 2011, 286, 22750–22757. [Google Scholar] [CrossRef] [Green Version]
  86. Selvanathan, S.P.; Graham, G.T.; Erkizan, H.V.; Dirksen, U.; Natarajan, T.G.; Dakic, A.; Yu, S.; Liu, X.; Paulsen, M.T.; Ljungman, M.E.; et al. Oncogenic fusion protein EWS-FLI1 is a network hub that regulates alternative splicing. Proc. Natl. Acad. Sci. USA 2015, 112, E1307–E1316. [Google Scholar] [CrossRef] [Green Version]
  87. Gorthi, A.; Romero, J.C.; Loranc, E.; Cao, L.; Lawrence, L.A.; Goodale, E.; Iniguez, A.B.; Bernard, X.; Masamsetti, V.P.; Roston, S.; et al. EWS–FLI1 increases transcription to cause R-loops and block BRCA1 repair in Ewing sarcoma. Nature 2018, 555, 387–391. [Google Scholar] [CrossRef]
  88. Ahmed, N.S.; Harrell, L.M.; Wieland, D.R.; Lay, M.A.; Thompson, V.F.; Schwartz, J.C. Fusion protein EWS-FLI1 is incorporated into a protein granule in cells. RNA 2021, 27, 920–932. [Google Scholar] [CrossRef] [PubMed]
  89. Boulay, G.; Sandoval, G.J.; Riggi, N.; Iyer, S.; Buisson, R.; Naigles, B.; Awad, M.E.; Rengarajan, S.; Volorio, A.; McBride, M.J.; et al. Cancer-Specific Retargeting of BAF Complexes by a Prion-like Domain. Cell 2017, 171, 163–178.e19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Chong, S.; Dugast-Darzacq, C.; Liu, Z.; Dong, P.; Dailey, G.M.; Cattoglio, C.; Heckert, A.; Banala, S.; Lavis, L.; Darzacq, X.; et al. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 2018, 361, eaar2555. [Google Scholar] [CrossRef]
  91. Chong, S.; Graham, T.G.; Dugast-Darzacq, C.; Dailey, G.M.; Darzacq, X.; Tjian, R. Tuning levels of low-complexity domain interactions to modulate endogenous oncogenic transcription. Mol. Cell 2022, 82, 2084–2097.e5. [Google Scholar] [CrossRef] [PubMed]
  92. Schwentner, R.; Papamarkou, T.; Kauer, M.O.; Stathopoulos, V.; Yang, F.; Bilke, S.; Meltzer, P.S.; Girolami, M.; Kovar, H. EWS-FLI1 employs an E2F switch to drive target gene expression. Nucleic Acids Res. 2015, 43, 2780–2789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Shimizu, R.; Tanaka, M.; Tsutsumi, S.; Aburatani, H.; Yamazaki, Y.; Homme, M.; Kitagawa, Y.; Nakamura, T. EWS-FLI 1 regulates a transcriptional program in cooperation with Foxq1 in mouse Ewing sarcoma. Cancer Sci. 2018, 109, 2907–2918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Montoya, C.; Rey, L.; Rodriguez, J.; Fernandez, M.J.; Troncoso, D.; Canas, A.; Moreno, O.; Henriquez, B.; Rojas, A. Epigenetic control of the EWSFLI1 promoter in Ewing’s sarcoma. Oncol. Rep. 2020, 43, 1199–1207. [Google Scholar]
  95. Giorgi, C.; Boro, A.; Rechfeld, F.; Lopez-Garcia, L.A.; Gierisch, M.E.; Schafer, B.W.; Niggli, F.K. PI3K/AKT signaling modulates transcriptional expression of EWS/FLI1 through specificity protein 1. Oncotarget 2015, 6, 28895–28910. [Google Scholar] [CrossRef] [Green Version]
  96. Aryee, D.N.; Niedan, S.; Kauer, M.; Schwentner, R.; Bennani-Baiti, I.M.; Ban, J.; Muehlbacher, K.; Kreppel, M.; Walker, R.L.; Meltzer, P.; et al. Hypoxia modulates EWS-FLI1 transcriptional signature and enhances the malignant properties of Ewing’s sarcoma cells in vitro. Cancer Res. 2010, 70, 4015–4023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Grohar, P.J.; Kim, S.; Rivera, G.O.R.; Sen, N.; Haddock, S.; Harlow, M.L.; Maloney, N.K.; Zhu, J.; O’Neill, M.; Jones, T.L.; et al. Functional Genomic Screening Reveals Splicing of the EWS-FLI1 Fusion Transcript as a Vulnerability in Ewing Sarcoma. Cell Rep. 2016, 14, 598–610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Keskin, T.; Bakaric, A.; Waszyk, P.; Boulay, G.; Torsello, M.; Cornaz-Buros, S.; Chevalier, N.; Geiser, T.; Martin, P.; Volorio, A.; et al. LIN28B Underlies the Pathogenesis of a Subclass of Ewing Sarcoma LIN28B Control of EWS-FLI1 Stability. Cell Rep. 2020, 30, 4567–4583.e5. [Google Scholar] [CrossRef]
  99. Sun, H.; Lin, D.-C.; Cao, Q.; Guo, X.; Marijon, H.; Zhao, Z.; Gery, S.; Xu, L.; Yang, H.; Pang, B.; et al. CRM1 Inhibition Promotes Cytotoxicity in Ewing Sarcoma Cells by Repressing EWS-FLI1–Dependent IGF-1 Signaling. Cancer Res. 2016, 76, 2687–2697. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, M.; Xie, Y.; Girnita, L.; Nilsson, G.; Dricu, A.; Wejde, J.; Larsson, O. Regulatory role of mevalonate and N-linked glycosylation in proliferation and expression of the EWS/FLI-1 fusion protein in Ewing’s sarcoma cells. Exp. Cell Res. 1999, 246, 38–46. [Google Scholar] [CrossRef]
  101. Girnita, L.; Wang, M.; Xie, Y.; Nilsson, G.; Dricu, A.; Wejde, J.; Larsson, O. Inhibition of N-linked glycosylation down-regulates insulin-like growth factor-1 receptor at the cell surface and kills Ewing’s sarcoma cells: Therapeutic implications. Anticancer Drug Des. 2000, 15, 67–72. [Google Scholar]
  102. Klevernic, I.V.; Morton, S.; Davis, R.J.; Cohen, P. Phosphorylation of Ewing’s sarcoma protein (EWS) and EWS-Fli1 in response to DNA damage. Biochem. J. 2009, 418, 625–634. [Google Scholar] [CrossRef]
  103. Bachmaier, R.; Aryee, D.N.; Jug, G.; Kauer, M.; Kreppel, M.; Lee, K.A.; Kovar, H. O-GlcNAcylation is involved in the transcriptional activity of EWS-FLI1 in Ewing’s sarcoma. Oncogene 2009, 28, 1280–1284. [Google Scholar] [CrossRef] [Green Version]
  104. Schlottmann, S.; Erkizan, H.V.; Barber-Rotenberg, J.S.; Knights, C.; Cheema, A.; Üren, A.; Avantaggiati, M.L.; Toretsky, J.A. Acetylation Increases EWS-FLI1 DNA Binding and Transcriptional Activity. Front. Oncol. 2012, 2, 107. [Google Scholar] [CrossRef] [Green Version]
  105. Gierisch, M.E.; Pfistner, F.; Lopez-Garcia, L.A.; Harder, L.; Schäfer, B.W.; Niggli, F.K. Proteasomal Degradation of the EWS-FLI1 Fusion Protein Is Regulated by a Single Lysine Residue. J. Biol. Chem. 2016, 291, 26922–26933. [Google Scholar] [CrossRef] [Green Version]
  106. Elzi, D.J.; Song, M.; Hakala, K.; Weintraub, S.T.; Shiio, Y. Proteomic Analysis of the EWS-Fli-1 Interactome Reveals the Role of the Lysosome in EWS-Fli-1 Turnover. J. Proteome Res. 2014, 13, 3783–3791. [Google Scholar] [CrossRef] [Green Version]
  107. Su, S.; Chen, J.; Jiang, Y.; Wang, Y.; Vital, T.; Zhang, J.; Laggner, C.; Nguyen, K.T.; Zhu, Z.; Prevatte, A.W.; et al. SPOP and OTUD7A Control EWS-FLI1 Protein Stability to Govern Ewing Sarcoma Growth. Adv. Sci. 2021, 8, e2004846. [Google Scholar] [CrossRef]
  108. Gierisch, M.E.; Pedot, G.; Walser, F.; Lopez-Garcia, L.A.; Jaaks, P.; Niggli, F.K.; Schäfer, B.W. USP19 deubiquitinates EWS-FLI1 to regulate Ewing sarcoma growth. Sci. Rep. 2019, 9, 951. [Google Scholar] [CrossRef] [Green Version]
  109. Toretsky, J.A.; Erkizan, V.; Levenson, A.; Abaan, O.D.; Parvin, J.D.; Cripe, T.P.; Rice, A.M.; Lee, S.B.; Üren, A. Oncoprotein EWS-FLI1 Activity Is Enhanced by RNA Helicase A. Cancer Res 2006, 66, 5574–5581. [Google Scholar] [CrossRef] [Green Version]
  110. Brenner, J.C.; Feng, F.Y.; Han, S.; Patel, S.; Goyal, S.V.; Bou-Maroun, L.M.; Liu, M.; Lonigro, R.; Prensner, J.R.; Tomlins, S.A.; et al. PARP-1 inhibition as a targeted strategy to treat Ewing’s sarcoma. Cancer Res. 2012, 72, 1608–1613. [Google Scholar] [CrossRef] [Green Version]
  111. Bertolotti, A.; Melot, T.; Acker, J.; Vigneron, M.; Delattre, O.; Tora, L. EWS, but Not EWS-FLI-1, Is Associated with Both TFIID and RNA Polymerase II: Interactions between Two Members of the TET Family, EWS and hTAF II 68, and Subunits of TFIID and RNA Polymerase II Complexes. Mol. Cell Biol. 1998, 18, 1489–1497. [Google Scholar] [CrossRef] [Green Version]
  112. Watson, D.K.; Robinson, L.; Hodge, D.R.; Kola, I.; Papas, T.S.; Seth, A. FLI1 and EWS-FLI1 function as ternary complex factors and ELK1 and SAP1a function as ternary and quaternary complex factors on the Egr1 promoter serum response elements. Oncogene 1997, 14, 213–221. [Google Scholar] [CrossRef]
  113. Spahn, L.; Petermann, R.; Siligan, C.; Schmid, J.A.; Aryee, D.N.; 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, 4583–4587. [Google Scholar]
  114. Kim, S.; Denny, C.T.; Wisdom, R. Cooperative DNA Binding with AP-1 Proteins Is Required for Transformation by EWS-Ets Fusion Proteins. Mol. Cell. Biol. 2006, 26, 2467–2478. [Google Scholar] [CrossRef] [Green Version]
  115. Srivastava, S.; Nataraj, N.B.; Sekar, A.; Ghosh, S.; Bornstein, C.; Drago-Garcia, D.; Roth, L.; Romaniello, D.; Marrocco, I.; David, E.; et al. ETS Proteins Bind with Glucocorticoid Receptors: Relevance for Treatment of Ewing Sarcoma. Cell Rep. 2019, 29, 104–117.e4. [Google Scholar] [CrossRef] [Green Version]
  116. Vasileva, E.; Warren, M.; Triche, T.J.; Amatruda, J.F. Dysregulated heparan sulfate proteoglycan metabolism promotes Ewing sarcoma tumor growth. Elife 2022, 11, e69734. [Google Scholar] [CrossRef]
  117. Molnar, C.; Reina, J.; Herrero, A.; Heinen, J.P.; Méndiz, V.; Bonnal, S.; Irimia, M.; Sánchez-Jiménez, M.; Sánchez-Molina, S.; Mora, J.; et al. Human EWS-FLI protein recapitulates in Drosophila the neomorphic functions that induce Ewing sarcoma tumorigenesis. Proc. Natl. Acad. Sci. Nexus 2022, 1, pgac222. [Google Scholar] [CrossRef]
  118. Jedlicka, P. Ewing Sarcoma, an enigmatic malignancy of likely progenitor cell origin, driven by transcription factor oncogenic fusions. Int. J. Clin. Exp. Pathol. 2010, 3, 338–347. [Google Scholar]
  119. Toomey, E.C.; Schiffman, J.D.; Lessnick, S.L. Recent advances in the molecular pathogenesis of Ewing’s sarcoma. Oncogene 2010, 29, 4504–4516. [Google Scholar] [CrossRef] [Green Version]
  120. Mackintosh, C.; Madoz-Gurpide, J.; Ordonez, J.L.; Osuna, D.; Herrero-Martin, D. The molecular pathogenesis of Ewing’s sarcoma. Cancer Biol. Ther. 2010, 9, 655–667. [Google Scholar] [CrossRef]
  121. Tirode, F.; Surdez, D.; Ma, X.; Parker, M.; Le Deley, M.C.; Bahrami, A.; Zhang, Z.; Lapouble, E.; Grossetête-Lalami, S.; Rusch, M.; et al. Genomic Landscape of Ewing Sarcoma Defines an Aggressive Subtype with Co-Association of STAG2 and TP53 Mutations. Cancer Discov. 2014, 4, 1342–1353. [Google Scholar] [CrossRef] [Green Version]
  122. Aynaud, M.-M.; Mirabeau, O.; Gruel, N.; Grossetête, S.; Boeva, V.; Durand, S.; Surdez, D.; Saulnier, O.; Zaïdi, S.; Gribkova, S.; et al. Transcriptional Programs Define Intratumoral Heterogeneity of Ewing Sarcoma at Single-Cell Resolution. Cell Rep. 2020, 30, 1767–1779.e6. [Google Scholar] [CrossRef] [Green Version]
  123. Jagodzińska-Mucha, P.; Sobczuk, P.; Mikuła, M.; Raciborska, A.; Dawidowska, A.; Kulecka, M.; Bilska, K.; Szumera-Ciećkiewicz, A.; Kluska, A.; Piątkowska, M.; et al. Mutational landscape of primary and recurrent Ewing sarcoma. Contemp. Oncol. 2021, 25, 241–248. [Google Scholar] [CrossRef]
  124. Brohl, A.S.; Ms, R.P.; Turner, C.E.; Wen, X.; Song, Y.K.; Wei, J.S.; Calzone, K.A.; Khan, J. Frequent inactivating germline mutations in DNA repair genes in patients with Ewing sarcoma. Anesth. Analg. 2017, 19, 955–958. [Google Scholar] [CrossRef]
  125. Crompton, B.; Stewart, C.; Taylor-Weiner, A.; Alexa, G.; Kurek, K.; Calicchio, M.; Kiezun, A.; Carter, S.; Shukla, S.; Mehta, S.; et al. Abstract 999: The genomic landscape of pediatric Ewing sarcoma. Cancer Discov. 2014, 4, 1326–1341. [Google Scholar] [CrossRef] [Green Version]
  126. Gillani, R.; Camp, S.Y.; Han, S.; Jones, J.K.; Chu, H.; O’Brien, S.; Young, E.L.; Hayes, L.; Mitchell, G.; Fowler, T.; et al. Germline predisposition to pediatric Ewing sarcoma is characterized by inherited pathogenic variants in DNA damage repair genes. Am. J. Hum. Genet. 2022, 109, 1026–1037. [Google Scholar] [CrossRef]
  127. Stoll, G.; Surdez, D.; Tirode, F.; Laud, K.; Barillot, E.; Zinovyev, A.; Delattre, O. Systems biology of Ewing sarcoma: A network model of EWS-FLI1 effect on proliferation and apoptosis. Nucleic Acids Res. 2013, 41, 8853–8871. [Google Scholar] [CrossRef] [Green Version]
  128. Sohn, E.J.; Li, H.; Reidy, K.; Beers, L.F.; Christensen, B.L.; Lee, S.B. EWS/FLI1 Oncogene Activates Caspase 3 Transcription and Triggers Apoptosis In vivo. Cancer Res. 2010, 70, 1154–1163. [Google Scholar] [CrossRef] [Green Version]
  129. van der Ent, W.; Jochemsen, A.G.; Teunisse, A.F.; Krens, S.F.; Szuhai, K.; Spaink, H.P.; Hogendoorn, P.C.; Snaar-Jagalska, B.E. Ewing sarcoma inhibition by disruption of EWSR1-FLI1 transcriptional activity and reactivation of p53. J. Pathol. 2014, 233, 415–424. [Google Scholar] [CrossRef]
  130. Xiao, J.; Glasgow, E.; Agarwal, S. Zebrafish Xenografts for Drug Discovery and Personalized Medicine. Trends Cancer 2020, 6, 569–579. [Google Scholar] [CrossRef]
  131. Okada, S.; Vaeteewoottacharn, K.; Kariya, R. Application of Highly Immunocompromised Mice for the Establishment of Patient-Derived Xenograft (PDX) Models. Cells 2019, 8, 889. [Google Scholar] [CrossRef] [Green Version]
  132. Wan, L.; Neumann, C.A.; LeDuc, P.R. Tumor-on-a-chip for integrating a 3D tumor microenvironment: Chemical and mechanical factors. Lab Chip 2020, 20, 873–888. [Google Scholar] [CrossRef]
  133. Hoffman, R.M. Patient-derived orthotopic xenografts: Better mimic of metastasis than subcutaneous xenografts. Nat. Rev. Cancer 2015, 15, 451–452. [Google Scholar] [CrossRef]
  134. Stewart, E.; Federico, S.M.; Chen, X.; Shelat, A.A.; Bradley, C.; Gordon, B.; Karlstrom, A.; Twarog, N.R.; Clay, M.R.; Bahrami, A.; et al. Orthotopic patient-derived xenografts of paediatric solid tumours. Nature 2017, 549, 96–100. [Google Scholar] [CrossRef]
Figure 1. Regulatory mechanisms controlling EWSR1::FLI1. Proper EWSR1::FLI1 protein expression is controlled by multi-layer mechanisms, including regulation of transcription by epigenetic regulations, transcription factors and miRNA/lncRNAs, translational regulations, post-translational regulation and various binding proteins. This figure is generated using BioRender.
Figure 1. Regulatory mechanisms controlling EWSR1::FLI1. Proper EWSR1::FLI1 protein expression is controlled by multi-layer mechanisms, including regulation of transcription by epigenetic regulations, transcription factors and miRNA/lncRNAs, translational regulations, post-translational regulation and various binding proteins. This figure is generated using BioRender.
Cancers 15 00382 g001
Figure 2. A summary of EWSR1-FLI1 post-translational modifications. Function of EWSR1-FLI1 proteins are regulated by various post-translational modifications including mono-ubiquitination, poly-ubiquitination, phosphorylation, acetylation and O-GlcNAcylation in cells as indicated here.
Figure 2. A summary of EWSR1-FLI1 post-translational modifications. Function of EWSR1-FLI1 proteins are regulated by various post-translational modifications including mono-ubiquitination, poly-ubiquitination, phosphorylation, acetylation and O-GlcNAcylation in cells as indicated here.
Cancers 15 00382 g002
Table 1. A summary of FET–ETS fusion oncogenes in Ewing sarcoma.
Table 1. A summary of FET–ETS fusion oncogenes in Ewing sarcoma.
FETETSFusion GeneFrequencyTranslocation
EWSR1FLI1EWSR1-FLI185%t(11; 22)(q24; q12)
EWSR1ERGEWSR1-ERG10%t(21; 12)(q22; q12)
EWSR1FEVEWSR1-FEV<1%t(2; 22)(q33; q12)
EWSR1ETV1EWSR1-ETV1<1%t(7; 22)(p22; q12)
EWSR1E1AFEWSR1-E1AF<1%t(17; 22)(q21; q12
FUSFEVFUS-FEV<1%t(2; 16)(q35; p11)
FUSERGFUS-ERG<1%t(16; 21)(p11; q22)
ETSFETFusion geneFrequencyTranslocation
FLI1EWSR1FLI1-EWSR1TBDt(22; 11)(q12; q24)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yu, L.; Davis, I.J.; Liu, P. Regulation of EWSR1-FLI1 Function by Post-Transcriptional and Post-Translational Modifications. Cancers 2023, 15, 382. https://doi.org/10.3390/cancers15020382

AMA Style

Yu L, Davis IJ, Liu P. Regulation of EWSR1-FLI1 Function by Post-Transcriptional and Post-Translational Modifications. Cancers. 2023; 15(2):382. https://doi.org/10.3390/cancers15020382

Chicago/Turabian Style

Yu, Le, Ian J. Davis, and Pengda Liu. 2023. "Regulation of EWSR1-FLI1 Function by Post-Transcriptional and Post-Translational Modifications" Cancers 15, no. 2: 382. https://doi.org/10.3390/cancers15020382

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop